You are on page 1of 371

Development and Demonstration of a New Non-

Equilibrium Rate-Based Process Model for the


Hot Potassium Carbonate Process

by

Su Ming Pamela Ooi

Thesis submitted for the degree of


Doctor of Philosophy

in

The University of Adelaide


School of Chemical Engineering
July 2008
Summary

Summary
Chemical absorption and desorption processes are two fundamental operations in the process
industry. Due to the rate-controlled nature of these processes, classical equilibrium stage models are
usually inadequate for describing the behaviour of chemical absorption and desorption processes. A
more effective modelling method is the non-equilibrium rate-based approach, which considers the
effects of the various driving forces across the vapour-liquid interface.

In this thesis, a new non-equilibrium rate-based model for chemical absorption and desorption is
developed and applied to the hot potassium carbonate process CO2 Removal Trains at the Santos
Moomba Processing Facility. The rate-based process models incorporate rigorous thermodynamic
and mass transfer relations for the system and detailed hydrodynamic calculations for the column
internals. The enhancement factor approach was used to represent the effects of the chemical
reactions.

The non-equilibrium rate-based CO2 Removal Train process models were implemented in the Aspen
Custom Modeler® simulation environment, which enabled rigorous thermodynamic and physical
property calculations via the Aspen Properties® software. Literature data were used to determine the
parameters for the Aspen Properties® property models and to develop empirical correlations when the
default Aspen Properties® models were inadequate. Preliminary simulations indicated the need for
adjustments to the absorber column models, and a sensitivity analysis identified the effective
interfacial area as a suitable model parameter for adjustment. Following the application of adjustment
factors to the absorber column models, the CO2 Removal Train process models were successfully
validated against steady-state plant data.

The success of the Aspen Custom Modeler® process models demonstrated the suitability of the non-
equilibrium rate-based approach for modelling the hot potassium carbonate process. Unfortunately,
the hot potassium carbonate process could not be modelled as such in HYSYS®, Santos’s preferred
simulation environment, due to the absence of electrolyte components and property models and the
limitations of the HYSYS® column operations in accommodating chemical reactions and non-
equilibrium column behaviour. While importation of the Aspen Custom Modeler® process models into
HYSYS® was possible, it was considered impractical due to the significant associated computation
time.

To overcome this problem, a novel approach involving the HYSYS® column stage efficiencies and
hypothetical HYSYS® components was developed. Stage efficiency correlations, relating various
operating parameters to the column performance, were derived from parametric studies performed in
Aspen Custom Modeler®. Preliminary simulations indicated that the efficiency correlations were only
necessary for the absorber columns; the regenerator columns were adequately represented by the
default equilibrium stage models. Hypothetical components were created for the hot potassium
carbonate system and the standard Peng-Robinson property package model in HYSYS® was

i
Summary

modified to include tabular physical property models to accommodate the hot potassium carbonate
system. Relevant model parameters were determined from literature data. As for the Aspen Custom
Modeler® process models, the HYSYS® CO2 Removal Train process models were successfully
validated against steady-state plant data.

To demonstrate a potential application of the HYSYS® process models, dynamic simulations of the
two most dissimilarly configured trains, CO2 Removal Trains #1 and #7, were performed. Simple first-
order plus dead time (FOPDT) process transfer function models, relating the key process variables,
were derived to develop a diagonal control structure for each CO2 Removal Train. The FOPDT model
is the standard process engineering approximation to higher order systems, and it effectively
described most of the process response curves for the two CO2 Removal Trains. Although a few
response curves were distinctly underdamped, the quality of the validating data for the CO2 Removal
Trains did not justify the use of more complex models than the FOPDT model.

While diagonal control structures are a well established form of control for multivariable systems, their
application to the hot potassium carbonate process has not been documented in literature. Using a
number of controllability analysis methods, the two CO2 Removal Trains were found to share the same
optimal diagonal control structure, which suggested that the identified control scheme was
independent of the CO2 Removal Train configurations. The optimal diagonal control structure was
tested in dynamic simulations using the MATLAB® numerical computing environment and was found
to provide effective control. This finding confirmed the results of the controllability analyses and
demonstrated how the HYSYS® process model could be used to facilitate the development of a
control strategy for the Moomba CO2 Removal Trains.

In conclusion, this work addressed the development of a new non-equilibrium rate-based model for the
hot potassium carbonate process and its application to the Moomba CO2 Removal Trains. Further
work is recommended to extend the model validity over a wider range of operating conditions and to
expand the dynamic HYSYS® simulations to incorporate the diagonal control structures and/or more
complex control schemes.

ii
Statement

Statement
This work contains no material which has been accepted for the award of any other degree or diploma
in any university or other tertiary institution and, to the best of my knowledge and belief, contains no
material previously published or written by another person, except where due reference has been
made in the text.

I give consent to this copy of my thesis, when deposited in the University Library, being made
available for loan and photocopying, subject to the provisions of the Copyright Act 1968.

Su Ming Pamela Ooi


Adelaide, 16/07/2008

iii
Acknowledgements

Acknowledgements
I wish to express my sincere appreciation to all organisations and persons who have contributed
council and assistance during this work.

In particular, I would like to thank Santos for providing funding and technical support for this project.
Special thanks are extended to Keith Humphris, Len Cowen, Ian Smith, Ross Mullner, Claire Barber,
Randall Yeates and Mark Moss for their advice and assistance with providing technical data.

I would also like to thank my academic supervisors, Associate Professor Brian O’Neill, Dr Chris Colby
and Professor Keith King, for their encouragement, advice and guidance. Special thanks are
extended to Dr Robin Thiele who provided invaluable advice regarding rate-based modelling.

Finally, I would like to express my deepest gratitude to my boyfriend Simon and to my parents for
providing endless support and encouragement, without which I could not have completed this thesis.

iv
Table of Contents

Table of Contents
Nomenclature …........... xvii

Chapter 1: Introduction …........... 1


1.1 Project Objectives …........... 2
1.2 Thesis Structure …........... 2

Chapter 2: Literature Review …........... 4


2.1 The Hot Potassium Carbonate Process …........... 5
2.2 The Non-Equilibrium Rate-Based Approach …........... 18
2.3 Electrolyte Thermodynamics …........... 30
2.4 Process Simulation Platform …........... 32
2.5 Multivariable Process Control ….......... 34
2.6 Summary ….......... 45

Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium


Carbonate System ….......... 46
3.1 Summary of Property Models ….......... 47
3.2 The Electrolyte NRTL Model ….......... 50
3.3 Summary ….......... 58

Chapter 4: Aspen Custom Modeler® Process Model Development ….......... 59


4.1 Process Model Equations ….......... 60
4.2 Preliminary CO2 Train Simulations ….......... 69
4.3 Column Model Adjustments ….......... 73
4.4 CO2 Train Model Validation ….......... 83
4.5 Summary ….......... 91

Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies ….......... 92
5.1 Solution Operating Parameters ….......... 93
5.2 Raw Gas Operating Parameters ….......... 97
5.3 Column Operating Parameters ….......... 100
5.4 Summary ….......... 107

Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS® ….......... 108
6.1 The Modelling Approach ….......... 109
6.2 Thermodynamic and Physical Property Models ….......... 115
6.3 Summary ….......... 120

Chapter 7: The Absorber and Regenerator Column Models ….......... 121


7.1 Absorber Column Models ….......... 122
7.2 Regenerator Column Models ….......... 124
7.3 Preliminary Column Model Simulations ….......... 127
7.4 Column Stage Efficiency Correlations ….......... 135

v
Table of Contents

7.5 Summary ….......... 141

Chapter 8: HYSYS® CO2 Removal Train Process Model Development ….......... 142
8.1 Ancillary Operation Models ….......... 143
8.2 Steady-State CO2 Train Models ….......... 147
8.3 Model Validation ….......... 151
8.4 Summary ….......... 158

Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7 ….......... 159
9.1 Dynamic CO2 Train Models ….......... 160
9.2 Process Case Studies ….......... 169
9.3 Summary ….......... 175

Chapter 10: Process Control Studies for CO2 Trains #1 and #7 ….......... 176
10.1 Selection of Diagonal Control Structure ….......... 177
10.2 Selection of Diagonal Control Structure Configuration ….......... 184
10.3 Analysis of Diagonal Control Structure Performance ….......... 187
10.4 BLT Tuning ….......... 190
10.5 Diagonal Control Structure Dynamic Simulations ….......... 192
10.6 Summary ….......... 198

Chapter 11: Conclusions and Recommendations …......... 199


11.1 Conclusions …......... 200
11.2 Recommendations …......... 203

References …......... 204

Appendix A: Thermodynamic Model Equations ….......... A-1


A.1 Reference States ….......... A-2
A.2 The Electrolyte NRTL Model ….......... A-3
A.3 Cubic Equations of State ….......... A-8

Appendix B: Property Models for Aspen Custom Modeler® ….......... A-10


B.1 Thermodynamic Property Models ….......... A-11
B.2 Physical and Transport Property Models ….......... A-17

Appendix C: Electrolyte NRTL Adjustable Parameters ….......... A-38


C.1 Parameter Values ….......... A-39
C.2 Data Regression Procedure ….......... A-41
C.3 Data Regression Results ….......... A-44

Appendix D: Aspen Custom Modeler® Simulation Results ….......... A-49


D.1 The Different Modelling Approaches ….......... A-50
D.2 Model Adjustments ….......... A-52
D.3 CO2 Train Model Validation ….......... A-56

Appendix E: Hypothetical K2CO3* HYSYS® Component Properties ….......... A-62

vi
Table of Contents

E.1 Base Properties ….......... A-63


E.2 Additional Point Properties ….......... A-66
E.3 Temperature Dependent Properties ….......... A-67

Appendix F: Property Models for HYSYS® ….......... A-70


F.1 Thermodynamic Property Models ….......... A-71
F.2 Physical and Transport Property Models ….......... A-72

Appendix G: Enhanced PR Binary Interaction Parameters ….......... A-84


G.1 Data Regression Procedure ….......... A-85
G.2 Data Regression Results ….......... A-87

Appendix H: HYSYS® Simulation Results ….......... A-88


H.1 Preliminary Column Model Simulations ….......... A-89
H.2 Steady-State CO2 Train Models ….......... A-95
H.3 CO2 Train Model Validation ….......... A-101

Appendix I: Process Control Studies of the CO2 Trains ….......... A-107


I.1 Selection of Diagonal Control Structure ….......... A-108
I.2 Selection of Diagonal Control Structure Configuration ….......... A-111
I.3 Analysis of Diagonal Control Structure Performance ….......... A-114
I.4 BLT Tuning ….......... A-115
I.5 Diagonal Control Structure Dynamic Simulations ….......... A-117

vii
List of Figures

List of Figures
Figure 2.1.1: A simple form of the hot potassium carbonate process. ….......... 5
Figure 2.1.2: Basic CO2 train process flow diagrams. ….......... 7
Figure 2.1.3: Reaction flow scheme for the hot potassium carbonate process. ….......... 10
Figure 2.2.1: The two-film model for simultaneous mass and energy transfer. ….......... 20
Figure 3.2.1: Comparison between the Electrolyte NRTL predictions and the
experimental data. ….......... 54
Figure 3.2.2: CO2 partial pressure over K2CO3 solution as a function of CO2
loading. ….......... 54
Figure 3.2.3: Comparison between the Electrolyte NRTL predictions and the
experimental data. ….......... 57
Figure 3.2.4: H2S partial pressure over K2CO3 solution as a function of
equivalent H2S content. ….......... 57
Figure 4.1.1: Equilibrium stage for Model 1 (adapted from Thiele (2007)). ….......... 62
Figure 4.1.2: Non-equilibrium stage for Model 2 (adapted from Thiele (2007)). ….......... 63
Figure 4.1.3: Non-equilibrium stage for Model 3 (adapted from Thiele (2007)). ….......... 64
Figure 4.2.1: Preliminary Aspen Custom Modeler® simulation column
configurations. ….......... 69
Figure 4.2.2: Results of the absorber discretisation simulation runs for CO2
train #1. ….......... 70
Figure 4.2.3: Results of the regenerator discretisation simulation runs for CO2
train #1. ….......... 70
Figure 4.2.4: Results of the different modelling approaches for CO2 trains #1
and #7. ….......... 72
Figure 4.3.1: Temperature profiles for CO2 trains #1 and #7. ….......... 74
Figure 4.3.2: Effect of the liquid phase enthalpy correction on the absorber
profiles. ….......... 75
Figure 4.3.3: Sensitivity analysis results for the absorber column model
(Model 2). ….......... 79
Figure 4.3.4: Sensitivity analysis results for the regenerator column model
(Model 2). ….......... 80
Figure 4.3.5: Effect of the solution reboiler steam flow on the regenerator
column model (Model 2). ….......... 81
Figure 4.3.6: Effect of the effective interfacial area adjustment factor on the
absorber CO2 and H2S vapour phase profiles. ….......... 82
Figure 4.4.1: Simplified CO2 train configurations for the Aspen Custom
Modeler® simulations. ….......... 83
Figure 4.4.2: CO2 and H2S vapour and liquid phase profiles for the first set of
plant data. ….......... 88
Figure 4.4.3: CO2 and H2S vapour and liquid phase profiles for the second set
of plant data. ….......... 87
Figure 4.4.4: Vapour and liquid phase temperature profiles for the two sets of
plant data. ….......... 88
Figure 5.1.1: Effect of the solution flow rate on the performance of the CO2
trains. ….......... 94

viii
List of Figures

Figure 5.1.2: Effect of the solution strength on the performance of the CO2
trains. ….......... 95
Figure 5.1.3: Effect of the solution CO2 loading on the performance of the CO2
trains. ….......... 96
Figure 5.2.1: Effect of the raw gas flow rate on the performance of the CO2
trains. ….......... 98
Figure 5.2.2: Effect of the raw gas CO2 content on the performance of the CO2
trains. ….......... 99
Figure 5.3.1: Effect of the absorber temperature on the performance of the
CO2 trains. ….......... 101
Figure 5.3.2: Effect of the regenerator temperature on the performance of the
CO2 trains. ….......... 102
Figure 5.3.3: Effect of pressure on the performance of the CO2 trains. ….......... 104
Figure 5.3.4: Effect of the steam flow rate to the regenerator solution reboilers
on the performance of the CO2 trains. ….......... 105
Figure 5.3.5: Effect of the makeup water flow rate to the regenerator overhead
catchpots on the performance of the CO2 trains. ….......... 106
Figure 6.1.1: The different definitions of the column stage efficiency  in
HYSYS®. ….......... 114
Figure 6.2.1: Comparison between the enhanced PR predictions and the
experimental data. ….......... 118
Figure 6.2.2: Comparison between the enhanced PR predictions and the
experimental data. ….......... 119
Figure 7.1.1: Process flow diagram of an absorber column model in HYSYS®. ….......... 122
Figure 7.2.1: Process flow diagrams of the two most dissimilar regenerator
column models in HYSYS®. ….......... 126
Figure 7.3.1: Equilibrium stage simulation results for the absorber and
regenerator columns. ….......... 128
Figure 7.3.2: Sensitivity analysis results for the absorber models. ….......... 130
Figure 7.3.3: Effect of the column stage efficiencies on the absorber
composition and temperature profiles. ….......... 131
Figure 7.3.4: Sensitivity analysis results for the regenerator models. ….......... 132
Figure 7.3.5: Effect of the column stage efficiencies on the regenerator
composition and temperature profiles. ….......... 133
Figure 7.3.6: Effect of the reboiler steam flow on the regenerator column
performance. ….......... 134
Figure 7.4.1: Effect of the correlated overall stage efficiencies on the steady-
state absorber columns. ….......... 137
Figure 7.4.2: Effect of the correlated overall stage efficiencies on the steady-
state behaviour of the dynamic absorber columns. ….......... 137
Figure 7.4.3: An example HYSYS® spreadsheet for calculating the absorber
overall stage efficiencies. ….......... 139
Figure 8.1.1: HYSYS® process flow diagram of the CO2 train absorption
circuits. ….......... 143
Figure 8.1.2: HYSYS® process flow diagram of the solution pumpset model. ….......... 144
Figure 8.1.3: HYSYS® spreadsheet for the pumpset calculations. ….......... 145

ix
List of Figures

Figure 8.2.1: Process flow diagram for the steady-state HYSYS® model of
CO2 train #1. ….......... 149
Figure 8.2.2: Process flow diagram for the steady-state HYSYS® model of
CO2 train #7. ….......... 150
Figure 8.3.1: CO2 and H2S vapour and liquid phase profiles for the first set of
data. ….......... 153
Figure 8.3.2: CO2 and H2S vapour and liquid phase profiles for the second set
of data. ….......... 154
Figure 8.3.3: Vapour and liquid phase temperature profiles for the two sets of
plant data. ….......... 155
Figure 9.1.1: Process flow diagram of the simplified dynamic HYSYS® model
for CO2 train #1. ….......... 163
Figure 9.1.2: Process flow diagram of the simplified dynamic HYSYS® model
for CO2 train #7. ….......... 164
Figure 9.1.3: Flow control loop responses. ….......... 166
Figure 9.1.4: Temperature control loop responses. ….......... 167
Figure 9.1.5: Liquid level control loop responses. ….......... 168
Figure 9.2.1: Process response curves for a 2% magnitude step change in the
raw gas flow rate at 0 min. ….......... 171
Figure 9.2.2: Process response curves for a 2% magnitude step change in the
lean solution flow rate at 0 min. ….......... 171
Figure 9.2.3: Process response curves for a 2% magnitude step change in the
reboiler steam flow rate at 0 min. ….......... 172
Figure 9.2.4: Process response curves for a 2% magnitude step change in the
regenerator liquid level at 0 min. ….......... 172
Figure 9.2.5: Process response curves for a 2% magnitude step change in the
raw gas CO2 content at 0 min. ….......... 173
Figure 10.1.1: Frequency plots of the MRI and CN. ….......... 181
Figure 10.1.2: Frequency plots of DCN and DC for CO2 train #1. ….......... 182
Figure 10.1.3: Frequency plots of DCN and DC for CO2 train #7. ….......... 183
Figure 10.2.1: Frequency plots for the RGA elements. ….......... 185
Figure 10.3.1: Frequency plots for |PRGAij| and |CLDGij|. ….......... 189
Figure 10.4.1: Plots of the scalar function W. ….......... 191
Figure 10.5.1: Frequency plots of |1+GOL,i(s)| and |PRGAij|. ….......... 195
Figure 10.5.2: CO2 train #1 closed-loop step response curves at the high gas
throughput conditions. ….......... 196
Figure 10.5.3: CO2 train #7 closed-loop step response curves at the high gas
throughput conditions. ….......... 197
Figure B.1.1: Comparison between the predicted and experimental solution
heat capacities. ….......... A-16
Figure B.2.1: Comparison between the predicted and experimental solution
mass densities. ….......... A-20
Figure B.2.2: Comparison between the predicted and experimental solution
viscosities. ….......... A-25
Figure B.2.3: Comparison between the predicted and experimental solution
surface tensions. ….......... A-28

x
List of Figures

Figure B.2.4: Comparison between the predicted and experimental solution


thermal conductivities. ….......... A-33
Figure D.1.1: Results of the different modelling approaches for CO2 trains #2
to #4. ….......... A-50
Figure D.1.2: Results of the different modelling approaches for CO2 trains #5
and #6. ….......... A-51
Figure D.2.1: Temperature profiles for CO2 trains #2 to #4. ….......... A-52
Figure D.2.2: Temperature profiles for CO2 trains #5 and #6. ….......... A-53
Figure D.2.3: Effect of the effective interfacial area adjustment factor on the
absorber CO2 and H2S vapour phase profiles. ….......... A-55
Figure D.3.1: CO2 and H2S vapour and liquid phase column profiles for the
first set of plant data. ….......... A-56
Figure D.3.2: CO2 and H2S vapour and liquid phase column profiles for the
first set of plant data. ….......... A-57
Figure D.3.3: CO2 and H2S vapour and liquid phase column profiles for the
second set of plant data. ….......... A-58
Figure D.3.4: CO2 and H2S vapour and liquid phase column profiles for the
second set of plant data. ….......... A-59
Figure D.3.5: Column temperature profiles for the first set of plant data. ….......... A-60
Figure D.3.6: Column temperature profiles for the second set of plant data. ….......... A-61
Figure F.2.1: Comparison between the predicted and experimental solution
mass densities. ….......... A-74
Figure F.2.2: Comparison between the predicted and experimental solution
viscosities. ….......... A-77
Figure F.2.3: Comparison between the solution surface tensions predicted by
the tabular model and the empirical correlation. ….......... A-79
Figure F.2.4: Comparison between the solution thermal conductivities
predicted by the tabular model and the empirical correlation. ….......... A-83
Figure H.1.1: Equilibrium stage simulation results for the absorber and
regenerator columns. ….......... A-89
Figure H.1.2: Equilibrium stage simulation results for the absorber and
regenerator columns. ….......... A-90
Figure H.1.3: Effect of the correlated overall stage efficiencies on the steady-
state absorber columns. ….......... A-91
Figure H.1.4: Effect of the correlated overall stage efficiencies on the steady-
state absorber columns. ….......... A-92
Figure H.1.5: Effect of the correlated overall stage efficiencies on the steady-
state behaviour of the dynamic absorber columns. ….......... A-93
Figure H.1.6: Effect of the correlated overall stage efficiencies on the steady-
state behaviour of the dynamic absorber columns. ….......... A-94
Figure H.2.1: Process flow diagram for the steady-state model of CO2 train #2. ….......... A-96
Figure H.2.2: Process flow diagram for the steady-state model of CO2 train #3. ….......... A-97
Figure H.2.3: Process flow diagram for the steady-state model of CO2 train #4. ….......... A-98
Figure H.2.4: Process flow diagram for the steady-state model of CO2 train #5. ….......... A-99
Figure H.2.5: Process flow diagram for the steady-state model of CO2 train #6. ….......... A-100

xi
List of Figures

Figure H.3.1: CO2 and H2S vapour and liquid phase column profiles for the
first set of plant data. ….......... A-101
Figure H.3.2: CO2 and H2S vapour and liquid phase column profiles for the
first set of plant data. ….......... A-102
Figure H.3.3: CO2 and H2S vapour and liquid phase column profiles for the
second set of plant data. ….......... A-103
Figure H.3.4: CO2 and H2S vapour and liquid phase column profiles for the
second set of plant data. ….......... A-104
Figure H.3.5: Column temperature profiles for the first set of plant data. ….......... A-105
Figure H.3.6: Column temperature profiles for the second set of plant data. ….......... A-106
Figure I.5.1: CO2 train #1 closed-loop step response curves at the medium
gas throughput conditions. ….......... A-121
Figure I.5.2: CO2 train #7 closed-loop step response curves at the medium
gas throughput conditions. ….......... A-122
Figure I.5.3: CO2 train #1 closed-loop step response curves at the low gas
throughput conditions. ….......... A-123
Figure I.5.4: CO2 train #7 closed-loop step response curves at the low gas
throughput conditions. ….......... A-124

xii
List of Tables

List of Tables
Table 2.1.1: Nameplate capacity of the CO2 trains. ….......... 8
Table 2.1.2: Typical operating data for the CO2 trains from 2002. ….......... 9
Table 2.1.3: Acid gas absorption reactions in the hot potassium carbonate
process. ….......... 10
Table 2.1.4: Ion contribution factors (Pohorecki and Moniuk, 1988). ….......... 13
Table 2.1.5: Acid gas desorption reactions in the hot potassium carbonate
process. ….......... 13
Table 2.1.6: CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system equilibria.
….......... 15
Table 2.1.7: Temperature dependence of the equilibrium and Henry’s Law
constants. ….......... 16
Table 2.1.8: Liquid phase relations and vapour-liquid equilibria expressions. ….......... 16
Table 2.2.1: The MESH equations for a stage i and j = 1…NC components. ….......... 19
Table 2.2.2: Mass transfer relations (Taylor and Krishna, 1993). ….......... 21
Table 2.2.3: Mass transfer coefficient and effective interfacial area
correlations (Onda et al., 1968ab). ….......... 27
Table 2.2.4: Hydrodynamic relations (Stichlmair et al., 1989). ….......... 29
Table 2.2.5: Packing characteristics and constants for metal random
packings. ….......... 29
Table 2.5.1: Common dynamic process behaviour (Stephanopoulos, 1984;
Wade, 2004). ….......... 35
Table 2.5.2: Ziegler-Nichols and Tyreus-Luyben controller tuning rules. ….......... 37
Table 2.5.3: Liquid level PID controller tuning rules (Wade, 2004). ….......... 37
Table 3.1.1: Vapour phase thermodynamic and physical property models. ….......... 47
Table 3.1.2: Liquid phase thermodynamic and physical property models. ….......... 48
Table 3.1.3: Property data sources for the hot potassium carbonate system. ….......... 49
Table 3.2.1: Adjustable binary parameters for the Electrolyte NRTL model. ….......... 50
Table 3.2.2: Electrolyte NRTL parameters for the CO2-K2CO3-KHCO3-H2O
system. ….......... 53
Table 3.2.3: Electrolyte NRTL parameters for the CO2-H2S-K2CO3-KHCO3-
KHS-K2S-H2O system. ….......... 56
Table 4.3.1: The average variation associated with the model parameter
values for the CO2 train absorbers and regenerators. ….......... 77
Table 4.3.2: Effective interfacial area adjustment factor values and their effect
on the CO2 train absorbers. ….......... 82
Table 4.4.1: CO2 train simulation results for the first set of plant data in Table
2.1.2. ….......... 89
Table 4.4.2: CO2 train simulation results for the second set of plant data in
Table 2.1.2. ….......... 90
Table 6.1.1: Property estimation methods for the Miscellaneous class of
hypothetical components. ….......... 110
Table 6.2.1: Thermodynamic and physical property models. ….......... 115

xiii
List of Tables

Table 6.2.2: Enhanced PR parameter values for the CO2-K2CO3-H2O system. ….......... 117
Table 6.2.3: Enhanced PR parameter values for the CO2-H2S-K2CO3-H2O
system. ….......... 119
Table 7.4.1: Coefficients for the steady-state column stage efficiency
correlations. ….......... 136
Table 7.4.2: Coefficients for the dynamic column stage efficiency
correlations. ….......... 136
Table 7.4.3: An example HYSYS® macro for updating the absorber overall
stage efficiencies. ….......... 140
Table 8.3.1: CO2 train simulation results for the first set of plant data in Table
2.1.2. ….......... 156
Table 8.3.2: CO2 train simulation results for the second set of plant data in
Table 2.1.2. ….......... 157
Table 9.1.1: Flow control loop characteristics and controller settings for CO2
trains #1 and #7. ….......... 166
Table 9.1.2: Temperature control loop characteristics and controller settings
for CO2 trains #1 and #7. ….......... 167
Table 9.1.3: Liquid level controller settings for CO2 trains #1 and #7. ….......... 167
Table 9.2.1: Process transfer functions for CO2 train #1. ….......... 173
Table 9.2.2: Process transfer functions for CO2 train #7. ….......... 174
Table 10.1.1: Process transfer function matrices for the diagonal control
structures. ….......... 179
Table 10.1.2: Sensitivity analysis indices at steady-state. ….......... 180
Table 10.2.1: Steady-state results for the interaction and stability analyses for
the RGF-RSF control structure. ….......... 185
Table 10.2.2: Reordered process transfer function matrices for the RGF-RSF
diagonal control structure. ….......... 186
Table 10.3.1: Steady-state values for the PRGA and CLDG for the selected
configuration for the RGF-RSF diagonal control structure. ….......... 188
Table 10.4.1: BLT tuning parameters for the RGF-RSF diagonal control
structures for CO2 trains #1 and #7. ….......... 191
Table B.1.1: Component critical properties (Poling et al., 2001). ….......... A-11
Table B.1.2: Ideal gas heat capacity coefficients and enthalpies of formation
(Poling et al., 2001). ….......... A-12
Table B.1.3: Ionic species thermodynamic properties (Zemaitis et al., 1986). ….......... A-15
Table B.1.4: Temperature dependence of Henry’s Law constants. ….......... A-15
Table B.1.5: Criss-Cobble entropy parameters (Criss and Cobble, 1964ab). ….......... A-16
Table B.1.6: Atmospheric solution heat capacity data. ….......... A-16
Table B.1.7: Parameter values for the Aspen Properties® heat capacity
polynomial. ….......... A-16
Table B.2.1: Antoine equation coefficients (Rowley et al., 1998). ….......... A-17
Table B.2.2: Parameter values for the modified Rackett equation (Spencer
and Danner, 1972). ….......... A-19
Table B.2.3: Atmospheric solution mass density data. ….......... A-19

xiv
List of Tables

Table B.2.4: Pair parameter values for the Clarke Aqueous Electrolyte
Volume model. ….......... A-19
Table B.2.5: Component characteristic volumes. ….......... A-21
Table B.2.6: Coefficients for the DIPPR vapour viscosity model (Rowley et al.,
1998). ….......... A-23
Table B.2.7: Coefficients for the Andrade liquid viscosity equation (Reid et al.,
1977). ….......... A-24
Table B.2.8: Atmospheric solution viscosity data. ….......... A-25
Table B.2.9: Parameter values for the Jones-Dole viscosity equation. ….......... A-25
Table B.2.10: Coefficients for the DIPPR surface tension equation (Rowley et
al., 1998). ….......... A-27
Table B.2.11: Atmospheric solution surface tension data. ….......... A-28
Table B.2.12: Surface tension correlation coefficients. ….......... A-28
Table B.2.13: Coefficients for the DIPPR vapour thermal conductivity model
(Rowley et al., 1998). ….......... A-30
Table B.2.14: Coefficients for the DIPPR liquid thermal conductivity equation
(Rowley et al., 1998). ….......... A-32
Table B.2.15: Atmospheric solution thermal conductivity data. ….......... A-32
Table B.2.16: Liquid phase thermal conductivity correlation coefficients. ….......... A-32
Table B.2.17: Normal boiling points and the corresponding liquid molar
volumes (Poling et al., 2001). ….......... A-35
Table B.2.18: Ionic conductivities at infinite dilution. ….......... A-37
Table B.2.19: Diffusivities in water at 25°C. ….......... A-37
Table C.1.1: The Electrolyte NRTL adjustable parameters used in this work. ….......... A-39
Table C.1.2: The Electrolyte NRTL adjustable parameters used in this work. ….......... A-40
Table C.3.1: Statistical results for the CO2-K2CO3-KHCO3-H2O system data
regression runs. ….......... A-45
Table C.3.2: Suitable parameter value sets for the CO2-K2CO3-KHCO3-H2O
system. ….......... A-46
Table C.3.3: Statistical results for the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O
system data regression runs. ….......... A-47
Table C.3.4: Suitable parameter value sets for the CO2-H2S-K2CO3-KHCO3-
KHS-K2S-H2O system. ….......... A-48
Table D.2.1: Alternative mass transfer coefficient and effective interfacial area
correlations. ….......... A-54
Table E.3.1: Temperature dependent property correlation coefficients. ….......... A-67
Table F.2.1: Coefficient values for the HYSYS® liquid density tabular model. ….......... A-73
Table F.2.2: Coefficient values for the HYSYS® liquid viscosity tabular
model. ….......... A-77
Table F.2.3: Coefficient values for the HYSYS® liquid surface tension tabular
model. ….......... A-79
Table F.2.4: Coefficient values for the HYSYS® liquid thermal conductivity
tabular model. ….......... A-83
Table G.2.1: Statistical results for the CO2-K2CO3-H2O system data
regression runs. ….......... A-87

xv
List of Tables

Table G.2.2: Statistical results for the CO2-H2S-K2CO3-H2O system data


regression runs. ….......... A-87
Table I.2.1: System poles and zeros for the SGC-RSF control structure. ….......... A-111

xvi
Nomenclature

Nomenclature
Latin Letters
A – Step size or amplitude of limit cycle
*
A – Latini component parameter
3
Aca m /kmol Clarke Aqueous Electrolyte Volume parameter
AI – Debye-Hückel parameter
AAD % Average absolute deviation
a – Activity
6 2
a m ·bar/kmol Cubic equation of state mixture parameter
2 3
a m /m Specific surface area
2
aa W·m /K·kmol Riedel anion parameter
2
ac W·m /K·kmol Riedel cation parameter
2 3
aI m /m Effective interfacial area
6 2
aj m ·bar/kmol Cubic equation of state parameter for species j
aT,j kJ/kmol·K Criss-Cobble entropy parameter
Bca L/mol Breslau-Miller electrolyte parameter
3
b m /kmol Cubic equation of state mixture parameter
3
b m /kmol Ion contribution factor
3
ba,1 m /kmol Breslau-Miller anion parameter
3
ba,2 m /kmol·K Breslau-Miller anion parameter
bc,1 m3/kmol Breslau-Miller cation parameter
3
bc,2 m /kmol·K Breslau-Miller cation parameter
3
bj m /kmol Cubic equation of state parameter for species j
bT,j – Criss-Cobble entropy parameter
3
C kmol/m Molar concentration or molar density
3
mol/cm
3
Ĉ H2S mol/m Equivalent H2S content

Cj – Electrolyte NRTL parameter (zj for ions and 1 for molecular species)
C $jw – Reduced volume integral of species j at infinite dilution in water

Cp kJ/kmol·K Heat capacity


cal/mol·K
kJ/kg·K
Btu/lbmol·°R

Ĉ p kJ/kg·K Mass heat capacity

T kJ/kmol·K Average value of C p between 25°C and temperature T


Cp
25

CR – Reduced molar density


Cv J/mol·K Heat capacity at constant volume
CLDG – Closed-loop disturbance gain matrix
CN – Condition number
3
ct kmol/m Total mole concentration
D m Diameter
2
D cm /s Effective diffusion coefficient or diffusivity
m2/s

xvii
Nomenclature

2
Djk m /s Binary diffusion coefficient for species pair j-k
2
Dw cm /s Diffusion coefficient in water
2
jk m /s Maxwell-Stefan diffusivity for the binary species pair j-k
DC – Disturbance cost
DCN – Disturbance condition number
DR – Decay ratio
DRGA – Dynamic relative gain array
d – Height of first overshoot
d – Process load or disturbance
d(s) – Vector of process disturbances
dh m Hydraulic diameter
dN m Nominal packing size
dp m Particle diameter
df – Degrees of freedom
2
E kW/m Energy flux across the vapour-liquid interface
Estage – HYSYS® column overall stage efficiency
EMurph – Murphree vapour efficiency
EF – Enhancement factor
-19
e C Charge of an electron (1.60219×10 C)
F C/mol Faraday’s constant (96 485 C/mol)
F kmol/s Feed molar flow
F kmol/s Flow rate
F – F-Test result
F – BLT detuning factor
F1, F2,F3 – Chung functions
Fc – Fractional conversion of K2CO3 to KHCO3 and KHS
FCO2 – CO2 loading
2- -
Fj – Fractional conversion of CO3 to HCO3 due to species j absorption
Fmax m3/min Maximum flow through control valve
'Fin % Maximum percentage step change in inflow
f atm Fugacity
bar
Pa
fo – Particle friction factor
fT – Twu function at temperature T
fx,0 – Ely-Hanley function
G – Electrolyte NRTL parameter
G kmol/s Vapour phase molar flow
G – Transfer function
G(s) – Transfer function matrix
3
Ĝ sm /h Vapour phase standard volumetric flow

G(s) – Transfer function matrix with paired elements along the diagonal
E
G kJ/kmol Symmetric excess Gibbs energy
E*
G kJ/kmol Un-symmetric excess Gibbs energy
G(s) – Process transfer function

xviii
Nomenclature

2 2
g m/s Gravitational acceleration (9.81m/s )
H m Height
3
Hj bar·m /kmol Henry’s Law constant for species j in solution
Hj bar Henry’s Law constant for species j in pure water
3
H wj bar·m /kmol Henry’s Law constant for species j in water

Ha – Hatta number
HETP m Height of packing equivalent to a theoretical plate
h kJ/kmol Enthalpy
kJ/kg
h – Height of relay
h – Twu function
hx,0 – Ely-Hanley function
f
'h kJ/kmol Enthalpy or heat of formation
kJ/mol
'hvap kJ/kmol Enthalpy of vaporisation
'hfk kJ/mol Joback contribution for group k to the enthalpy of formation
I – Identity matrix
3
Ic kmol/m Molar concentration based ionic strength
Ix – Mole fraction based ionic strength
IAE – Integral of the absolute error
J – Objective function for the linear quadratic regulator problem
2
J kmol/m ·s Diffusion flux
K – Chemical equilibrium constant
kmol/m3
2 6
kmol /m
K – Gain
K’ – Pseudo-equilibrium constant
K’p – Integrator gain
k J/K Boltzmann constant (1.38066×10-23 J/K or 1.38066×10-16 erg/K)
erg/K
k m/s Mass transfer coefficient
k 1/s Reaction rate constant
m3/kmol·s
6 2
m /kmol ·s
kjk – Cubic equation of state binary interaction parameter for species pair j-k
k’ 1/s Pseudo-first-order reaction rate constant
k’’ 1/s Kinetic coefficient
L kmol/s Liquid phase molar flow
Lc – Closed-loop log modulus
'Lmax % Maximum allowable percentage deviation from setpoint
3
L̂ m /h Liquid phase volumetric flow

Le – Lewis number
M kmol Material holdup
MAD % Maximum absolute deviation
MIC – Morari index of integral controllability
MRI – Morari Resiliency Index

xix
Nomenclature

MW kg/kmol Molecular weight


g/mol
m – Manipulated variable
3
m kmol/m Solution molarity (total K2CO3 concentration)
2
N kmol/m ·s Molar flux across the vapour-liquid interface
N – Number
N – Order of multivariable system
Neqm – Number of equilibrium stages
Nk – Number of UNIFAC groups of type k for Joback method
No 1/mol Avogadro’s number (6.02205×1023 1/mol)
NC – Number of components
ND – Number of data points
NI – Niederlinski index
n kmol Number of moles
n – Number of species in the system
P atm Pressure or partial pressure
bar
Pa
psia
s
P bar Vapour pressure
Po min Period of oscillation
Pu min Ultimate period or limit cycle period
'P Pa Pressure drop
kPa
PRGA – Performance relative gain array
p – Pitzer-Debye-Hückel closest approach parameter (14.9)
p – p-value
Q – Objective function for the data regression runs
Q kJ/s Heat flow or duty
kW
QDR – Quarter decay ratio
R kJ/kmol·K Gas constant (8.3144 kJ/kmol·K or 0.0831447 m3·bar/K·kmol or 82.06
3
m ·bar/K·kmol atm·cm3/mol·K)
3
atm·cm /mol·K
R kmol/m3·s Molar reaction rate
RSP – Ratio setpoint
RGA – Relative gain array
RRMSQE – Residual root mean square error
Re – Reynolds number
rj m Born radius of species j
S f25 kJ/kmol·K Infinite dilution entropy at 25°C

S fT kJ/kmol·K Infinite dilution entropy at temperature T

SG – Specific gravity
SSQ – Sum of squared errors
'SG – Change in specific gravity
s – Laplace transform variable (s = i·Z)

xx
Nomenclature

T K Temperature
°C
°R
°F
Tref K Reference temperature (298.15 K)
T* – Dimensionless temperature
t min Time
U kJ Energy holdup
U(s) – Process input transfer function
u – Process input
3
V m Volume
3
V cm /mol Molar volume
3
m /kmol
f
Vca m3/kmol Clarke Aqueous Electrolyte Volume parameter

Ve L/mol Breslau-Miller effective volume


VRo , VRG – COSTALD reduced volumes

v m/s Velocity
* 3
v cm /mol Characteristic volume
~
v – Reduced molar volume
f 3
v cm /mol Partial molar volume at infinite dilution in pure water
m3/kmol
W – Scalar BLT function
Wi – Weight of data group i
WSSQ – Weighted sum of squares
wf – Weight fraction
wfK 2CO3 – Equivalent weight fraction of K2CO3

X – Effective local mole fraction


x – Liquid phase mole fraction
x – Component mole fraction
Y(s) – Process output transfer function
y – Vapour phase mole fraction
y – Component mass fraction
y – Process output or response
Z – Compressibility
Z – Chung parameter
Z cSt Twu kinematic viscosity parameter
Z° 1/atm Isothermal compressibility at infinite dilution in water
(0)
Z – Pitzer compressibility function for spherical molecules
(1)
Z – Pitzer compressibility deviation function
ZRA – Rackett parameter
z – Charge number
z – Feed mole fraction
z – Ionic charge
z – Secondary process variable

xxi
Nomenclature

Greek Letters
D – Chung parameter
D – Electrolyte NRTL non-randomness factor
D – Latini component parameter
2
D kW/m ·K Heat transfer coefficient
Dc – Riedel critical point parameter
Dj – Cubic equation of state alpha function for species j
E – Chung parameter
E – Latini component parameter
E – Packing specific constant
3
E CO2 m /kmol Contribution factor for CO2 absorption

G m Film thickness
G Debye Dipole moment
Gp – Chapman-Enskog-Brokaw polar parameter
H – Error
2
H C /J·m Dielectric constant
H erg Characteristic energy
I ° Phase lag
rad
I m3/m3 Phase volumetric holdup
3 3
I m /m Packing voidage
Ijk – Wilke viscosity function for species pair j-k
* m·K/W Stiel-Thodos parameter
J – Activity coefficient
J – Latini component parameter
J – Symmetric activity coefficient
*
J – Un-symmetric activity coefficient
K – Dimensionless film coordinate
K – Column stage efficiency
M kV Electrical potential
M – Fugacity coefficient
N – Chung association factor
O – Relative gain
O – Eigenvalue
O min IMC tuning parameter
O W/m·K Thermal conductivity
O – Vector of eigenvalues
O' W/m·K Ely-Hanley translational thermal conductivity contribution
O” W/m·K Ely-Hanley internal thermal conductivity contribution
2
Of m /·kmol Ionic conductivity at infinite dilution in water

P kJ·m/kmol Chemical potential

xxii
Nomenclature

P cP Dynamic viscosity
kg/m·s
Pa·s
'Pca cP Jones-Dole viscosity contribution term for electrolyte ca
Q cSt Kinematic viscosity
2
m /s
T min Dead time
3
U g/m Mass density
3
kg/m
V – Standard error associated with a data point
V(s) – Singular value
V Å Characteristic length
Vc N/m Critical surface tension parameter
'Vca N/m Onsager-Samaras surface tension contribution term for electrolyte ca
VL N/m Surface tension
dyne/cm
W – Electrolyte NRTL binary interaction energy parameter
W min Natural period of oscillation
W min Time constant
X – Stoichiometric coefficient
:D,jk – Diffusion coefficient integral for species pair j-k
Z – Acentricity
Z rad/min Frequency
1/6 1/2 1/2 2/3
[ K ·kmol /kg ·atm Dean-Stiel parameter
< – Chung function
\b – Riedel parameter
] – Damping factor

xxiii
Nomenclature

Subscripts
 Final or at infinity or at steady-state
0 Reference fluid
Abs Absorber
a, a’ Anion
aq Aqueous
av Average
B Bandwidth
b Normal boiling point
CL Closed-loop
CLR Closed-loop regulator
CLS Closed-loop servo
Cond Condenser
c Column
c Controller
c Critical
c, c’ Cation
ca Electrolyte consisting of cation c and anion a
co Cross-over
D Derivative
d Disturbance
diag Diagonal matrix
dry Dry packing
eq Equilibrium
est Estimated
expt Experimental
F Feed
f Film
f Forward reaction
G Vapour phase
H High gas throughput
H2O Water
I Integral
I Vapour-liquid interface
IG Ideal gas
i Stage
irr Irrigated packing
j Component or species
j Data point
jk Component or species pair j-k
K2CO3 Potassium carbonate
k Component or species
L Liquid phase
L Low gas throughput

xxiv
Nomenclature

LS Lean solution
lc Local composition
M Medium gas throughput
m, m’ Molecular species
m Measured
max Maximum
min Minimum
OL Open-loop
o Pre-loading
PV Process variable or measured process variable
p Process
R Reaction
Reb Reboiler
RG Raw gas
r Reverse reaction
ref Reference
SP Setpoint
s Solvent
T Temperature
T Total or mixture
t Total or mixture
u Ultimate
vap Vapour
w Water
x Fluid of interest

xxv
Nomenclature

Superscripts
 At infinite dilution in water or in solvent
* Un-symmetric convention
+ Adjusted matrix with positive diagonal elements
-1 Inverse
Born Born model
Chem Chemical
c Molar concentration basis
E Excess
eq Chemical equilibrium
ex Excess
f Formation
H Conjugate transpose
LP Low pressure or atmospheric pressure
o Reference
NRTL Non-Random Two-Liquid model
PDH Pitzer-Debye-Hückel model
Phys Physical
res Residual
s
s Saturation or at vapour pressure P
T Transpose
w Water
ZN Ziegler-Nichols

xxvi
Chapter 1: Introduction

CHAPTER 1

INTRODUCTION

Chemical absorption and desorption are two fundamental process operations. Involving the selective
transfer, facilitated by chemical reactions, of components between a vapour (or gas) mixture and a
liquid mixture, they play a significant role in the process industry, particularly in gas treatment
applications (Astarita et al., 1983).

One such gas treatment method is the widely used hot potassium carbonate process (Benson and
Field, 1959) for removing acidic gas impurities, such as carbon dioxide (CO2) and hydrogen sulfide
(H2S). A typical chemical absorption and desorption system, the hot potassium carbonate process
displays complex process behaviour, which arises from the combined influence of many different
driving forces, including multicomponent diffusion, chemical interactions and temperature gradients.
In practice, equilibrium is rarely achieved due to the rate-controlled nature of chemical absorption and
desorption (Chakravarty et al., 1985).

As a result of its assumption of thermodynamic equilibrium, the classical equilibrium stage approach is
considered inadequate for describing chemical absorption and desorption processes like the hot
potassium carbonate process. Instead, a more physically consistent approach is the non-equilibrium
rate-based method (Kenig et al., 2001). This approach separates the material and energy balances
for the two phases by only assuming equilibrium at the phase interface. It therefore enables a more
thorough consideration of the mass and heat transfer relations, the chemical reaction kinetics and the
column hydrodynamics.

In this thesis, a new non-equilibrium rate-based model for describing chemical absorption and
desorption is developed for the hot potassium carbonate process. Unlike previous models, this model
incorporates rigorous thermodynamic, heat and mass transfer relations and column hydrodynamic
calculations for both the absorber and regenerator columns. It also considers both the reaction rate
expression and enhancement factor approaches for representing the effects of the chemical reactions.
To demonstrate this new model, it is applied to the hot potassium carbonate process CO2 Removal
Trains at the Moomba Processing Facility, one of the key operations for Santos, a major Australian oil
and gas exploration and production company.

1
Chapter 1: Introduction

Santos does not currently have detailed process models of the Moomba CO2 Removal Trains, despite
the CO2 Removal Trains being one of the bottlenecks of the processing facility. Consequently, the
process models developed in this thesis have the potential to be particularly useful tools for Santos,
provided they are implemented in HYSYS®, Santos’s preferred simulation environment. Accurate and
detailed process models can enable the evaluation of different operating conditions for process
optimisation purposes without disrupting the operation of the actual plant. They can also save time
and human resources since process simulations can be relatively quick and easy to perform
compared to actual plant tests.

To demonstrate a potential application of the process models developed in this thesis, the models are
used to analyse the multivariable controllability of the CO2 Removal Trains. The results of this
analysis are used to develop diagonal control structures to fully automate the control of the individual
trains. Although diagonal control structures are the simplest and most widely used form of
multivariable control (Luyben and Luyben, 1997), their application to the hot potassium carbonate
process has not been previously published in the literature.

1.1 Project Objectives


In summary, the four objectives of this work are as follows:

1. To develop a new rigorous non-equilibrium rate-based model for the hot potassium carbonate
process;

2. To apply this model to the hot potassium carbonate process CO2 Removal Trains at the
Santos Moomba Processing Facility;

3. To develop accurate and detailed process models of the CO2 Removal Trains in HYSYS®, the
preferred simulation environment for Santos; and

4. To demonstrate a potential application of the HYSYS® process models by analysing the


multivariable controllability of the CO2 Removal Trains and developing diagonal control
structures for automating the control of the individual CO2 Removal Trains.

1.2 Thesis Structure


As a guide to the reader, the overall structure of this thesis is described below.

In Chapter 2, a review of the relevant literature is presented. The hot potassium carbonate process is
described and the operation of the Moomba CO2 Removal Trains is outlined. Next, the effective
simulation of the hot potassium carbonate process is explored: the non-equilibrium rate-based
modelling approach is discussed; electrolyte thermodynamics are considered; and the capabilities of

2
Chapter 1: Introduction

Santos’s preferred process simulation platform, HYSYS®, are examined. An overview of the simple
diagonal control structure for multivariable process control is also provided.

Due to the limitations of HYSYS® identified in Chapter 2, the new rigorous non-equilibrium rate-based
model for the hot potassium carbonate process was instead developed in the Aspen Custom Modeler
simulation environment. Chapter 3 describes the physical and thermodynamic property models that
were used to represent the behaviour of the hot potassium carbonate system, focussing particularly on
the regression of the electrolyte thermodynamic model parameters from literature data. Chapter 4
outlines the process model equations and discusses the necessary model adjustments identified from
preliminary column simulations. The benefits of increasing the model complexity and rigor are
evaluated, followed by the validation of the Aspen Custom Modeler® process models against steady-
state plant data.

Importation of the Aspen Custom Modeler® process models into HYSYS® was considered impractical
due to the large computation times associated with these process models. Consequently, a novel
approach, utilising column stage efficiencies and hypothetical components, was undertaken to
compensate for HYSYS®’s restricted capabilities in simulating the hot potassium carbonate process.
Chapter 5 outlines a series of parametric studies investigating the effects of various operating
parameters on the performance of the Aspen Custom Modeler® process models. The results of these
studies are used in Chapter 7 to develop the column stage efficiency correlations for the absorber and
regenerator process models. Chapter 6 discusses the creation of the hypothetical components and
the modifications to the standard HYSYS® property models to accommodate the hot potassium
carbonate system. Chapter 8 outlines the development and validation of the HYSYS® process
models of the Moomba CO2 trains.

To demonstrate a potential application of the HYSYS® process models, the controllability of the two
most dissimilar CO2 Removal Trains, Trains #1 and #7, was examined using the MATLAB® numerical
computing environment. Chapter 9 describes the derivation of simple first-order plus dead time
(FOPDT) process transfer function models from dynamic HYSYS® simulations for the two CO2
Removal Trains. Despite the wide use of the simple diagonal control structure for multivariable
control, its application to the hot potassium carbonate process has not been previously documented.
Consequently, the development and evaluation of an optimal diagonal control structure for CO2
Removal Trains #1 and #7 are presented in Chapter 10.

Finally, Chapter 11 outlines the conclusions for this thesis and the recommendations for future work.

3
Chapter 2: Literature Review

CHAPTER 2

LITERATURE REVIEW

The relevant literature for this thesis is reviewed in this chapter. To introduce the reader to the hot
potassium carbonate process and its application in the Santos Moomba Processing Facility, the
process background is briefly described and the operation and configuration of the Moomba CO2
Removal Trains is outlined. An overview of the process chemistry is also provided.

The next parts of the literature review consider the various aspects of developing an effective process
model for the hot potassium carbonate process. The non-equilibrium rate-based approach and its
application to chemical absorption and desorption processes, like the hot potassium carbonate
process, are reviewed. In this section, the relevant model equations are presented for the material
and energy balances, the mass and heat transfer relations and resistances, the column hydrodynamic
calculations and the chemical reaction kinetics. The following section briefly discusses electrolyte
thermodynamic models, which are necessary for the accurate representation of the non-ideal
behaviour of the electrolyte system governing the hot potassium carbonate process. It is noted that
rigorous electrolyte thermodynamics have only been previously implemented in modelling the
absorber column for the hot potassium carbonate process. Finally, the capabilities of HYSYS®,
Santos’ preferred simulation environment, are examined and a novel approach is considered for
simulating the hot potassium carbonate process in HYSYS®.

To complete the literature review, an overview of the simple diagonal control structure for multivariable
process control is provided. This includes a review of the various controllability analysis tools
associated with this form of multivariable control and a brief description of a simple method of tuning
such control structures. A survey of the available literature highlights the absence of any previous
research concerning the multivariable control of the hot potassium carbonate process.

4
Chapter 2: Literature Review

2.1 The Hot Potassium Carbonate Process


2.1.1 Process Background
The acidic nature of CO2 and H2S enables their removal through contact with alkaline solutions, such
as aqueous monoethanolamine (MEA), aqueous diethanolamine (DEA) and aqueous potassium
carbonate (K2CO3) solutions (Astarita et al., 1983). In 1959, U.S. Bureau of Mines researchers H. E.
Benson and J. H. Field invented a process which utilised these alkaline solutions to remove CO2 and
H2S from gas mixtures (Benson and Field, 1959). It was noted that this process performed best with
concentrated potassium carbonate solution, which became the preferred absorbent solution. A simple
process flowsheet for the hot potassium carbonate process is depicted in Figure 2.1.1.

Acid Gas

CONDENSER

Sweet Gas

Lean Solution

REGENERATOR
ABSORBER

PRESSURE
REDUCTION

Raw Gas

Rich Solution REBOILER

PUMP

Figure 2.1.1: A simple form of the hot potassium carbonate process.

The raw gas, containing the acid gas impurities, enters the bottom of the high pressure (>8 atm)
absorber countercurrent to the lean hot potassium carbonate solution which is fed to the top of the
absorber. Within the column, contact between the two phases results in the transfer of the acid gases
from the gas phase into the liquid phase. The sweet gas leaves through the top of the absorber while
the rich potassium carbonate solution, containing the absorbed acid gases, leaves from the bottom.
The rich solution is depressurised, causing some of the absorbed acid gas to flash off, before entering
the low pressure (1 – 3 atm) regenerator. The remaining acid gas is then removed in the regenerator
through reboil stripping. The removed acid gas is vented from the top of the regenerator while the
regenerated potassium carbonate solution is pumped back to the absorber.

The high pressure in the absorber is required to ensure that the acid gas partial pressure in the raw
gas exceeds the equilibrium pressure of the acid gas over the potassium carbonate solution (Benson
and Field, 1959). This is necessary for the absorption of the acid gas to take place. The high
operating pressure also enables the absorber to be run at temperatures close to the atmospheric
boiling point of the potassium carbonate solution (100° – 140°C) without excessive evaporation of the

5
Chapter 2: Literature Review

solution, thereby eliminating the need to heat the rich solution before it enters the regenerator.
Another advantage of the elevated absorber temperature is the increased solubility of potassium
carbonate and potassium bicarbonate. This enables the use of concentrated solutions between 20 to
40 wt% potassium carbonate, which facilitates greater acid gas removal. Kohl and Nielson (1997)
recommend a solution concentration of 30 wt% potassium carbonate as a suitable design value for
most applications.

Further research and development work has introduced many enhancements that have substantially
improved the process economics and extended the applicability of the hot potassium carbonate
process. One such enhancement is the addition of small quantities of amine or inorganic activators to
the potassium carbonate solution to increase the rate of acid gas absorption (Bartoo, 1984; Kohl and
Nielson, 1997). Other improvements involve modifications to the process flow scheme, such as the
inclusion of more complex split-flow absorbers and two-stage regenerators to increase the sweet gas
purity (Field et al., 1962; Benson and Parrish, 1984), and energy conservation features to improve the
process economy (Benson and McCrea, 1979; Grover, 1987).

To date, there are over 700 commercial installations of the hot potassium carbonate process
worldwide (UOP Gas Processing, 2000). Seven of these comprise the Raw Gas Conditioning (RGC)
Plant at the Santos Moomba Processing Facility. Located in the Central Australian desert, the
Moomba facility processes raw natural gas to reduce its CO2 and water content and to recover ethane
and heavier hydrocarbons. The removal of CO2 occurs in the RGC Plant, which consists of seven
parallel hot potassium carbonate process trains known as the CO2 Removal Trains (CO2 trains). The
configuration and operation of these CO2 trains are outlined in the following section.

2.1.2 The Santos Moomba CO2 Trains


The seven parallel Moomba CO2 trains are not identical, but share similar layouts, which are depicted
in Figure 2.1.2. The CO2 trains are based on a simple single-stage packed absorber and single-stage
packed regenerator configuration, similar to the flow scheme in Figure 2.1.1. Some energy
conservation features have been incorporated: power recovery turbines are coupled to the main
solution pumps for all seven trains, while gas-gas heat exchangers have been included for heat
recovery in CO2 trains #3 to #7. A proprietary amine activator, ACT-1, was previously added to the
potassium carbonate solutions for CO2 trains #1 to #5, but Santos has since discontinued its use.

The purpose of the Moomba CO2 trains is to remove sufficient CO2 to prevent the formation of dry ice
(solid CO2) in the cold sections of the downstream Liquids Recovery Plant (LRP) and to meet the
quality specification for sales gas, the final gas product from the Moomba Processing Facility. H2S
removal also occurs within the CO2 trains, but this is of minor consequence due to the relatively
insignificant levels of H2S in the raw gas.

6
Chapter 2: Literature Review

Potassium Carbonate Solution Flash Steam


Condenser Acid Gas

Overhead
Sour Water Catchpot
(Trains #3 & #4)

Water Cooler

ABSORBER REGENERATOR
th
(Only 1 bed (4 bed only in Trains #3 &#4)
in Trains #1
& #2)

Raw Gas Makeup


Gas-Gas Water
Heat
Exchanger
Solution
(Trains #3 & #4)
Reboiler
Sweet Gas

Sweet Gas LP Steam


Separator
Condensate
Receiver

Power Main Steam


Recovery Solution Turbine
Sour Water Turbine Pump Condensate

(a)

Potassium Carbonate Solution Flash Steam


(Trains #5 & #6) Condenser Acid Gas

Overhead
Catchpot

Sour Water

ABSORBER
(Only 2 beds REGENERATOR Makeup Water
in Train #5)

Flash Steam (Train #7)

Raw Gas LP Steam


Gas-Gas
Heat
Exchanger Solution
Reboiler

Sweet Gas

Sweet Gas
Separator
Condensate
Receiver

Power Main Steam


Recovery Solution Turbine
Sour Water Turbine Pump Condensate

(b)
Figure 2.1.2: Basic CO2 train process flow diagrams. (a) CO2 trains #1 to #4. (b) CO2 trains #5 to #7.

The raw gas entering the RGC Plant typically contains 16 to 20 mol% CO2 and 5 to 20 ppm H2S. This
variation is due to changes in the proportions of the raw gas supplied to Moomba from various gas
fields, whose individual CO2 content range from 5 to 45 mol%. The sweet gas leaving the RGC Plant
typically contains 2.5 to 3 mol% CO2 and 0.5 to 1 ppm H2S. However, due to the different absorber
designs, the CO2 content of sweet gas streams produced by the individual trains can vary between 0.1

7
Chapter 2: Literature Review

and 8 mol% depending on the operating conditions. The combined sweet gas from the RGC Plant is
fed to the LRP for separation into sales gas (which is predominantly methane), ethane, liquefied
petroleum gas (LPG) and condensate.

Due to the high CO2 content of the raw gas, the processing capacity of the CO2 trains is normally
limited by the CO2 removal capacity of the hot potassium carbonate process. Table 2.1.1 gives an
indication of the maximum capacities of each train for processing raw gas to produce sweet gas that
meets the Santos specification of 2.5 mol% CO2, and two sets of typical operating data from 2002 are
presented in Table 2.1.2. The performance of the CO2 trains is critical to the successful operation of
the Moomba Processing Facility since the RGC Plant is one of the facility’s bottlenecks. The gas flow
through the RGC Plant sets the gas flow through the Moomba Processing Facility and therefore the
sales gas, ethane and LPG production rates.

Table 2.1.1: Nameplate capacity of the CO2 trains.

CO2 Removal Raw Gas CO2 Content (mol%)


Capacity 16 18 20
(106 sm3/d) Maximum Raw Gas Capacity (106 sm3/d)
Train #1 0.30 2.17 1.89 1.67
Train #2 0.30 2.17 1.89 1.67
Train #3 0.50 3.61 3.15 2.79
Train #4 0.54 3.90 3.40 3.01
Train #5 0.58 4.19 3.65 3.23
Train #6 0.69 4.98 4.34 3.84
Train #7 0.96 6.93 6.04 5.35
Total 3.87 27.95 24.34 21.56
3 3
Note: sm /d refers to m /d at standard conditions, i.e. dry gas at 15°C and 1 atm.

8
Chapter 2: Literature Review
Table 2.1.2: Typical operating data for the CO2 trains from 2002.
CO2 Train Data Set #1 CO2 Train Data Set #2
#1 #2 #3 #4 #5 #6 #7 #1 #2 #3 #4 #5 #6 #7
Raw Gas
a
mol% CO2 18.5 18.5 18.5 18.5 18.5 18.5 18.5 19.2 19.2 19.2 19.2 19.2 19.2 19.2
ppm H2S 20 20 20 20 20 20 20 15 15 15 15 15 15 15
6 3 b
Flow (10 sm /d) 1.95 1.65 2.93 3.35 3.35 4.08 5.55 1.80 1.55 2.79 3.18 3.26 3.99 4.56
Temperature (°C) 38 38 95 95 95 95 95 35 35 93 93 93 93 93
Sweet Gas
a
mol% CO2 2.5 2.5 2.5 2.5 2.5 2.5 2.5 2.8 2.8 2.9 2.9 3.1 2.5 1.2
ppm H2S <4 <4 <4 <4 <4 <4 <4 <4 <4 <4 <4 <4 <4 <4
Temperature (°C) 110 110 110 110 110 110 110 112 112 113 112 112 113 113
Acid Gas
6 3 b
CO2 (10 sm /d) 0.32 0.27 0.48 0.55 0.55 0.67 0.91 0.30 0.26 0.47 0.53 0.54 0.68 0.83
c d d
Temperature (°C) 103 103 80 103 95 94 94 102 103 75 103 95 94 94
Lean Solution
e
CO2 loading 0.38 0.38 0.38 0.38 0.38 0.38 0.38 0.40 0.40 0.40 0.40 0.44 0.35 0.35
e
H2S loading 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
f
wt% K2CO3 27 27 30 30 30 30 30 28 28 30 29 30 31 30
3
Flow (m /h) 550 550 670 812 1000 1000 1068 481 475 638 758 923 940 946
9

Temperature (°C) 110 110 110 110 110 110 110 112 112 112 112 112 113 112
Rich Solution
e
CO2 loading 0.82 0.75 0.86 0.84 0.75 0.83 0.96 0.85 0.80 0.89 0.89 0.84 0.82 0.94
e -5 -5 -5 -5 -5 -5 -5 -5 -5 -5 -5 -5 -5 -5
H2S loading < 6×10 < 5×10 < 6×10 < 6×10 < 5×10 < 6×10 < 7×10 < 4×10 < 4×10 < 5×10 < 5×10 < 4×10 < 5×10 < 5×10
3
Flow (m /h) 570 567 700 845 1034 1040 1122 499 490 667 789 955 980 995
Temperature (°C) 110 110 117 117 117 117 117 112 112 118 119 118 120 120
Sour Water
ppm CO2 - - 500 500 500 500 500 - - 600 600 600 500 300
3
Flow (m /h) - - 1.3 1.5 1.5 1.8 2.6 - - 1.2 1.4 1.4 1.8 2.1
Absorber
Pressure (bar) 70 70 70 70 70 70 70 70 70 70 70 70 70 70
Pressure drop (bar) < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2
Regenerator
Pressure (bar) 1.3 1.3 1.3 1.3 1.3 1.3 1.3 1.3 1.3 1.3 1.3 1.3 1.3 1.3
Pressure drop (bar) < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1
Cooling Water Circuit
3
Water Flow (m /h) 456 456 818 1079 - - - 456 456 818 1079 - - -
Cooler Duty (MW) 6.89 6.89 14.24 17.80 - - - 6.89 6.89 14.24 17.80 - - -
Overhead Condenser
Temperature (°C) 60 60 60 60 60 60 60 60 60 60 60 60 60 60
Pressure (bar) 1.08 1.08 1.08 1.08 1.08 1.08 1.08 1.08 1.08 1.08 1.08 1.08 1.08 1.08
Solution Reboiler
g
Steam flow (t/h) 29.0 24.4 34.8 42.3 44.3 49.3 64.9 23.6 21.6 35.0 36.0 37.1 49.6 54.3
a
This refers to the CO2 content in the dry gas. b sm3/d refers to m3/d at standard conditions (dry gas at 15°C and 1 atm). c This is the vapour temperature entering the regenerator wash sections in CO2 trains #1 and
#2 and leaving the wash sections in trains #3 to #7. d The value for train #4 is unexpectedly high compared to the typical value of 80°C (Santos Ltd, 1998). e The CO2 and H2S loading refer to the moles of CO2 and
H2S per equivalent mole of K2CO3 in the solution. The equivalent moles of K2CO3 are defined as the total number of moles of K2CO3 in the solution if all the KHCO3, KHS and K2S in the solution are converted back
into K2CO3. f This refers to the equivalent weight percent of K2CO3 in the solution. g The low pressure steam enters the reboilers at 140°C and 3.6 bar, while the steam condensate leaves at 138°C and 3.4 bar.
Chapter 2: Literature Review

2.1.3 Process Chemistry


Figure 2.1.3 summarises the various absorption, desorption and equilibrium reactions that take place
in the hot potassium carbonate process. These reactions are discussed in detail in the following
sections.
CO2 H2S H2O

Gas Phase

Liquid Phase

+
K CO2 H2S H2O

OH- 2H2O H3O+ OH- H2O H2O H3O+


H2O

- -
HCO3 HS
- +
OH + H3O

OH- H2O H2O H3O+ OH- H2O H2O H3O+

2 2-
CO3 S

Figure 2.1.3: Reaction flow scheme for the hot potassium carbonate process.

2.1.3.1 Acid Gas Absorption


CO2 and H2S absorption takes place in the hot potassium carbonate process via a series of reactions,
as summarised in Table 2.1.3.

Table 2.1.3: Acid gas absorption reactions in the hot potassium carbonate process.

1. Physical Absorption
CO 2 ( g) mo CO 2 ( aq) (R2.1.1)
H2 S ( g) mo H2 S (aq) (R2.1.2)

2. Chemical Absorption
2 
CO 2  CO 3  H2 O mo 2 ˜ HCO3 (R2.1.3)
2  
H 2 S  CO 3 mo HCO 3  HS (R2.1.4)

Mechanism 1:
2 
CO 3  H 2 O mo HCO 3  OH  (R2.1.5)
 
CO 2  OH mo HCO 3 (R2.1.6)
 
H 2 S  OH mo HS  H 2 O (R2.1.7)
  2
HS  OH mo S  H2 O (R2.1.8)

Mechanism 2:

CO 2  2 ˜ H 2 O mo HCO 3  H3 O  (R2.1.9)
 
H 2 S  H 2 O mo HS  H3 O (R2.1.10)
 2 
HS  H 2 O mo S  H3 O (R2.1.11)
2  
CO 3  H3 O mo HCO 3  H 2 O (R2.1.12)

10
Chapter 2: Literature Review

The absorption process begins with the diffusion of the two acid gases to the gas-liquid interface
followed by physical absorption into the solution (reactions (R2.1.1) and (R2.1.2)). The aqueous CO2
and H2S then react with K2CO3 to give potassium bicarbonate (KHCO3) and potassium bisulfide
(KHS). Since K2CO3, KHCO3 and KHS are all strong electrolytes, they can be assumed to be fully
dissociated in water, and can therefore be represented by carbonate (CO32-), bicarbonate (HCO3-) and
bisulfide (HS-) ions (reactions (R2.1.3) and (R2.1.4)).

There are two possible mechanisms by which CO2 and H2S can react with the potassium carbonate
solution (Astarita et al., 1981). The CO32- ions can first react with water to generate hydroxide (OH-)
ions, which then react with CO2 and H2S to give HCO3- and HS- ions (reactions (R2.1.5) to (R2.1.7)).
The HS- ions also react with the OH- ions to give sulfide (S2-) ions (reaction (R2.1.8)). Alternatively,
CO2 and H2S react with water to produce HCO3-, HS- and hydronium (H3O+) ions, which then react
with the CO32- ions to form more HCO3- ions in what is known as the “acidic mechanism” (reactions
(R2.1.9), (R2.1.10) and (R2.1.12)). Further H3O+ ions and S2- ions are formed from the reaction
between the HS- and water (reaction (R2.1.11)).

Of the eight chemical absorption reactions, only the two CO2 reactions (R2.1.6) and (R2.1.9) are
kinetically controlled. The remainder involve simple proton transfers and are considered to occur
instantaneously. The instantaneous nature of the H2S reactions leads to a considerably faster
absorption rate for H2S compared to CO2, which enables the selective absorption of H2S over CO2
using the hot potassium carbonate process. However, the hot potassium carbonate process is not
suitable for treating gases for which H2S is the only acid gas impurity. Some CO2 has to be present in
the untreated gas in order to regenerate the potassium carbonate solution. Otherwise, all the K2CO3
in the solution will eventually be converted to KHS, which has been found to be essentially non-
regenerable (Tosh et al., 1960). This is further discussed in the next section on solution regeneration.

Reactions (R2.1.6) and (R2.1.9) provide the rate-controlling steps for CO2 absorption. At pH levels
exceeding 8, reaction (R2.1.6) is fast whereas reaction (R2.1.9) is extremely slow (Astarita et al.,
1981). Since hot potassium carbonate processes typically operate in the pH range of 9 to 11 (Kohl
and Neilson, 1997), the absorption rate of CO2 is predominantly governed by reaction (R2.1.6).
Reaction (R2.1.9) has a comparably negligible effect and can be ignored. It should also be noted that
within the pH range of interest, the equilibrium constants for reactions (R2.1.8) and (R2.1.11) are so
low that it can be assumed that essentially no S2- ions are formed and all the absorbed H2S is present
as either aqueous H2S or HS- ions (Astarita et al., 1983).

Neglecting reaction (R2.1.9), the overall reaction rate for the absorption of CO2 into hot potassium
carbonate solution can be expressed as (Danckwerts and Sharma, 1966):

R CO2
k '˜ C CO2  C CO2
eq
(2.1.1)

where k’ is the pseudo-first-order reaction constant, defined as:

11
Chapter 2: Literature Review

k' k OH ˜ C OH (2.1.2)

k OH is the forward rate constant for reaction (R2.1.6), C is the molar concentration, and CCO2 eq is

the concentration of CO2 at chemical equilibrium. CO2 can be considered to undergo a pseudo-first-
order reaction due to the buffering nature of carbonate-bicarbonate systems which ensures the
concentration of OH- ions near the surface of the liquid is not significantly depleted by the absorbed
CO2 and remains relatively constant.

Astarita and co-workers (1983) have presented the following equation for k OH :

2895
log k OH 13.635   0.08 ˜ Ic (2.1.3)
T

¦z
1 2
Ic ˜ j ˜ Cj (2.1.4)
2 j

where k OH is in m3/kmol·s, Ic is the solution ionic strength in kmol/m3, T is the temperature in K, z is

the ionic charge, and C is the molar concentration in kmol/m3. Equation (2.1.3) has been validated for
temperatures up to 110°C and for solution ionic strengths up to 8 kmol/m3. Since these conditions
correspond to the typical operating conditions of the hot potassium carbonate process, equation
(2.1.4) can be applied to this process with confidence.

Similar expressions for k OH have also been developed by other workers, such as:

¦b
2382 1 2
Pohorecki and Moniuk (1988): log k OH 11.916   ˜ j ˜ Cj ˜ zj (2.1.5)
T 2 j

6612
Kucka and co-workers (2002): ln k OH 31.121   E CO2 ˜ Ic (2.1.6)
T

The ion contribution factors bj for species relevant to the hot potassium carbonate process are given in
Table 2.1.4, and the temperature dependence of the contribution factor  CO2 for CO2 absorption into

potassium hydroxide (KOH) solutions is:


 CO2 0.00033968 ˜ T 2  0.021215 ˜ T  33.506 (2.1.7)

Equation (2.1.5) has only been validated at 20°C (although the first two infinite dilution terms have
been validated for up to 41°C) and is applicable to solution ionic strengths greater than 0.5 kmol/m3.
Equation (2.1.6) has been validated for temperatures up to 50°C and for solution ionic strengths up to
3 kmol/m3. These equations therefore may not be applicable to the hot potassium carbonate process,
unlike equation (2.1.3).

12
Chapter 2: Literature Review

Table 2.1.4: Ion contribution factors (Pohorecki and Moniuk, 1988).

Species b (m3/kmol)
+
K 0.220
OH- 0.220
HCO3- 0
a

CO32- 0.085
HS- 0
a

2- a
S 0
a
Not available so set as 0.

2.1.3.2 Solution Regeneration


The regeneration of the rich potassium carbonate solution requires the desorption of the absorbed
CO2 and H2S from the solution. Up to two-thirds of the absorbed acid gas is flashed off through the
significant reduction in pressure between the absorber and regenerator (Astarita et al., 1981), and the
remainder is desorbed through the reboil stripping of the partially regenerated solution. The CO2 and
H2S desorption reactions are summarised in Table 2.1.5.

Table 2.1.5: Acid gas desorption reactions in the hot potassium carbonate process.

1. Chemical Desorption
 2
2 ˜ HCO3 mo CO2  CO3  H2O (R2.1.13)
  2
HCO3  HS mo H2S  CO3 (R2.1.14)

Mechanism 1:

HCO 3 mo CO 2  OH  (R2.1.15)
2  
S  H2O mo HS  OH (R2.1.16)
 
HS  H 2 O mo H 2 S  OH (R2.1.17)
  2
HCO 3  OH mo CO 3  H2 O (R2.1.18)

Mechanism 2:

HCO 3  H3 O  mo CO 2  2 ˜ H 2 O (R2.1.19)
2  
S  H3O mo HS  H2O (R2.1.20)
 
HS  H 3 O mo H 2 S  H 2 O (R2.1.21)
 2 
HCO 3  H 2 O mo CO 3  H3 O (R2.1.22)

2. Physical Desorption
CO 2 ( aq) mo CO 2 ( g) (R2.1.23)
H2 S (aq) mo H2 S ( g) (R2.1.24)

The acid gas desorption process involves the reversal of reactions (R2.1.3) and (R2.1.4) to regenerate
K2CO3 and the two acid gases (reactions (R2.1.13) and (R2.1.14)), followed by the diffusion of the
evolved acid gases to the gas-liquid interface (reactions (R2.1.23) and (R2.1.24)). Reactions
(R2.1.13) and (R2.1.14) take place in the following sequence of steps. First, reactions (R2.1.6) to
(R2.1.8) and reactions (R2.1.9) to (R2.1.11) are reversed to generate CO2 and H2S from the HCO3-,

13
Chapter 2: Literature Review

HS- and S2- ions (reactions (R2.1.15) to (R2.1.17) and reactions (R2.1.19) to (R2.1.21), respectively).
Then reactions (R2.1.5) and (R2.1.12) are then reversed to regenerate the CO32- ions from the HCO3-
ions (reactions (R2.1.18) and (R2.1.22)).

Like their reverse reactions, reactions (R2.1.15) and (R2.1.19) are slow, rate-determining steps while
the remaining reactions are instantaneous proton transfers. Reaction (R2.1.19) is extremely slow
under the conditions of interest and its effect on the desorption rate of CO2 can be neglected
(Pohorecki and Kucharski, 1991).

It should be noted that reactions (R2.1.15) and (R2.1.19) will occur even if H2S is the only acid gas
impurity, as its absorption leads to the formation of both HCO3- and HS- ions. Consequently, with each
absorption-regeneration cycle, some HCO3- ions are lost from the system through the generation of
CO2. However, the absorption of H2S does not produce enough HCO3- ions to compensate for this
loss (Tosh et al., 1960). The regeneration of the potassium carbonate solution is dependent on the
presence of HCO3- ions, which are necessary in order for reactions (R2.1.18) and (R2.1.22) to
proceed. If no CO2 is absorbed from the untreated gas, there will eventually be insufficient HCO3- ions
in the system to regenerate the CO32- ions.

As for the absorption of CO2 into potassium carbonate solution, the desorption of CO2 from potassium
carbonate solutions can be treated as a pseudo-first-order reaction with the following overall reaction
rate:

R  CO 2
k '˜ CCO 2
eq
 CCO 2 (2.1.8)

where k’ is the pseudo-first-order rate constant as defined in equation (2.1.2), C is the molar
concentration, and the superscript eq denotes the condition of chemical equilibrium. Consequently,
the desorption reaction kinetics can be described by the expression for k OH (equation (2.1.3))

presented by Astarita and co-workers (1983).

An alternative approach has been proposed by Pohorecki and Kucharski (1991), who have developed
the following expression for k’:
C CO 2
k ' k "˜ 3
(2.1.9)
C HCO 
3

8514 2
ln k" 27.821   0.5456 ˜ Ic  0.02760 ˜ Ic (2.1.10)
T

The rate constant k’ and the kinetic coefficient k” are in 1/s, the molar concentrations C CO 2 and
3

C HCO  are in kmol/m3, the temperature T is in K, and the molar ionic strength Ic is in kmol/m3.
3

Equations (2.1.9) and (2.1.10) are valid for temperatures between 60 and 100°C and solution ionic

14
Chapter 2: Literature Review

strengths between 2 to 8 kmol/m3, which correspond to the typical operating conditions for the hot
potassium carbonate process.

2.1.3.3 CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O System Equilibria


Aside from the above absorption and desorption reactions, the chemistry of the hot potassium
carbonate process is also dependent on the equilibria governing the CO2-H2S-K2CO3-KHCO3-KHS-
K2S-H2O system, as summarised in Table 2.1.6.

Table 2.1.6: CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system equilibria.


Vapour-Liquid Equilibria
H 2 O ( g) mo H 2 O (aq) (R2.1.25)
H
CO 2 ( g) mCO

2
o CO 2 ( aq) (R2.1.26)
HH2 S
H 2 S ( g) mo H 2 S ( aq) (R2.1.27)

Liquid Phase Chemical Equilibria


K H2 O
2 ˜ H 2 O m o OH   H 3 O  (R2.1.28)
K CO 2  
CO 2  2 ˜ H 2 O m 
o HCO 3  H 3 O (R2.1.29)
 K  2
HCO 3  H 2 O mHCO
 3
o CO 3  H3 O  (R2.1.30)
K H2S  
H 2 S  H 2 O m o HS  H 3 O (R2.1.31)
K
HS   H 2 O mHSo S 2  H 3 O 

(R2.1.32)

The chemical dissociation equilibrium relations for five liquid equilibria (R2.1.28) to (R2.1.32) and the
temperature dependence of the equilibrium constants are given in Table 2.1.7. Also included are the
temperature dependence of the Henry’s Law constants for CO2 and H2S. The equilibrium and Henry’s
Law constants are taken on a mole fraction basis at infinite dilution in pure water and are valid for
temperatures between 0° and 150°C.

By taking CO2, H2S, H2O and K2CO3 to be the unspeciated liquid phase species in the CO2-H2S-
K2CO3-KHCO3-KHS-K2S-H2O system, the liquid phase ionic speciation is described by the equilibrium
relations in Table 2.1.7 and the liquid phase equations in Table 2.1.8. The corresponding vapour
phase composition is represented by the vapour-liquid phase equilibrium expressions in Table 2.1.8.
These equations can be extended to accommodate any inert molecular solute species, such as
nitrogen, methane and other hydrocarbons which are present in the Moomba CO2 trains.

15
Chapter 2: Literature Review

Table 2.1.7: Temperature dependence of the equilibrium and Henry’s Law constants.

Constant A B C D Equilibrium Relation d


B
ln K A   C ˜ ln T  D ˜ T T in K
T
a xH  ˜ x OH JH* 
*
˜ J OH
3O 3O
K H2O 132.899 -13445.9 -22.4773 0.0 2
˜ 2 (2.1.11)
x H2 O J H2 O

a xH  ˜ x HCO  J H* 
*
˜ J HCO 
3O 3O
K CO2 231.465 -12092.1 -36.7816 0.0 3
˜ 3
(2.1.12)
2 2
x CO 2 ˜ x H 2 O *
J CO 2
˜ J H2 O

a xH  ˜ x CO 2 J H* 
*
˜ J CO 2
3O 3O
K HCO  216.050 -12431.7 -35.4819 0.0 3
˜ 3
(2.1.13)
3 xHCO  ˜ x H2 O J HCO
*
 ˜ J H2 O
3 3

a xH  ˜ xHS  JH*  ˜ J
*
HS 
3O 3O
K H2S 214.582 -12995.4 -33.5471 0.0 ˜ (2.1.14)
x H 2 S ˜ x H2 O J H2 S ˜ J H2 O
*

b xH  ˜ xS2  J H*  ˜ J 2
*
3O 3O S
K HS -9.74196 -8585.47 0.0 0.0 ˜ (2.1.15)
xHS  ˜ xH2 O J HS  ˜ J H2 O
*

B
ln H A  C ˜ ln T  D ˜ T H in bar, T in K
T
c
HCO2 159.1997 -8477.711 -21.9574 0.00578075 -
a
HH2S 346.6251 -13236.800 -55.0551 0.05956500 -
a b c d *
Austgen and co-workers (1989) Drummond (1981) Chen (1980) x is the liquid phase mole fraction, J is the un-
symmetric activity coefficient, and J is the symmetric activity coefficient (Section A.1 in Appendix A contains a brief explanation
of the different activity coefficient reference states).

Table 2.1.8: Liquid phase relations and vapour-liquid equilibria expressions.

Liquid Phase Relations


Electroneutrality xH   xK x OH  x HCO   2 ˜ x CO 2   x HS   2 ˜ x S 2  (2.1.16)
3O 3 3

Overall Balance x H2O  x H O   x OH  x K   x CO 2  x HCO   x CO 2  x H2S  x HS   x S 2  1 (2.1.17)


3 3 3
+
K Balance xK 2 ˜ nK 2CO3 ˜ n total (2.1.18)
CO2 Balance x CO 2  x HCO
3
  x CO
3
2 n CO 2

 n K 2CO 3 ˜ n total (2.1.19)
H2S Balance x H2S  x HS  x S2  n H2S ˜ n total (2.1.20)
x H2O  2 ˜ x H O   2 ˜ x CO2  2 ˜ x HCO   3 ˜ x CO 2  2 ˜ x H2S  x HS 
3

 2 ˜ n CO
3 3
H2O Balance (2.1.21)
2
 3 ˜ n K 2CO3  2 ˜ n H2S  n H2O ˜ n total

Vapour-Liquid Equilibria Expressions


§ Vs
¨ L,H2O ˜ P  PH2O
s
·¸
For (R2.1.25) y H2O ˜ M H2O ˜ P x H2O ˜ J H2O ˜ M Hs O ˜ PHs O ˜ exp¨ ¸¸ (2.1.22)
2 2
¨ R˜T
© ¹

y CO2 ˜ M CO2 ˜ P x CO2 ˜ J CO2 ˜ H CO2 ˜ exp¨


*
§ v CO
¨
f
2
˜ P  PHs O ·¸
2

For (R2.1.26) ¸¸ (2.1.23)
¨ R˜T
© ¹

y H2S ˜ M H2S ˜ P x H2S ˜ J H* 2S ˜ HH2S ˜ exp¨


§ v Hf S ˜ P  P s ·
¨ 2 H2 O ¸

For (R2.1.27) ¸¸ (2.1.24)
¨ R˜T
© ¹
Notation: x is the liquid phase mole fraction in the speciated system; n is the number of moles in the unspeciated liquid
phase; ntotal is the total number of moles in the speciated liquid phase; y is the vapour phase mole fraction in equilibrium
with the speciated liquid phase; H is the Henry’s Law constant; VL is the liquid molar volume; v f is the partial molar
volume at infinite dilution in water; P is the system pressure; T is the absolute system temperature; R is the gas constant;
*
J and J are the symmetric and un-symmetric activity coefficients; and M is the fugacity coefficient. The superscript s
s
indicates the vapour pressure P .

16
Chapter 2: Literature Review

Since the inert species do not participate in any liquid phase chemical equilibria, the resulting material
balance and vapour-liquid equilibrium expression are:
xj n j ˜ n total for j = inert molecular species (2.1.25)


§ v fj ˜ P  P s
¨ H2O
·¸
y j ˜ Mj ˜P x j ˜ J *j ˜ H j ˜ exp¨ ¸¸ for j = inert molecular species (2.1.26)
¨ R˜T
© ¹

The left-hand side of the overall balance (2.1.17) must also be extended to include xj for any inert
molecular species.

In summary, the hot potassium carbonate process involves a multicomponent system with two-phase
mass transfer coupled with chemical reactions. Its simulation therefore requires the consideration of
multicomponent mass transport theory, chemical reaction theory, and the effects of various driving
forces such as multicomponent diffusion and chemical interactions. This is best achieved via the non-
equilibrium rate-based approach (Taylor and Krishna, 1993; Kenig et al., 2001, Brettschneider et al.,
2004), which is discussed in the following section.

17
Chapter 2: Literature Review

2.2 The Non-Equilibrium Rate-Based Approach


2.2.1 The MESH Equations
The non-equilibrium rate-based approach is based on a set of balance equations, commonly known as
the MESH (Material balance, Equilibrium, Summation and entHalpy balance) equations (Taylor and
Krishna, 1993). For comparison, the MESH equations for the simple equilibrium stage approach and
the non-equilibrium rate-based approach are both provided in Table 2.2.1. Both approaches assume
the division of the absorption or desorption column into artificial height segments, called stages.
However, in the equilibrium stage approach, thermodynamic equilibrium is assumed to exist between
the vapour and liquid streams leaving each stage (Seader and Henley, 1998). In contrast, the non-
equilibrium rate-based approach assumes that equilibrium only exists at the vapour-liquid interface,
hence the separate balances for each phase (Taylor and Krishna, 1993).

For tray columns, each equilibrium stage represents an ideal tray, while each non-equilibrium stage is
equivalent to a real tray. As thermodynamic equilibrium is rarely achieved in practice, the equilibrium
stage approach uses tray efficiencies to relate the ideal trays to real trays, such as the widely used
Murphree vapour efficiency (Murphree, 1925):
y i, j  y i1, j
E Murph,i for j 1...NC (2.2.9)
K i, j ˜ x i, j  y i1, j

which replaces equation (2.2.2) in the equilibrium stage MESH equations. In the case of packed
columns, the packed bed is divided into a series of stages over which equations (2.2.1) to (2.2.8) can
be applied. For the equilibrium stage approach, the number of equilibrium stages Neqm is related to the
real packed column height Hc via the “height of packing equivalent to a theoretical plate” or HETP:
Hc HETP ˜ N eqm (2.2.10)

While the MESH equations are sufficient for the equilibrium stage approach, the non-equilibrium rate-
based approach requires an additional set of mass and energy transfer relations. These relations
describe the rate of mass and energy transfer across the vapour-liquid interface and are used to
determine the interphase molar and energy fluxes.

It should be noted that the inclusion of tray efficiencies can enable the satisfactory application of the
equilibrium stage approach to chemical absorption and desorption processes, as will be demonstrated
in the later chapters of this work. However, these tray efficiencies had to be first derived from a
rigorous non-equilibrium model in order for the equilibrium stage approach to reasonably represent the
effect of chemical absorption and desorption in the hot potassium carbonate process. The added
complexity of the non-equilibrium rate-based approach facilitates a greater understanding of the mass
and energy transfer processes that occur in such chemical absorption and desorption systems.

18
Chapter 2: Literature Review
Table 2.2.1: The MESH equations for a stage I and j = 1…NC components.

Equilibrium Stage Approach Non-Equilibrium Rate-Based Approach

Gi, yi,j, Ti, hGi Li-1, xi-1,j, Ti-1, hLi-1


Gi, yi,j, Ti, hGi Li-1, xi-1,j, Ti-1, hLi-1
QGi Stage i QLi
NGi,j TIi NLi,j
Stage i
Stage Diagram: EGi xIi,j ELi
Fi, zFi,j, TFi, hFi Qi
FGi, zFGi,j, TFi, hFGi yIi,j FLi, zFLi,j, TFi, hFLi
Gi+1, yi+1,j, Ti+1, hGi+1 Li, xi,j, Ti, hLi
Gi+1, yi+1,j, Ti+1, hGi+1 Li, xi,j, Ti, hLi

wM Gi, j
FGi ˜z GFi, j  G i1 ˜y i1, j  G i ˜y i, j  N Gi, j ˜ a Ii ˜ Vi
wMi, j wt
Fi ˜zFi, j  Gi 1 ˜y i 1, j  Li 1 ˜x i 1, j
Material Balance: wt wMLi, j (2.2.1)
FLi ˜zFLi, j  Li 1 ˜x i 1, j  Li ˜xi, j  NLi, j ˜ aIi ˜ Vi
 Gi ˜y i, j  Li ˜x i, j wt
NGi, j NLi, j
19

Equilibrium Relation: y i, j K i, j ˜x i, j (2.2.2)


NC NC
Summation Equations: ¦y
j 1
i, j 1 and
¦x
j 1
i, j 1 (2.2.3)

wU Gi
FGi ˜h FGi  G i1 ˜h Gi1  G i ˜h Gi  Q Gi  E Gi ˜ a Ii ˜ Vi
wUi wt
Fi ˜hFi  Gi 1 ˜hGi 1  L i 1 ˜hLi 1
Enthalpy Balance: wt wULi (2.2.4)
FLi ˜hFLi  Li 1 ˜hLi 1  Li ˜hLi  QLi  ELi ˜ aIi ˜ Vi
 Gi ˜hGi  L i ˜hLi  Qi wt
EGi ELi

y i, j ˜C Gti ˜ I Gi ˜ Vi
Material Holdup: Mi, j y i,j ˜C Gti ˜ I Gi  x i,j ˜CLti ˜ ILi ˜ Vi MGi, j
(2.2.5)
MLi, j x i, j ˜CLti ˜ I Li ˜ Vi
U Gi h Gi ˜C Gti ˜ I Gi  Pi ˜ Vi
Energy Holdup: Ui h Gi ˜C Gti ˜ I Gi  h Li ˜C Lti ˜ I Li  Pi ˜ Vi (2.2.6)
ULi h Li ˜C Lti ˜ I Li ˜ Vi
2
S ˜ D i ˜ Hi
Stage Volume: Vi (2.2.7)
4
Stage Pressure: Pi Pi1  'Pi (2.2.8)
Notation: M is the component material holdup; F, G and L are the feed, vapour phase and liquid phase molar flow rates; N is the molar flux across the vapour-liquid interface; aI is the effective
interfacial area for mass transfer; x, y and zF are the liquid phase, vapour phase and feed mole fractions; K is the equilibrium constant; NC is the number of components in the system; U is
the energy holdup; h is the enthalpy; Q is the stage heat loss; E is the energy flux across the vapour-liquid interface; T is the temperature; Ct is the phase molar density; I is the phase
volumetric holdup; V is the stage volume; D is the stage diameter; H is the stage height; P is the stage pressure; and 'P is the stage pressure drop. The subscripts F, G, L and I denote the
feed, vapour phase, liquid phase and the vapour-liquid interface, while the subscripts i and j refer to stage i and component j.
Chapter 2: Literature Review

2.2.2 The Mass Transfer Relations


The mass transfer at the vapour-liquid interface can be described by a variety of different theoretical
concepts, including the two-film model (Lewis and Whitman, 1923), the Higbie (1935) penetration
model, and the Danckwerts (1951) surface renewal model. The two-film model is most commonly
employed due to its simplicity and the wide range of correlations available in literature for estimating
its parameters.

Interface

NG NL
y
yf xI

TG yI xf x
TGf
TI TLf TL
EG EL
Vapour Bulk Vapour Liquid Liquid Bulk
Film Film

Figure 2.2.1: The two-film model for simultaneous mass and energy transfer.

In this model, two stationary films are postulated to exist adjacent to the vapour-liquid interface, as
depicted in Figure 2.2.1. It is assumed that all resistance to mass transfer is concentrated in these
two stagnant films, and that mass transfer within the films takes place only by steady-state molecular
diffusion due to the insignificant thickness of the films (Taylor and Krishna, 1993). The concentration
gradient is the driving force for diffusion across the films, and physical equilibrium is achieved at the
interface. In the bulk phases beyond the films, it is assumed that there are such high levels of mixing
that chemical equilibrium exists. These key assumptions result in one-dimensional mass transfer
normal to the interface (Taylor and Krishna, 1993).

Pilot plant studies of a hot potassium carbonate process for the simultaneous absorption of CO2 and
H2S have shown that for CO2 absorption, the liquid film resistance to mass transfer predominates, i.e.
CO2 absorption is liquid-phase limited (Yih and Lai, 1987). However, H2S absorption and desorption
were shown to be primarily influenced by the vapour film resistance, making them vapour-phase
limited processes. Neither film resistance were found to dominate during CO2 desorption, hence both
vapour-phase and liquid-phase resistances must be considered. The film resistances to mass transfer
are described by mass transfer coefficients, which are further discussed in a later section.

The multicomponent diffusion in the vapour and liquid films can be described by a number of
theoretical equations, which are summarised in Table 2.2.2. The most rigorous of these mass transfer
relations are the Maxwell-Stefan equations, which relate the component diffusion fluxes to their
chemical and electrical potential gradients. However, given their complexity, the Maxwell-Stefan
equations are often simplified by assuming that the system is sufficiently dilute that the diffusional
interactions between the components can be ignored. This gives rise to Fick’s Law for the vapour

20
Chapter 2: Literature Review

phase and to the Nernst-Planck equation for the liquid phase, both of which can be further simplified
by neglecting the convective mass transfer due to the bulk flow. If the effect of the electrical potential
gradient is ignored, the Nernst-Planck equation reduces to Fick’s Law.

Table 2.2.2: Mass transfer relations (Taylor and Krishna, 1993).

Vapour Phase Mass Transfer Relations


y fj dP Gj NC y fj ˜ J Gk  y fk ˜ J Gj
G G ˜ R ˜ TG
˜
dK G ¦ C Gt ˜Ð Gjk
Maxwell-Stefan k 1
(2.2.11)
Equations NC
NGj J Gj  y fj ˜ ¦ j 1
NGj


dy fj
J Gj C Gt ˜ k Gj ˜ Gt ˜ k Gj ˜ y j  y Ij
C av
dK G
Fick’s Law (2.2.12)
NC NC y j  y Ij
NGj JGj  y fj ˜ ¦j 1
NGj JGj  y av
j ˜ ¦
j 1
NGj and y av
j
2

Simplified Fick’s
Law
N Gj C Gt ˜ k Gj ˜
dy fj
dK G

C Gt ˜ k Gj ˜ y j  y Ij (2.2.13)

Liquid Phase Mass Transfer Relations


x fj dP Lj x fj ˜ z j ˜ F dM
NC x fj ˜ JLk  x fk ˜ JLj
G L ˜ R ˜ TL
˜
dK L

G L ˜ R ˜ TL
˜
dK L ¦k 1
C Lt ˜Ð Ljk
NC dx fj
Maxwell Stefan
Equations dM R ˜ TL
¦z j ˜ D Lj ˜
dK L NC (2.2.14)

dK L

F
˜
j 1
NC
and ¦ x fj ˜ z j

¦z 2 j 1
j ˜ D Lj ˜ x fj
j 1

Nernst-Planck C Lt ˜ D Lj § dx fj x fj ˜ z j ˜ F dM ·
NLj  ˜¨  ˜ ¸  x j ˜ N solvent (2.2.15)
Equation GL ¨ dK R ˜ TL dK L ¸
© L ¹

Simplified Nernst- CLt ˜ DLj § dx fj x fj ˜ z j ˜ F dM ·


NLj  ˜ ¨¨  ˜ ¸ (2.2.16)
Planck Equation GL © dKL R ˜ TL dKL ¸¹


dx fj
JLj C Lt ˜ k Lj ˜ av
C Lt ˜ k Lj ˜ x Ij  x j
dK L
Fick’s Law (2.2.17)
NC NC x j  x Ij
NLj JLj  x fj ˜ ¦
j 1
NLj JLj  x av
j ˜ ¦
j 1
NLj and x av
j
2


Simplified Fick’s dx fj
NLj C Lt ˜ k Lj ˜ C Lt ˜ k Lj ˜ x Ij  x j (2.2.18)
Law dK L

Mass Transfer D Gj D Lj
k Gj and k Lj (2.2.19)
Coefficients GG GL

Notation: x and y are the liquid and vapour phase mole fractions; G is the film thickness; R is the gas constant; T is
the absolute temperature;  is the chemical potential;  is the dimensionless film coordinate; J is the diffusion flux; N
is the molar flux; Ct is the total mole concentration or molar density; jk is the Maxwell-Stefan diffusivity for the
binary component pair j-k; NC is the number of components in the system; k is the mass transfer coefficient; D is
the effective diffusivity; z is the ionic charge; F is Faraday’s constant; and M is the electrical potential. The
subscripts G and L denote the vapour and liquid phases; the subscripts f and I denote the film and the vapour-liquid
th
interface; and the subscripts j and solvent denote the j component and the solvent. The superscript av denotes the
average.

21
Chapter 2: Literature Review

Both the simplified Nernst-Planck equation and the simplified Fick’s Law have been successfully
applied to the non-equilibrium rate-based modelling of the hot potassium carbonate process. Al-
Ramdhan (2001) developed her sieve plate absorber model based on the former equation, while
others (Joshi et al., 1981; Staton, 1985; Suenson et al., 1985; Marini et al., 1985; Sanyal et al., 1988;
Rao, 1990, 1991; Todinca, 1995; Todinca et al., 1997; Rahimpour and Kashkooli, 2004) have used the
latter equation to model packed absorber columns. A number of packed regenerator column models
have also been developed based on the simplified Fick’s Law (Staton, 1985; Suenson et al., 1985;
Marini et al., 1985; Rao, 1991; Todinca, 1994, 1995). The successful application of both mass
transfer equations indicates that the mass transfer in this electrolyte system can be reasonably
represented without including the effect of the electrical potential gradient.

However, studies on other similar acid gas removal processes have shown that while the exclusion of
the electrical potential gradient from the mass transfer relations does not adversely affect the
predicted removal of CO2, the removal of H2S is significantly underestimated (Schneider and Górak,
2001; Thiele et al., 2005ab). Nevertheless, given that the hot potassium carbonate process of interest
is focussed on the removal of CO2 and involves negligible levels of H2S, the omission of the electrical
potential gradient is considered to be a valid model simplification for this work.

2.2.3 The Energy Transfer Relations


As for mass transfer, the two-film model is used to describe the energy transfer across the vapour-
liquid interface. Consequently, all the resistance to energy transfer is assumed to be concentrated in
thin, stagnant vapour and liquid films existing on either side of the interface. The energy transfer
within these films is assumed to occur only by steady-state heat conduction, driven by the temperature
gradient. As depicted in Figure 2.2.1, the temperatures are equal at the interface and are uniform in
the bulk phases. These assumptions are analogous to those for the two-film model for mass transfer,
and lead to the equivalent conclusion of one-dimensional energy transfer normal to the interface.

The energy fluxes across the vapour-liquid interface are represented by the following equations, which
are analogous to Fick’s Law (Taylor and Krishna, 1993; Kucka et al., 2003):
NC NC
dTGf
Vapour: EG D G ˜
dK G
 ¦j 1
N Gj ˜ h Gj D G ˜ TG  TI  ¦N
j 1
Gj ˜ h Gj (2.2.20)

NC NC
dTLf
Liquid: EL D L ˜
dK L
 ¦
j 1
NLj ˜ h Lj D L ˜ TI  TL  ¦N
j 1
Lj ˜ h Lj (2.2.21)

where D is the heat transfer coefficient, T is the temperature, and h is the partial molar enthalpy. The
subscripts G, L and I indicate the vapour phase, the liquid phase and the vapour-liquid interface. The
first terms in equations (2.2.20) and (2.2.21) are the conductive heat fluxes which are analogous to the
diffusion fluxes J for mass transfer, while the second terms represent the convective contributions to
the energy transfer.

22
Chapter 2: Literature Review

For a non-reactive process, the inclusion of the above mass and energy transfer relations complete
the set of non-equilibrium rate-based model equations, known as the MERQ (Material balance, Energy
balance, Rate and eQuilibrium) equations (Taylor and Krishna, 1993). The following section
discusses the modifications required to include the effect of the hot potassium carbonate process
liquid phase chemical reactions.

2.2.4 The Effect of Chemical Reactions


The effect of chemical reactions can be incorporated into the MERQ equations in one of three ways:
chemical equilibrium relations, reaction rate expressions or enhancement factors. It should be noted
that no changes are required for the energy balances, provided the enthalpies are calculated using
standard enthalpies of formation, as these directly account for the heats of reaction.

2.2.4.1 Chemical Equilibrium Relations


Consider a homogenous chemical reaction of the form:
R
X A1 A 1  ...  X A NA A NA o X P1 P1  ...  X PNP PNP (R2.2.1)

– x
NA NP NA NP
R kf ˜ – x
a
Aa ˜ J Aa X Aa
 kr ˜
p
Pp ˜ J Pp
XPp
k cf ˜ –C
a
Aa
XA a
 k rc ˜ –C
p
Pp
XPp
(2.2.22)

where R is the molar reaction rate, k is the mole fraction rate constant, kc is the molar concentration
rate constant, X is the stoichiometric coefficient, x is the mole fraction, J is the activity coefficient, C is
the molar concentration, and the subscripts f and r denote the forward and reverse reactions.

If reaction (R2.2.1) is very fast or instantaneous, its effect on the rate of mass transfer can be
described by the chemical equilibrium relations:

– x
NP NP

kf p
Pp ˜ J Pp
XPp

k cf
–C p
Pp
XPp

K eq NA
and K ceq (2.2.23)
k rc NA

– x
kr

a
Aa ˜ J Aa
XA a
–C
a
Aa
XA a

where Keq is the mole-fraction equilibrium constant and K ceq is the molar-concentration equilibrium

constant. In the case of slow or kinetically controlled reactions like the rate-controlling steps for the
hot potassium carbonate process, reaction rate expressions or enhancement factors have to be
employed. The chemical equilibrium relations alone are inadequate for the hot potassium carbonate
process, as demonstrated by Park and Edgar (1984).

23
Chapter 2: Literature Review

2.2.4.2 Reaction Rate Expressions


The reaction rate R for slow or kinetically controlled reactions can be applied over an entire phase or
just within the film region. For the bulk liquid phase, the liquid phase component material balance in
the non-equilibrium MESH equations is replaced by:
wMLi, j
wt

FLi ˜z FLi, j  L i1 ˜x i1, j  L i ˜x i, j  NLi, j ˜ a Ii  R i, j ˜ I Li ˜ Vi for j 1...NC (2.2.24)

In the liquid film, mass transfer and chemical reactions occur simultaneously so the effect of the
reactions is included by performing a differential component balance over the film (Kucka et al., 2003):

1 wNLi, j
˜  R i, j 0 for j 1...NC (2.2.25)
G L wKL

This differential balance accounts for the difference between the component molar fluxes at the
vapour-liquid interface and at the film-bulk phase boundary due to the chemical reactions, and
replaces the material flux balance in the non-equilibrium MESH equations. The assumption of a linear
film concentration profile reduces equation (2.2.25) to:
NLi, j N Gi, j  R iav
,j ˜ G L for j 1...NC (2.2.26)

It may be valid to assume chemical equilibrium in the bulk liquid phase and a linear film concentration
profile for the hot potassium carbonate process. Schneider and Górak (2001) found that these two
assumptions did not significantly affect the predicted removal of CO2 and H2S for an amine gas
sweetening process, unlike the assumption of chemical equilibrium in the film. The film reactions are
therefore a key parameter that must be included when modelling chemical absorption processes.

Reaction rate expressions have been utilised with success in a number of non-equilibrium rate-based
absorption models for various acid gas removal processes (Kenig et al., 1999; Kucka et al., 2003;
Ebrahimi et al., 2003), but have not been implemented for the hot potassium carbonate process.

2.2.4.3 Enhancement Factors


Enhancement factors are the alternative method of incorporating the effect of slow or kinetically-
controlled reactions into the MERQ equations. These are theoretically derived factors which describe
the acceleration of the interfacial diffusion fluxes due to chemical reactions:
D Lj wC j
 ˜
Chem
JLj GL wKL
Interface
EFj for j 1...NC (2.2.27)
JPhys
Lj
av
k Lj ˜ C Lt
˜ x Ij  x j

where J is the diffusion flux, D is the diffusivity, G is the film thickness,  is the dimensionless film
av
coordinate, C is the molar concentration, k is the mass transfer coefficient, C Lt is the average liquid

24
Chapter 2: Literature Review

phase molar density, x is the liquid phase mole fraction, and NC is the number of components. The
subscripts I and L indicate the vapour-liquid interface and the liquid phase, while the superscripts
Chem and Phys indicate the presence and the absence of chemical reactions.

The application of such enhancement factors requires the liquid phase mass transfer relations in the
MERQ equations to be replaced by:
NC
NLj av
EFj ˜ C Lt
˜ k Lj ˜ x Ij  x j  x av
j ˜ ¦N Lj for j 1...NC (2.2.28)
j 1

Numerous enhancement factors have been proposed in literature (van Swaaij and Versteeg, 1992).
One of the most widely used is the expression developed by Danckwerts and Kennedy (1954) from
the surface renewal model:

K ' 1
K'1 ˜ 1  Ha A ˜
2
K'
EFA (2.2.29)
2 K '1
K ' 1  Ha A ˜
K'

CP DA ˜ k
K' and Ha A (2.2.30)
CA k LA

where K’ is the pseudo-equilibrium constant, k is the first-order or pseudo-first-order reaction rate


constant, k L is the liquid phase mass transfer coefficient, and Ha is the dimensionless Hatta number.
This enhancement factor is applicable to reversible first-order and pseudo-first-order reactions, like the
rate-controlling reactions in the hot potassium carbonate process. Danckwerts (1970) observed that
the predictions made by this enhancement factor were in numerical agreement with those by the
equivalent two-film theory enhancement factor. It is therefore applicable to the flux equation (2.2.28),
which was derived using the two-film theory.

The Hatta number compares the maximum rate of reaction in the liquid film to the maximum rate of
mass transfer through the liquid film, and can therefore be used to characterise the reaction kinetic
regime (Danckwerts, 1970). According to Charpentier (1981), reactions with Ha<0.3 are slow with
respect to the mass transfer and occur only in the bulk liquid phase, whereas reactions with Ha>3 are
fast with respect to the mass transfer and are completed within the liquid film. Moderately fast
reactions, characterised by 0.3 Ha 3, take place in both the bulk liquid phase and the liquid film.

There is a view that enhancement factors may be inadequate for chemical reaction systems involving
reversible reactions due to the use of inappropriate assumptions during their derivation (Schneider
and Górak, 2001; Noeres et al., 2003). This has been refuted by de Leye and Froment (1986) and
Mayer (2002), who compared the use of enhancement factors and reaction rate expressions for the
simulation of amine gas sweetening processes. Both methods produced very similar results, despite
the significantly greater computation time required by the more rigorous reaction rate expressions.

25
Chapter 2: Literature Review

This may be the reason why the previous models developed for the hot potassium carbonate process
(Joshi et al., 1981; Staton, 1985; Suenson et al., 1985; Marini et al., 1985; Sanyal et al., 1988; Rao,
1990, 1991; Todinca, 1994, 1995; Todinca et al., 1997; Al-Ramdhan, 2001; Rahimpour and Kashkooli,
2004) have used enhancement factors instead of reaction rate expressions to represent the effect of
chemical reactions. For confirmation, the use of reaction rate expressions and enhancement factors
will be compared in this work for the hot potassium carbonate process.

2.2.5 Mass and Energy Transfer Coefficients


The solution of the MERQ equations requires the mass and energy transfer coefficients, for which
there exist a number of empirical and semi-theoretical correlations. Since the Moomba CO2 trains
consist only of randomly packed columns, correlations for tray columns and columns with structured
packings are not considered here.

2.2.5.1 Mass Transfer Coefficients and the Interfacial Area


To effectively describe the mass transfer between the vapour and liquid phases in a randomly packed
column, it is imperative to know the vapour and liquid film resistances to mass transfer and the area
over which the two phases are in intimate contact such that mass transfer can occur. The film
resistances are represented by the vapour and liquid component mass transfer coefficients, while the
area of phase contact is known as the effective interfacial area. Numerous empirical and semi-
theoretical correlations for estimating these parameters are available in literature, a comprehensive
review of which has been presented by Wang and co-workers (2005).

After careful consideration of the available mass transfer coefficient and effective interfacial area
correlations, the widely used correlations proposed by Onda and co-workers (1968ab) will be used in
this work due to their general applicability and simplicity. These correlations are summarised in Table
2.2.3. The sole packing-specific constant required for these correlations is the critical surface tension
parameter Vc, which accounts for the wettability of different packing materials. For metal packings,
such as those used in the Moomba CO2 trains, Vc is equal to 0.075 N/m.

26
Chapter 2: Literature Review

Table 2.2.3: Mass transfer coefficient and effective interfacial area correlations (Onda et al., 1968ab).

NOTE:
This table is included on page 27 of the print copy of
the thesis held in the University of Adelaide Library.

2.2.5.2 Energy Transfer Coefficients


In an analogous manner to mass transfer, the vapour and liquid film resistances to energy transfer are
described by the vapour and liquid energy transfer coefficients. For simultaneous mass and energy
transfer, as in the case of the hot potassium carbonate process, the component energy transfer
coefficients D can be determined from the corresponding mass transfer coefficients k via the Chilton-
Colburn analogy (Bird and co-workers, 1960):
2
D Gj § OG · 3
˜¨ ¸
2
Vapour: U G ˜ Ĉ pG ˜ Le G 3 U G ˜ Ĉ pG (2.2.35)
k Gj ¨ U ˜ Ĉ ˜ D ¸
© G pG Gj ¹
2
D Lj § OL · 3
˜¨ ¸
2
Liquid: U L ˜ Ĉ pL ˜ Le L 3 U L ˜ Ĉ pL (2.2.36)
k Lj ¨ U ˜ Ĉ ˜ D ¸
© L pL Lj ¹

where is the mass density, Ĉ p is the mass heat capacity, Le is the dimensionless Lewis number, O

is the thermal conductivity, and D is the diffusivity. The subscripts G and L refer to the vapour and
liquid phases, while the subscript j denotes the component j.

The phase energy transfer coefficients can then be obtained from the mole-fraction-weighted
averages of the component coefficients:
NC NC
DG ¦y
j 1
j ˜ D Gj and DL ¦x
j 1
j ˜ D Lj (2.2.37)

27
Chapter 2: Literature Review

where x and y are the liquid and vapour phase mole fractions, and NC is the number of components in
the system.

While the above mass and energy transfer correlations enable the calculation of the mass and energy
transfer coefficients, the solution of the MERQ equations is incomplete without knowledge of two key
hydrodynamic parameters. These are discussed in the following section.

2.2.6 Packed Column Hydrodynamics


The hydrodynamic behaviour of randomly packed columns, such as the absorber and regenerator
columns in Moomba CO2 trains, can be described by two key parameters: the liquid volumetric holdup
and the column pressure drop. The former represents the quantity of liquid that accumulates within
the packing during column operation, while the latter describes the resistance to vapour flow due to
the packing and liquid volumetric holdup. A number of empirical and semi-theoretical correlations
have been proposed for describing the hydrodynamic behaviour of randomly packed columns, a
detailed review of which has been provided by Kister (1992).

The particle model hydrodynamic correlations presented by Stichlmair and co-workers (1989) will be
utilised in this work due to their simplicity and greater theoretical consistency than the corresponding
channel model hydrodynamic correlations (Stichlmair and Fair, 1998). As summarised in Table 2.2.4,
the Stichlmair correlations include a separate expression for the liquid volumetric holdup in the pre-
loading operating region, which was only validated for the air-water system. Although potassium
carbonate solution has a higher surface tension and viscosity than water, this expression may be valid
for the hot potassium carbonate process. Surface tension has a negligible effect on the liquid
volumetric holdup for high surface tension (>0.027 N/m) liquids (Strigle, 1994), and the equation has
been shown to be applicable for liquid viscosities up to 5 cP (Stichlmair et al., 1989).

Three packing-specific constants are required for the Stichlmair hydrodynamic correlations. These are
easily derived from the packing characteristics and the dry pressure drop curves published by packing
manufacturers. Table 2.2.5 gives the constants for a number of metal random packings, some of
which are used in the Moomba CO2 trains.

In summary, the non-equilibrium rate-based approach rigorously considers the effects of the chemical
potential, electrical potential and temperature gradients and the column hydrodynamics on the transfer
of mass and energy between the vapour and liquid phases. This makes it particularly suitable for
modelling rate-controlled processes like the hot potassium carbonate process. Since the hot
potassium carbonate process involves an aqueous electrolyte system, a key requirement for
developing an accurate and detailed process model is an effective representation of the complex non-
ideal electrolyte system thermodynamic behaviour. This is discussed in the next section.

28
Chapter 2: Literature Review

Table 2.2.4: Hydrodynamic relations (Stichlmair et al., 1989).


1
Pre-loading Liquid Volumetric § v 2 ˜a · 3
I Lo 0.555 ˜ ¨ L 4.65 ¸ (2.2.38)
Holdup ¨g˜I ¸
© ¹
§ § 'Pirr · ·¸
2

Liquid Volumetric Holdup IL I Lo ˜ ¨1  20 ˜ ¨¨ ¸¸ (2.2.39)


¨ ¸
© © H ˜ UL ˜ g ¹ ¹
Vapour Volumetric Holdup IG I  IL (2.2.40)
'Pdry 3 ˜ fo§ 1 I · U ˜ v 2
˜ ¨ 4.65 ¸ ˜ G G
H 4 ¨I ¸ dp
© ¹
C1 C2
fo   C3
Re G Re G
Dry Bed Pressure Drop (2.2.41)
U G ˜ v G ˜ dp
Re G
PG
6 ˜ 1  I
dP
a
2c
 4.65
'Pirr § 1  I  IL · 3 § I ·
¨¨ ¸¸ ˜ ¨¨1  L ¸¸
'Pdry © 1 I ¹ © I ¹
Irrigated Bed Pressure Drop (2.2.42)
1 §C C2 ·
c ˜¨ 1
 ¸
fo ¨ ReG 2 ˜ Re ¸
© G ¹
Vapour Head Contribution 'Pvap UG ˜ IG ˜ g ˜ H (2.2.43)
Overall Pressure Drop 'P 'Pirr  'Pvap (2.2.44)
Note: All variables are in SI units. Notation: I is the packing voidage; IL is the liquid volumetric holdup; IG is the vapour
volumetric holdup; ILo is the pre-loading liquid volumetric holdup; v is the flow velocity; a is the packing specific surface
area; g is the gravitational constant; 'P is the pressure drop over the packed bed; 'Pvap is the pressure drop due to the
static head of vapour in the packing; H is the height of packing; U is the mass density; dp is the packing particle diameter;
fo is the particle friction factor; Re is the Reynolds number; P is the viscosity; and C1, C2 and C3 are packing-specific
constants. The subscripts G and L refer to the vapour and liquid phases, while the subscripts irr and dry denote the
irrigated packed bed and the dry packed bed.

Table 2.2.5: Packing characteristics and constants for metal random packings.

Packing Type Number dN (m) a (m2/m3) I(m3/m3) C1 C2 C3


a
Cascade Mini-Ring® 2 0.044 164 0.950 63.90 -12.68 2.45
(CMR™) 3 0.063 105 0.960 38.05 -5.23 2.04
4 0.090 79 0.960 38.25 -3.51 2.03
Flexiring® a 2 0.050 102 0.960 15.98 -1.56 2.54
3.5 0.090 65 0.970 63.78 -8.78 2.82
Hy-Pak® a 2 0.060 88 0.971 47.31 -10.81 2.96
3 0.090 60 0.973 32.90 1.98 2.61
Intalox® Metal Tower Packing a 40 0.040 169 0.973 6.90 -1.32 1.78
(IMTP®) 50 0.050 81 0.978 3.88 -0.50 2.43
70 0.070 48 0.981 658.85 -78.21 4.36
Nutter Ring b 2 0.050 96 0.979 123.51 -28.46 4.35
2.5 0.063 83 0.982 91.74 -19.36 4.09
3 0.080 66 0.984 156.22 -21.33 3.98
Pall Ring® c 2 0.050 102 0.960 15.98 -1.56 2.54
3.5 0.090 65 0.970 63.78 -8.78 2.82
a b
Packing constants were derived from the dry pressure drop capacity curves provided by Koch-Glitsch (2003ab). Packing
c
constants were derived from the dry pressure drop capacity curves provided by Sulzer Chemtech (2006). Pall Rings® are
equivalent to Flexirings® (Kister, 1992).

29
Chapter 2: Literature Review

2.3 Electrolyte Thermodynamics


The non-ideal behaviour exhibited by electrolyte systems arises from the short-range molecular
interactions between the non-ionic species and the strong long-range electrostatic interactions
between the ionic species. The former type of interactions can be effectively described by traditional
thermodynamic models, such as the Redlich-Kwong equation of state (Redlich and Kwong, 1949), the
Non-Random Two-Liquid (NRTL) model (Renon and Prausnitz, 1968), and the Universal Quasi-
Chemical (UNIQUAC) model (Abrams and Prausnitz, 1975). However, these models are unable to
accommodate the additional complexity and non-idealities resulting from the electrostatic interactions.

Consequently, specialised electrolyte thermodynamic models were developed to enable a more


accurate description of electrolyte systems, as described in the excellent reviews provided by Zemaitis
and co-workers (1986), Renon (1996), Loehe and Donohue (1997) and Anderko and co-workers
(2002). Most of these electrolyte thermodynamic models are extensions of traditional thermodynamic
models, and can be divided into two broad categories: electrolyte activity coefficient models and
electrolyte equations of state. Electrolyte activity coefficient models are applied only to the liquid
phase, and require a traditional equation of state for the vapour phase. In contrast, electrolyte
equations of state are applied to both phases. The majority of the published electrolyte
thermodynamic models fall into the former category, including the highly popular Pitzer (Pitzer, 1973)
and Electrolyte NRTL (Chen and Evans, 1986) models. There are relatively few electrolyte equations
of state available; the most well-known of these is the Fürst-Renon equation of state (Fürst and
Renon, 1993).

An evaluation of the available electrolyte thermodynamic models identified the Electrolyte NRTL, the
Pitzer and the Extended UNIQUAC (Sander et al., 1986ab) models as the most suitable liquid phase
thermodynamic models for the hot potassium carbonate process (Ooi et al., 2005). Key factors
considered in this evaluation included: applicability to the hot potassium carbonate process,
robustness and computational accuracy, model simplicity and ease of implementation, and availability
in commercial process simulation software. The three models satisfied the above criteria, and have
been applied to vapour-liquid equilibria calculations for the CO2-K2CO3-KHCO3-H2O system with
success (Chen, 1980; Thomsen and Rasmussen, 1999; Al-Ramdhan, 2001; Cullinane and Rochelle,
2004; Hilliard, 2005; Ooi et al., 2005).

The major disadvantage with such rigorous electrolyte thermodynamic models is that these models all
involve a number of adjustable parameters. For example, the Electrolyte NRTL model includes two
types of adjustable parameters (the interaction energy parameter W and the non-randomness factor D),
whereas the Pitzer Model requires up to six different types (ionic interaction parameters B, C, ) and
<, and solute interaction parameters O and P). In order for these electrolyte thermodynamic models to
accurately describe the dependency of the individual component activity coefficients on the liquid
(and/or vapour) phase composition, these adjustable parameters have to be regressed from
experimental data. If insufficient data are available, rigorous thermodynamic models may perform no

30
Chapter 2: Literature Review

better than a simplified thermodynamic model that ignores the composition dependence of the activity
coefficients.

One such simplified thermodynamic model was presented by Astarita and co-workers (1983) for gas
treatment processes involving chemical absorption and desorption. It has been successfully adapted
for the hot potassium carbonate process and has been used in a number of non-equilibrium rate-
based models for this process (Joshi et al., 1981; Staton, 1985; Sanyal et al., 1988; Rao, 1990, 1991;
Rahimpour and Kashkooli, 2004). A similarly simple thermodynamic approach, which assumes
constant activity coefficients of unity, has also been applied to the non-equilibrium rate-based
modelling of the hot potassium carbonate process (Suenson et al., 1985; Marini et al., 1985; Todinca,
1994, 1995; Todinca et al., 1997). Only the absorber column model presented by Al-Ramdhan (2001)
for the hot potassium carbonate process includes rigorous electrolyte thermodynamics.

Despite the above reliance on simplified thermodynamic models, it is believed that there is sufficient
data on the hot potassium carbonate process to justify the use of a rigorous electrolyte thermodynamic
model. Consequently, in this work, the Electrolyte NRTL model will be used to perform the liquid
phase thermodynamic calculations for the non-equilibrium rate-based process model, while the
Soave-Redlich-Kwong (SRK) equation of state (Soave, 1972) will be applied to the vapour phase
calculations. The model equations for these two thermodynamic models are included in Sections A.2
and A.3 in Appendix A.

Having reviewed the various aspects of the non-equilibrium rate-based approach and selected the
electrolyte thermodynamic model, the final requirement to be considered in the development of an
accurate and detailed process model of the hot potassium carbonate process is the process
simulation platform. This is discussed in the following section.

31
Chapter 2: Literature Review

2.4 Process Simulation Platform


The commercial process simulation package HYSYS® is presently the primary process simulator
utilised by Santos due to its particular affinity for modelling oil and gas processing systems. It is
therefore the desired simulation platform for the Moomba CO2 train process models. However, there
are two significant disadvantages with using HYSYS® in this current work: the standard HYSYS®
component and property package libraries do not include electrolyte components or suitable
electrolyte property models 1, and HYSYS® employs the equilibrium stage approach with stage
component efficiencies for the calculation of column operations.

The absence of electrolyte components and appropriate electrolyte property models can be overcome
through HYSYS®’s customisation features, which enable the creation of hypothetical components and
custom property models. It is therefore theoretically possible to create electrolyte components and to
incorporate electrolyte property models into HYSYS® to enable the simulation of the hot potassium
carbonate process. However, such custom property models may not be sufficiently robust or reliable.
Alternatively, a commercial electrolyte package, such as the one from OLI Systems, can be added to
HYSYS®. However, these are limited to steady-state simulations and were therefore not considered
for this work.

A more viable option is to import custom unit operations with integrated property calculations into
HYSYS®. Such models can be developed in Aspen Custom Modeler®, an equation-based simulator,
which will enable the column operations to be modelled with the non-equilibrium rate-based approach.
Thermodynamic and physical property calculations can be incorporated into these models by
interfacing Aspen Custom Modeler® with the Aspen Properties® property calculation software, which
contains electrolytes and electrolyte models in its component and property model databanks. Aspen
Custom Modeler®, paired with Aspen Properties®, is therefore suitable for developing individual
process models of the unit operations in the hot potassium carbonate process for importation into
HYSYS®. However, this approach was considered impractical for two reasons: the high degree of
modelling rigor required to properly interface the imported models with HYSYS® to ensure sufficient
robustness and reliability; and more importantly, the significant computation time associated with the
Aspen Custom Modeler® models.

The absorber and regenerator columns are the key unit operations for the hot potassium carbonate
process. The performance of the overall process is dependant on these two columns, which are
heavily influenced by the non-ideal electrolyte system behaviour due to the process reactions. The
other incidental unit operations are less affected by the presence of electrolytes and can be modelled
using standard HYSYS® unit operations and hypothetical HYSYS® electrolyte components. Although

1
HYSYS® does have a proprietary Amines Property Package which is based on the Kent-Eisenberg model (Kent
and Eisenberg, 1976) for the CO2-H2S-Amine system. However, this model is specific to this particular electrolyte
system and is therefore not applicable to the hot potassium carbonate process.

32
Chapter 2: Literature Review

not theoretically correct, the thermodynamic and physical properties for these hypothetical
components can be calculated with a standard HYSYS® property package by adjusting the model
parameters to fit experimental data. This novel approach can be extended to the absorber and
regenerator columns by customising the standard HYSYS® column operations through the use of
stage efficiencies to represent the effects of the liquid phase reactions. These customised models can
then be validated by comparison with the more rigorous Aspen Custom Modeler® column models and
plant data.

Consequently, to satisfy the first three objectives of this work, rigorous non-equilibrium rate-based
column models will be developed in Aspen Custom Modeler®, while complete process models of the
Moomba CO2 trains will be developed in HYSYS®. Due to the computational time constraints, the
absorber and regenerator columns in the HYSYS® process models will be represented by customised
standard HYSYS® column operations, which will be validated against the rigorous Aspen Custom
Modeler® models and plant data.

It should be noted that there are available commercial software options which are suitable for
modelling electrolyte systems, such as ChemCAD, Aspen Plus®, VMGSim and gPROMS. While
these software packages apparently enable the simulation of the hot potassium carbonate process,
either through inbuilt electrolyte property packages 1 or via commercial electrolyte packages 2, they
were not used in this work as the selection of the final simulation platform was dictated by Santos.
The other alternative considered was to write a completely custom simulation program in Fortran or a
similar programming language. However, this approach was not pursued due to the availability of
equation-oriented simulation platforms like Aspen Custom Modeler®, which allow a similar level of
custom coding but provide access to inbuilt solver routines.

1
ChemCAD’s Electrolytes Package is based on the Electrolyte NRTL and Pitzer models, while Aspen Plus® has
Aspen Properties® inbuilt. VMGSim’s AMINES++ package was originally designed specifically for amine
electrolyte systems, but has apparently been extended to enable the simulation of acid gas removal using amines
with potassium carbonate solution. Unlike the other process simulators, the basis of the AMINES++ package is
not available in accessible literature and it is therefore difficult to judge its suitability for the hot potassium
carbonate process.
2
gPROMS can be integrated with Aspen Properties® or the OLI Systems commercial electrolyte package for
electrolyte thermodynamic and physical property calculations. It is similar to Aspen Custom Modeler® in that it is
an equation-oriented simulation environment that can facilitate the development of process models from first
principles.

33
Chapter 2: Literature Review

2.5 Multivariable Process Control


The final objective of this work is to demonstrate a potential application of the HYSYS® process model
by using it to analyse the controllability of the Moomba CO2 trains. This is a multivariable control
problem due to the multiple manipulated and controlled variables associated with the hot potassium
carbonate process. While a number of sophisticated multivariable control strategies have been
developed, the simple conventional diagonal (or decentralised) control structure remains the most
widely used approach for multivariable control (Luyben and Luyben, 1997). Consequently, this work
will focus on the development of diagonal control structures for the CO2 trains.

The diagonal control structure treats a multivariable or MIMO (multiple-output multiple-input) system
as a collection of multiple SISO (single-input single-output) control loops, in which one controlled
variable is paired with one manipulated variable. The design of such a control structure requires some
understanding of the dynamic characteristics of the process of interest. Luyben and Luyben (1997)
provide a concise review of the available methods for identifying the process dynamics, the two most
simple and popular of which are discussed in the following section.

2.5.1 Process Dynamic Behaviour


2.5.1.1 Step Response Models
The process dynamic behaviour can be determined from the process response to step inputs. Table
2.5.1 illustrates the most common types of dynamic process behaviour.

For the purpose of process control design and analysis, most chemical processes can be represented
by the simple FOPDT (first-order plus dead time) process model (Luyben and Luyben, 1997).
Integrating systems are a special case of a first-order system, and can be represented by a modified
form of the FOPDT model, known as the IPDT (integrating plus dead time) model. Oscillatory
processes (0 < ]< 1) are best described by the USOPDT (underdamped second-order plus dead
time) model; however, the FODPT model can approximate an USOPDT process via the following
parameters (Panda et al., 2004):
W
Kp Kp and Wp 2˜]˜W and T  T USOPDT (2.5.1)
2˜]

where TUSOPDT is the dead time for the USOPDT process. These simplified process models can also
be used to approximate systems with an inverse response, since an inverse response has a
deteriorating effect on control similar to that of a dead time (Skogestad, 2003). For such systems, the
e-T·s term represents the effective time delay due to the dead time and the inverse response.

34
Chapter 2: Literature Review

Table 2.5.1: Common dynamic process behaviour (Stephanopoulos, 1984; Wade, 2004).

Process Transfer Function Step Response Curve Step Response Function


y( t )
A Tangent at steepest point

Kp
First-Order Plus Dead Time
(FOPDT) 0.632·Kp § § t T · ·
¨ ¸
y( t ) ¨ ¨ W ¸¸
K p ˜ ¨1  e © p ¹ ¸
K p ˜ e  T˜ s A ¨¨ ¸¸
Gp ( s) © ¹
Wp ˜ s  1
t2
T t
t1 Wp = min{t1, t2}

y( t ) Tangent at steepest point


A
Integrating Plus Dead Time
(IPDT)

K p ˜ e  T ˜s y( t )
Gp ( s) K 'p ˜ t  T
Wp ˜ s A
K 'p ˜e  T˜s
s T t

K ' p ˜T

y( t )
A § § t T ·
 ] ˜¨¨ ¸¸ ·
1 ¨ ¸
e © W ¹
Kp+d
] y( t )
Underdamped Second-Order
1 2 ˜ S 2 ¨
Kp ˜ 1 ˜ sin Z ˜ t  T  I ¸
Plus Dead Time (USOPDT) ln DR
A ¨ ¸
Kp+d·DR ¨ 1 ]2 ¸
Kp © ¹
K p ˜ e  T˜s
Gp ( s ) 1 ]2
W2 ˜ s2  2 ˜ W ˜ ] ˜ s  1 Po ˜ 1  ] 2 Z
W W
2˜S
§ 1 ]2 ·
where 0  ]  1 ¨ ¸
T Po
t I tan 1 ¨ ¸¸
¨ ]
© ¹

Notation: Gp(s) is the process transfer function; y(t) is the step response function; Kp is the steady-state process gain; Wp
is the process time constant; K’p is the integrator gain; W is the natural period of oscillation; ] is the damping factor;T is
the dead time; A is the step size; DR is the decay ratio; Po is the period of oscillation; and d is the height of the first
overshoot.

2.5.1.2 Frequency Response Methods


Alternatively, the process dynamics can be identified from the process response to specific
frequencies. The most famous frequency domain method is the one proposed by Ziegler and Nichols
(1942), which is based on determining the point of marginal stability. This point can be found by
placing the process under proportional control and increasing the controller gain until a constant-
amplitude limit-cycle response is obtained. The controller gain at this point is the ultimate gain Ku and
the period of oscillation is the ultimate period Pu. A major disadvantage of this method is that the
system is driven towards instability, which may be dangerous in practice.

An improved frequency domain method was recently proposed by Åström and Hagglünd (1984). This
auto-tune variation (ATV) method inserts a relay as the process feedback controller in order to induce

35
Chapter 2: Literature Review

a small limit-cycle disturbance between the manipulated and controlled variables. The period of the
resulting limit-cycle oscillation is the ultimate period Pu, while the ultimate gain Ku can be calculated
from the oscillation amplitude A and the height of the relay h:
4˜h
Ku (2.5.2)
A˜S

Unlike the Ziegler-Nichols method, the ATV method can ensure the process is not unduly upset
through the careful selection of the relay height. However, the ATV method is only suitable for
processes with significant dead time (Svrcek et al., 2006).

Once the process dynamics have been identified by one of the above methods, they can be used to
design SISO controller algorithms, which are discussed in the following section.

2.5.2 SISO Controller Algorithms


2.5.2.1 PID Controllers
PID controllers are the simplest and most widely used method of process control in the chemical
processing industry (Svrcek et al., 2006). Detailed analyses of PID control theory are provided by
Seborg and co-workers (1989) and Stephanopoulos (1984), while Svrcek and co-workers (2006)
present a more concise overview for practical application purposes.

Numerous forms of PID controllers have been proposed (O’Dwyer, 2003), the simplest of which is the
parallel structure:
§ 1 ·
G c ( s) K c ˜ ¨¨1   W D ˜ s ¸¸ (2.5.3)
© WI ˜ s ¹

where Gc(s) is the controller transfer function, Kc is the controller gain, WI is the integral time constant,
and WD is the derivative time constant. Derivative action is not widely used in practice as it amplifies
any process noise, which can lead to large and unnecessary control action (Luyben and Luyben,
1997, Svrcek et al., 2006).

PID controllers are usually tuned by trial-and-error or via formal tuning rules. The former approach is
described in detail by Wade (2004), while O’Dwyer (2003) has compiled an extensive survey of the
formal tuning rules available in literature. The most well-known are the simple classical rules
developed by Ziegler and Nichols (1942, 1943), which can produce reasonable control performance,
but tend to be too aggressive for most applications as they give large controller gains and short
integral times (Svrcek et al., 2006). More conservative tuning rules have since been developed, the
simplest of which are the rules proposed by Tyreus and Luyben (1992). These produce less
underdamped control loops, and are therefore often preferred over the Ziegler-Nichols rules (Svrcek et

36
Chapter 2: Literature Review

al., 2006). Table 2.5.2 summarises the Ziegler-Nichols and Tyreus-Luyben tuning rules, while Table
2.5.3 presents PID tuning rules that are specific to liquid level control loops.

Table 2.5.2: Ziegler-Nichols and Tyreus-Luyben controller tuning rules.

Controller Kc WI (min) WD (min) Comments


Ziegler-Nichols (Ziegler and Nichols, 1942, 1943)
Wp K T
P or u – – d 1.0
Kp ˜ T 2 Wp
0.9 ˜ Wp Ku Pu T
PI or 3.33 ˜ T or – d 1.0
Kp ˜ T 2.2 1.2 Wp
1.2 ˜ Wp Ku Pu Pu T
PID or 2 ˜ T or 0.5 ˜ T or d 1.0
Kp ˜ T 1.7 2 8 Wp
Tyreus-Luyben (Tyreus and Luyben, 1992; Luyben, 1996)
0.49 K
PI or u 8.75 ˜ T or 2.2 ˜ Pu –
Kp ˜ T 3.2
0.72 K Pu
PID or u 8.75 ˜ T or 2.2 ˜ Pu 0.64 ˜ T or
Kp ˜ T 2.2 6.3

Table 2.5.3: Liquid level PID controller tuning rules (Wade, 2004).

NOTE:
This table is included on page 37 of the print copy of
the thesis held in the University of Adelaide Library.

2.5.2.2 Advanced Controller Algorithms


PID controllers do not always provide sufficient control action to produce the desired process
response. In such situations, improved process control may be obtained by the introduction of more
complex feedback control algorithms (e.g. cascade control or selective control) and/or feedforward
control action. Such advanced classical control algorithms are discussed in depth in a number of
texts, including those by Seborg and co-workers (1989), Stephanopoulos (1984) and Shinskey (1979).

For particularly difficult control problems, such as non-linear processes and non-stationary processes,
enhanced process control may be provided by modern control algorithms. The most popular of these
include adaptive, inferential, model predictive and intelligent control algorithms. Concise overviews of
these modern control algorithms are provided by Åström and Hagglünd (1995) and Seborg and co-
workers (1989), while more detailed discussions have been presented by Wittenmark (1997), Joseph

37
Chapter 2: Literature Review

(1999), Garcia and co-workers (1989), Lee and Cooley (1997), Qin and Badgwell (2003) and
Stephanopoulos and Han (1996).

For the purposes of this work, advanced controller algorithms will not be considered and only PID
controllers will be utilised in the development of the diagonal control structures for the CO2 trains.

2.5.3 Selection of Diagonal Control Structure


For a MIMO process like the hot potassium carbonate process, there exist a number of alternative
sets of controlled (output) and manipulated (input) variables that can be used to develop a diagonal
control structure. The selection of different sets of input and output variables for a given process will
result in different control structure alternatives. Typically, the set of controlled variables is selected on
the basis of engineering judgement (Luyben and Luyben, 1997), while the set of manipulated variables
is chosen using the analysis methods described below.

2.5.3.1 The Condition Number and the Morari Resiliency Index


Singular value decomposition (Golub and Kahan, 1965) is a useful tool for analysing the control loop
sensitivity of a MIMO process. Since process sensitivity affects the control system behaviour, singular
value decomposition is often used to select the set of manipulated variables for control.

One useful index derived from the singular value decomposition of the process transfer matrix Gp(s) is
the Morari resiliency index (MRI) (Yu and Luyben, 1986). This is the minimum singular value Vmin(s) of
Gp(s) and is a measure of the inherent ability of the process to handle uncertainties. The smaller the
MRI, the more sensitive the process is to uncertainties. Consequently, the set of manipulated
variables should be selected to give the largest MRI over the frequency range of interest. For
feedback control, the frequency range of particular importance is from about 0.01 to 1 rad/min
(Skogestad et al., 1990).

The singular value decomposition also gives the condition number (CN) (Grosdidier et al., 1985):
Vmax (s)
CN (2.5.4)
Vmin(s)

where Vmax(s) is the ratio of the maximum singular value of Gp(s). This index provides a quantitative
indication of the system sensitivity. A process is said to be well-conditioned and relatively insensitive
to uncertainties if CN is close to 1. Large condition numbers (CN>10) imply unbalanced sensitivity to
process uncertainties and may indicate control problems (Skogestad and Postlethwaite, 1996).
Consequently, the set of manipulated variables with the smallest CN should be selected.

It should be noted that CN and the MRI are scale-dependent and are therefore functions of the units of
Gp(s). To circumvent this problem, Gp(s) should be scaled based on physical grounds, e.g. by dividing
each variable by its allowed range (Skogestad and Postlethwaite, 1996).

38
Chapter 2: Literature Review

2.5.3.2 The Disturbance Condition Number and the Disturbance Cost


The sensitivity of a MIMO process to disturbances also affects the control system behaviour, and
should also be considered in the selection of the manipulated variables. Two useful indices are the
disturbance condition number (DCN) (Skogestad and Morari, 1987) and the disturbance cost (DC)
(Lewin, 1996):
1
Gp (s) ˜ Gd (s) ˜ d(s)
DCN 2
˜ Vmax (s) (2.5.5)
Gd (s) ˜ d(s) 2

1
DC Gp (s) ˜ Gd(s) ˜ d(s) (2.5.6)
2

where Gd(s) is the process disturbance transfer function matrix, and d(s) is the process disturbance
vector. Both DCN and DC provide an indication of the extent of control action required to reject a
disturbance. Small values for DCN (~1) and DC indicate the system is relatively insensitive to the
disturbance and hence less control action is needed to eliminate its effect. Consequently, the set of
manipulated variables should be selected to give the smallest DCN and DC. Like CN and the MRI,
DCN and DC are scale-dependent, so it is important that Gd(s) and d(s) are well-scaled.

It should be noted that the above four indices are dependent on the choice of controlled and
manipulated variables, but are unaffected by permutations in the control loop pairings. Therefore,
these indices can only be used to discriminate between alternative diagonal control structures and are
not applicable for selecting the configuration (variable pairings) for the diagonal control structure.

2.5.4 Selection of Diagonal Control Structure Configuration


Ideally, the diagonal control structure resulting from the selected controlled and manipulated variables
should have minimal interaction between its SISO control loops (Svrcek et al., 2006). Since there are
N! feasible control loops for an N×N MIMO system, the following analysis methods are used to identify
suitable diagonal control structure configurations.

2.5.4.1 The Relative Gain Array


The most well-known and extensively applied variable pairing tool is the relative gain array (RGA),
which was proposed by Bristol (1966) as a measure of the SISO control loop interactions in a diagonal
control structure. For an N×N MIMO process, the RGA is a dimensionless matrix whose ijth element Oij
is the relative gain of a controlled variable yi to a manipulated variable mj:

§ wyi ·
¨ ¸
Oij
©
§ wyi
wm j ¹m
·
(s)
Gp,ij (s) ˜ Gp
1 T
ij
for i,j = 1, …, N (2.5.7)
¨ ¸
© w m j¹y

39
Chapter 2: Literature Review

The relative gains are unaffected by scaling and provide a quantitative comparison of how each
manipulated variable affects each controlled variable. In general, variable pairings should be selected
such that the relative gains are positive and as close to 1 as possible over the frequency range of
interest (Skogestad and Postlethwaite, 1996). This minimises the interaction between the loops,
thereby preventing stability problems caused by interaction.

2.5.4.2 The Niederlinski Index and the Morari Index of Integral Controllability
Once suitable SISO control loop pairings have been determined from the RGA, the rows and columns
of Gp(s) can be reordered such that the paired elements are located along the diagonal to correspond
with the diagonal control structure. The reordered process transfer matrix Gp(s) enables the
application of the Niederlinski index (NI) (Niederlinski, 1971):

NI

det Gp (0) (2.5.8)
N

–G
i 1
p,ii (0)

to analyse the stability of the selected control loop pairings at steady-state (s=0). If all the SISO
controllers contain integral action and have positive loop gains, a negative NI proves that the specified
diagonal control structure will definitely be closed-loop unstable (Grosdidier et al., 1985). Therefore,
any variable pairings that give a negative NI should be eliminated. For a 2×2 system, a positive NI
satisfies the necessary conditions for closed-loop stability. However, for higher order systems, a
positive NI indicates the system may or may not be unstable, and its stability should therefore be
tested by dynamic simulation.

An alternative measure of the stability of a diagonal control structure is the Morari index of integral
controllability (MIC) (Grosdidier et al., 1985; Yu and Luyben, 1986):
MIC O G 
(0)
(2.5.9)
p

where O G 
(0)
is the vector of the eigenvalues of the matrix Gp+(0), which is obtained from Gp(0) by
p

adjusting the signs so that all the diagonal elements are positive. If all the SISO controllers contain
integral action and have positive loop gains, a negative eigenvalue for Gp+(0) will produce an unstable
diagonal control structure. Consequently, any variable pairings that result in a negative MIC element
should be eliminated.

2.5.5 Analysis of Diagonal Control Structure Performance


After an appropriate diagonal control structure configuration has been identified from the above
methods, its performance can be analysed through the use of the performance relative gain array
(PRGA) and the closed-loop disturbance gain (CLDG) matrix (Skogestad and Postlethwaite, 1996):
1
PRGA Gp,diag(s) ˜ Gp (s) (2.5.10)

40
Chapter 2: Literature Review

1
CLDG Gp,diag(s) ˜ Gp (s) ˜ Gd (s) (2.5.11)

where Gp,diag(s) is a diagonal matrix consisting of the diagonal elements of Gp(s).

For acceptable disturbance rejection and setpoint tracking performance, the following condition must
be satisfied for each control loop i, disturbance k and setpoint j (Skogestad and Postlethwaite, 1996):

^
1  GOL,i (s) ! max CLDG ik , PRGA ij
k, j
` (2.5.12)

GOL,i (s) Gp,ii(s) ˜ Gc,i (s) (2.5.13)

where GOL,i(s) is the open-loop transfer function and Gc,i(s) is the controller transfer function for loop i.
Consequently, disturbances and setpoint changes corresponding to large CLDG and PRGA elements
will be more difficult to control. It should be noted that |1+GOL,i(s)| need only be larger than |PRGAij| at
frequencies where the setpoints are tracked, which is usually limited to low frequencies (Z<1 rad/min)
(Skogestad and Postlethwaite, 1996). Similarly, having |1+GOL,i(s)| larger than |CLDGikj| is only
required at frequencies at which control is needed for disturbance rejection, i.e. when |CLDGikj| is
greater than 1 (Skogestad and Postlethwaite, 1996).

The validity of the above analyses should always be verified through dynamic simulation, which
requires the individual SISO controllers to be tuned. The next section describes such a tuning method
for diagonal control structures.

2.5.6 Tuning of Diagonal Control Structures


Since the early 1970s, a number of different controller tuning methods have been proposed for
diagonal control structures (Huang et al., 2003). The most well-known of these is the biggest log-
modulus tuning (BLT) method (Luyben, 1986), which will be applied in this work because of its
simplicity.

The BLT method involves the following steps:


1. The PI Ziegler-Nichols settings for each SISO controller for the selected control loop pairings
are calculated. For FOPDT process models, these settings are determined from the ultimate
gain Ku and ultimate period Pu, as shown in Table 2.52. Ku and Pu can be obtained from the
cross-over frequency Zco as follows:

Ii 180q
tan1  Wp,i ˜ Zco,i   Ti ˜ Zco,i for i = 1,…,N (2.5.14)

Ku,i Wp,i ˜ Zco,i 2  1


sign Kp,i ˜ for i = 1,…,N (2.5.15)

2˜S
Pu,i for i = 1,…,N (2.5.16)
Zco,i

41
Chapter 2: Literature Review

where I is the phase lag, Wp is the process time constant, T is the process dead time and Kp is
the process gain.

2. A detuning factor F is assumed (typically between 1.5 and 4) and is used to determine the
settings for each SISO controller from the calculated Ziegler-Nichols settings:

KcZN
,i
Kc,i for i = 1,…,N (2.5.17)
F
WI,i WIZN
,i ˜ F for i = 1,…,N (2.5.18)

where Kc,j are the controller gains, WIj are the controller integral time constants, and N is the
order of the MIMO system. The superscript ZN indicates the Ziegler-Nichols settings.

3. Based on the assumed value of F and the resulting controller settings, a multivariable Nyquist
plot is created of the scalar function W:
W
1  det I  Gp (s) ˜ Gc (s) (2.5.19)

where I is the identity matrix. This plot is used to check if the system is closed-loop unstable
by determining whether it encircles the critical (-1, 0) point. If the system is closed-loop stable
(i.e. it doesn’t encircle the critical point), then W is used to determine the closed-loop log
modulus Lc:
W
Lc 20 ˜ log (2.5.20)
1 W

4. F is adjusted until the biggest log-modulus Lc,max is equal to 2·N, where Lc,max is the peak in the
plot of Lc over the entire frequency range.

The above method has been found to yield a reasonable compromise between robustness and
performance (Luyben and Luyben, 1997), and can be extended to include derivative action and/or
individually weighted SISO loops (Monica et al., 1988).

2.5.7 Control of Acid Gas Removal Processes


A literature survey was performed to identify any previous applications of multivariable process control
techniques for the hot potassium carbonate process. While there is some published research on the
control of acid gas removal processes, it focuses on the removal of acid gases by amine solvents. No
research has been published that is specific to the hot potassium carbonate process. However, for
the interest of the reader, the available literature on the control of acid gas removal processes is
summarised below.

42
Chapter 2: Literature Review

A significant quantity of the available literature on the control of acid gas removal processes is based
on work performed on the large-scale computer-controlled CO2-MEA pilot plant at Imperial College in
London. Much of this research concerned the development and application of modern control
algorithms to obtain improved control for the pilot plant’s pressure, level, flow and composition control
loops. This included the multivariable controller based on the Nyquist array method (Rosenbrock,
1974) that was successfully designed and implemented by Albrecht and co-workers (1980) to control
the lean and rich solvent compositions. This multivariable controller was found to perform
considerably better than two independent SISO controllers. Due to the similarities between the CO2-
MEA process and the hot potassium carbonate process, this finding suggests that a multivariable
controller may be required for the hot potassium carbonate process.

A number of workers also considered the implementation of adaptive control on the Imperial College
pilot plant. These included Fortescue and co-workers (1981), who developed a self-tuning regulator
with a forgetting factor. This control scheme was applied to the pilot plant’s pressure loops, which
demonstrated its viability as an alternative to conventional PI controllers that require periodic manual
retuning due to process non-linearities. Based on this work, Ydstie and co-workers (1985) introduced
an element of model predictive control, resulting in a more robust and flexible self-tuning regulator.
The implementation of this control algorithm on the pilot plant’s pressure, flow and level control loops
showed that it could perform as well as a well-tuned PID controller over a wide range of operating
conditions. The extended horizon self-tuning regulator was further extended by Kershenbaum and
Pérez-Correa (1989) to include a feedforward component. The resulting feedforward/feedback self-
tuning regulator was applied to the absorber level control loop, and was found to be more robust and
effective than the simple feedback self-tuning regulator for the control of processes with large and
variable time delays.

Some work was also performed on the application of linear and non-linear model predictive control
algorithms on the Imperial College pilot plant, an overview of which was presented by Kershenbaum
(2000). It was found that both the linear DMC (dynamic matrix control) and non-linear receding
horizon model predictive control algorithms were very reliable in controlling the sweet gas composition
of the plant for both setpoint tracking and disturbance rejection in the absence of pilot plant/model
mismatches. However, poorer controller performance was observed when the pilot plant conditions
differed from those used to develop the process model. The performance deterioration was more
significant for the non-linear controller, and resulted in instability for sufficiently large mismatches
between the pilot plant and the process model. The division of the operating region into several zones
with different process models was found to partially compensate for this limitation.

More recently, the control of a smaller-scale CO2-amine pilot plant was studied by van der Lee and co-
workers (2001, 2003, 2004), with a particular focus on reducing the plant energy requirements. It was
found that a combination of model predictive and PID control could provide good disturbance rejection
and handling of setpoint changes, as well as lowering the energy consumption of the pilot plant. More

43
Chapter 2: Literature Review

effective disturbance rejection was achieved through the implementation of an on-line monitoring
system for the solvent CO2 loading, which enabled the liquid composition to be introduced as a
feedforward component in the control strategy.

It was noted that only a limited amount of work has been published in refereed journals on the control
of industrial-scale acid gas removal plants 1. However, it is expected that the control schemes
developed for the above pilot plants would be relatively insensitive to scale-up. This was
demonstrated by Alatiqi and co-workers (1993) who analysed the controllability of two industrial CO2-
MEA plants to identify a suitable diagonal control structure. A similar analysis was also performed on
the process transfer function matrix obtained by Albrecht and co-workers (1980) for the Imperial
College pilot plant, which resulted in the same loop pairings and closed-loop stability predictions as for
the industrial plants.

1
Some papers have been presented on the control of industrial scale acid gas removal plants at specialist
conferences, like the Laurence Reid Natural Gas Conditioning Conferences. Notable contributors to this subject
include William Poe and co-workers.

44
Chapter 2: Literature Review

2.6 Summary
In summary, there is a significant body of work that has been performed on the non-equilibrium rate-
based modelling of chemical absorption and desorption processes. However, the preceding literature
review has identified the following two gaps in the literature regarding the non-equilibrium rate-based
modelling of the hot potassium carbonate process:

¾ None of the previous models presented for the hot potassium carbonate process have used
reaction rate expressions to represent the effect of chemical reactions.

¾ None of the previous models presented for the hot potassium carbonate process have
considered rigorous electrolyte thermodynamics for both the absorber and regenerator
columns.

A considerable amount of work has also been published regarding multivariable process control, and
the control of acid gas removal processes has been studied in some detail. However, the above
review of the available literature noted the following gap concerning the process control of the hot
potassium carbonate process:

¾ No previous work has been published regarding the multivariable control of the hot potassium
carbonate process.

Consequently, in fulfilling the project objectives stated in Section 1.1, the succeeding chapters of this
thesis will attempt to address the above gaps in the literature, in accordance with the thesis structure
outlined in Section 1.2.

From Figure 1.2.1, it can be observed that the main body of this thesis is divided into four key parts,
the first being the above literature review. The second part, consisting of Chapters 3 to 5, focuses on
the work performed in Aspen Custom Modeler® to develop and apply a new non-equilibrium rate-
based model for the hot potassium carbonate process. Chapters 6 to 8 comprise the third part of this
thesis, and concentrate on the work performed in HYSYS® to develop process models of the CO2
trains using a novel approach. The fourth and final part of this thesis, Chapters 9 and 10, addresses
the application of the HYSYS® process models for analysing the multivariable controllability of the
CO2 trains.

45
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

CHAPTER 3

THERMODYNAMIC AND
PHYSICAL PROPERTIES OF THE
HOT POTASSIUM CARBONATE
SYSTEM

This chapter is the first of three chapters focussing on the simulation work performed in the Aspen
Custom Modeler® simulation environment to develop a new non-equilibrium rate-based model for the
hot potassium carbonate process.

As discussed in the preceding literature review, the hot potassium carbonate process involves a
multicomponent electrolyte system, which requires specialised property models to accurately
represent its complex non-ideal behaviour. In this chapter, a brief overview is provided of the various
property models that were used in this work to calculate the thermodynamic and physical properties
for the hot potassium carbonate system. Particular attention is paid to the regression of the adjustable
parameters for the Electrolyte NRTL model, which was used to determine the key property of interest:
the liquid phase activity coefficients.

46
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

3.1 Summary of Property Models


The thermodynamic and physical property calculations for modelling the hot potassium carbonate
process in Aspen Custom Modeler® were primarily performed using the Aspen Properties® property
calculation software. Where the property models provided in the Aspen Properties® databanks were
not suitable, the property calculations were instead performed using either alternative models from
literature or empirical correlations.

The property models used in this work are listed in Tables 3.1.1 and 3.1.2. Also included in these
tables are the references from which the model equations and/or model parameter values were taken.
Table 3.1.3 details the literature data used in the regression of some of the property model
parameters.

Table 3.1.1: Vapour phase thermodynamic and physical property models.

Property Model References


Fugacity coefficients SRK Soave (1972) a
b
Poling and co-workers (2001)
Enthalpy Ideal gas heat capacity Poling and co-workers (2001) b
a
Aspen Technology Inc (2003)
SRK Soave (1972) a
b
Heat capacity Via the vapour phase enthalpy Poling and co-workers (2001)
b
Molecular weight Weighted average Poling and co-workers (2001)
a
Molar density SRK Soave (1972)
a
Dynamic viscosity Dean-Stiel Dean and Stiel (1965)
DIPPR Rowley and co-workers (1998) a. b
Wilke Wilke (1950) a
a, b
Thermal conductivity DIPPR Rowley and co-workers (1998)
Stiel-Thodos Stiel and Thodos (1964) a
Wassiljewa-Mason-Saxena Poling and co-workers (2001) b
Mason and Saxena (1958) a
a
Diffusivity Blanc’s Law Reid and co-workers (1977)
a
Chapman-Enskog-Brokaw Poling and co-workers (2001)
a
Dawson-Khoury-Kobayashi Reid and co-workers (1977)
a b
Reference for model. Reference for model parameters.

47
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

Table 3.1.2: Liquid phase thermodynamic and physical property models.

Property Model References


a
Activity coefficients Electrolyte NRTL Chen and Evans (1986)
Enthalpy Electrolyte NRTL Chen and Evans (1986) a
b
Zemaitis and co-workers (1986)
Henry’s Law Chen (1980) b
Austgen and co-workers (1989) b
Fernández-Prini and co-workers (2003) b
b
de Hemptinne and co-workers (2000)
Ideal gas heat capacity Poling and co-workers (2001) b
Aspen Technology Inc (2003) a
Watson Reid and co-workers (1977) a
a, b
Heat capacity Aspen Properties® polynomial Aspen Technology Inc (2003)
b
Molecular weight Weighted average Poling and co-workers (2001)
b
Zemaitis and co-workers (1986)
a, b
Vapour pressure Extended Antoine Rowley and co-workers (1998)
Molar density Amagat’s Law Aspen Technology Inc (2003) a
Clarke Aqueous Electrolyte Chen and co-workers (1983) a
Rackett Spencer and Danner (1972) a
a
Partial molar volume Brelvi-O’Connell Brelvi and O’Connell (1972)
b
Austgen and co-workers (1989)
Zemaitis and co-workers (1986) b
Dynamic viscosity Andrade Reid and co-workers (1977) a
a
Breslau-Miller Breslau and Miller (1972)
Breslau and co-workers (1974) a
a
Jones-Dole Jones and Dole (1929)
Surface tension Empirical correlation -
Thermal conductivity Empirical correlation -
a
Diffusivity Nernst Horvath (1985)
Robinson and Stokes (1965)
Onda and co-workers (1971)
Ratcliff-Holdcroft Ratcliff and Holdcroft (1963) a
Stokes-Einstein Poling and co-workers (2001) a
Wang (1965) b
b
Ferrell and Himmelblau (1967)
b
Witherspoon and Saraf (1965)
Hayduk and Laurie (1974) b
Versteeg and van Swaaij (1988) b
Tamimi and co-workers (1994) b
a b
Reference for model. Reference for model parameters.

48
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

Table 3.1.3: Property data sources for the hot potassium carbonate system.

Property References
Vapour-liquid equilibria Tosh and co-workers (1959)
Tosh and co-workers (1960)
Liquid heat capacity Aseyev and Zaytsev (1996)
UOP Gas Processing (1998)
Liquid Mass density Armand Products Company (1998)
Chernen’kaya and Revenko (1975)
Bocard and Mayland (1962)
UOP Gas Processing (1998)
Liquid dynamic viscosity Chernen’kaya and Revenko (1975)
Correia and co-workers (1980)
Gonçalves and Kestin (1981)
Bocard and Mayland (1962)
UOP Gas Processing (1998)
Surface tension Armand Products Company (1998)
UOP Gas Processing (1998)
Bedekar (1955)
Liquid thermal conductivity Chernen’kaya and Revenko (1973)
UOP Gas Processing (1998)

The hot potassium carbonate process is governed by the behaviour of the CO2-H2S-K2CO3-KHCO3-
KHS-K2S-H2O system, in particular, its liquid phase thermodynamics. Consequently, the key property
of interest is the activity coefficients which, in this work, were calculated using the Electrolyte NRTL
model. The regression of the adjustable model parameters for the Electrolyte NRTL model is
discussed in the next section. For the purpose of brevity, detailed discussions of the other
thermodynamic and physical property models have not been included in the main body of this thesis,
but are instead located in Appendix B.

49
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

3.2 The Electrolyte NRTL Model


The Electrolyte NRTL model, as described in Section A.2 in Appendix A, was used to determine the
activity coefficients for this work. This model includes two forms of adjustable binary parameters, W
and D, which must be determined from experimental data to ensure an accurate representation of the
liquid phase thermodynamics. There are three types of these binary parameters: molecule-molecule,
molecule-electrolyte and electrolyte-electrolyte (with a common cation or anion). A complete list is
given in Table 3.2.1.

Table 3.2.1: Adjustable binary parameters for the Electrolyte NRTL model.

Binary Interaction Pair Energy Parameter W Non-randomness Factor D


Molecule-Molecule Wmm’ Wm’m Dmm’ Dm’m
Molecule-Electrolyte Wm,ca Wca,m Dm,ca Dca,m
Electrolyte-Electrolyte Wca,c’a Wc’a,ca Dca,c’a Dc’a,ca
Wac,a’c Wa’c.ac Dac,a’c Da’c.ac

It should be noted that not all the adjustable binary parameters are independent, as indicated by the
following relations:
W ca,c 'a W c 'a,ca and W ac,a'c  W a'c,ac (3.2.1)

D mm ' D m'm and D m,ca D ca,m (3.2.2)

D ca,c 'a D c 'a,ca and D ac,a'c D a'c,ac (3.2.3)

As suggested by Chen and Evans (1986), a default value of 0.2 was set for the non-randomness
factors, and the energy parameters were assumed to have the following dependence on temperature:

B § T ref  T § T · ·¸
W A  C˜¨  ln ¨ ref ¸¸ (3.2.4)
T ¨ T ©T ¹¹
©

where T is the absolute temperature and Tref is equal to 298.15 K. While values for the adjustable
parameters A, B and C can be determined from the regression of data, it is not always possible to
obtain statistically significant results for the adjustable parameters for all the binary interaction pairs
(Chen and Evans, 1986). It is therefore important to identify the key binary interaction pairs when
determining parameters for a system of interest.

For the hot potassium carbonate process, the system of interest is the CO2-H2S-K2CO3-KHCO3-KHS-
K2S-H2O system, which comprises of ten species: two cations (K+ and H3O+), five anions (HCO3-,
CO32-, HS-, S2- and OH-) and three molecular species (H2O, CO2 and H2S). Compared with the
concentrations of the other ions, the concentrations of H3O+ and OH- ions are negligible. Similarly, the
CO2 and H2S concentrations are insignificant when compared with that of H2O. Therefore, the four
key molecule-electrolyte binary interaction pairs are H2O-(K+,HCO3-), H2O-(K+,CO32-), H2O-(K+,HS-)
and H2O-(K+,S2-). The interactions between H2O and the two acid gases are also important, so the

50
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

parameters for the molecule-molecule binary interaction pairs of H2O-CO2 and H2O-H2S are also of
significance.

It should be noted that six possible salts can be formed from the cations and anions present in this
system: K2CO3, KHCO3, KHS, K2S, KOH and H2CO3. Electrolyte-electrolyte parameters can be
important when dealing with salt precipitation; however, the salt concentrations considered in this work
did not necessitate the inclusion of the respective salt precipitation equilibrium reactions. For this
reason, the electrolyte-electrolyte interactions were ignored and the corresponding parameters were
assumed to be zero.

The Aspen Properties® databanks include values for the adjustable parameters for the CO2-H2S-
K2CO3-KHCO3-KHS-K2S-H2O system. However, these values were found to give a poor fit to the
literature data. Consequently, the adjustable parameter values for the four key molecule-electrolyte
binary interaction pairs were re-regressed from literature data using the Aspen Properties® Data
Regression System (DRS). Adjustable parameter values for the other binary interaction pairs were
taken from the Aspen Properties® databanks. The full set of adjustable parameters used in this work
is provided in Tables C.1.1 and C.1.2 in Appendix C.

The re-regressed adjustable parameter values were determined via two sets of DRS regression runs,
which are described in the following sections. The data utilised in these regression runs consisted of
temperature T, pressure P, liquid phase mole fraction x and vapour phase mole fraction y values. The
maximum likelihood method of Britt and Luecke (1973) was used to minimise the objective function Q:


ª§ est
Ǭ T j  T j
exp t 2 · §

est
¸ ¨ Pj  P j
exp t 2 ·
¸
º
»
«¨ 2 ¸¨ 2 ¸ »
«¨© V Tj ¸ ¨ V Pj ¸ »
¹ © ¹
Q ¦W ˜¦ « » (3.2.5)

i
« NC1 § x est  x exp t · NC1 § y est  y exp t
2 2 ·»
i j
¨ j,k ¸ ¨ j,k ¸»
«
¦ ¦
j,k j,k
¨ ¸  ¨ ¸»
« V x j,k
2
V y j,k
2
«¬ k 1 ¨© ¸ k 1 ¨ ¸»
¹ © ¹¼

where Wi is the weight of data group i, V is the standard error associated with a data point, NC is the
number of components, the subscripts est and expt signify the estimated and experimental values for
each variable, the subscript j signifies the data point j, and the subscript k signifies the component k.

The objective function was minimised through the manipulation of the estimated values of the
variables and the model parameters. This minimisation process was subject to the vapour-liquid
equilibrium constraints, the chemical equilibrium constraints for all the reactions included in the
system, and the parameter bounds.

51
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

3.2.1 The CO2-K2CO3-KHCO3-H2O System


The first set of DRS regression runs was performed using Tosh and co-workers’ (1959) vapour-liquid
equilibrium data for the CO2-K2CO3-KHCO3-H2O system to obtain adjustable parameter values for the
H2O-(K+,HCO3-) and H2O-(K+,CO32-) binary interaction pairs.

Tosh and co-workers (1959) measured the partial pressure of CO2 over aqueous potassium
carbonate-potassium bicarbonate solutions as a function of the CO2 loading (or the fractional
conversion of K2CO3 to KHCO3), the solution K2CO3 equivalent weight percent 1 and temperature. 20,
30 and 40 wt% equivalent K2CO3 solutions were studied over a temperature range of 70° to 130°C
and a CO2 loading range of 0.10 to 0.83. These conditions correspond to the typical operating
conditions for the Moomba CO2 trains.

The liquid phase composition was determined using the apparent component approach. This method
ignores the solution chemistry and only considers the resulting molecular components present in the
liquid phase. For the CO2-K2CO3-KHCO3-H2O system, the apparent liquid phase components were
taken to be CO2, K2CO3 and H2O. The apparent liquid phase composition was determined from the
vapour-liquid equilibrium data using the following equations:
wf K 2CO3
n K 2CO3 (3.2.6)
MWK 2CO3

1  wf K 2CO3
n H2 O (3.2.7)
MWH2O

n CO2 n K 2CO3 ˜ FCO2 (3.2.8)

nj
xj for j = CO2, K2CO3 and H2O (3.2.9)
¦nj
j

where n is the number of moles, wf K 2CO3 is the equivalent weight fraction of K2CO3 of the solution,

MW is the molecular weight, F CO2 is the CO2 loading, and x is the liquid phase mole fraction.

The vapour phase constraint for the DRS regression runs was that only CO2 was present (Al-
Ramdhan, 2001; Hilliard, 2005), and the system pressure was taken to be the partial pressure of CO2.
The standard errors associated with the temperature, pressure and liquid component mole fractions
were set as 0.1°C, 5% and 0.1%, respectively. Since the vapour phase was assumed to be pure CO2,
the standard errors for the vapour component mole fractions were set as 0%.

1
The equivalent weight percent of K2CO3 is defined as the total concentration of K2CO3 in the solution if all the
KHCO3 in the solution is converted back into K2CO3.

52
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

A series of regression runs were performed, and the resulting optimal set of adjustable parameter
values is presented in Table 3.2.2. The procedure which was used to obtain the optimal parameter
set is outlined in Appendix C. Figures 3.1.1 and 3.1.2 compare the predicted CO2 partial pressure
values against the literature data. The average absolute deviation between the predicted and
literature values was 12.4%. This is considered to be acceptable since the relative uncertainty for
measurements of partial pressure is normally between 5 to 10%, but can be as high as 15%
(Thomsen and Rasmussen, 1999). In comparison, the average absolute deviation between the
predicted and literature values was 39.2% when the Aspen Properties® databank parameter values
were used.

Table 3.2.2: Electrolyte NRTL parameters for the CO2-K2CO3-KHCO3-H2O system.


a
Parameter Component j Component k Parameter Value Standard Deviation
+ 2-
A H2O (K ,CO3 ) -5.020 1.096
+ 2-
A (K ,CO3 ) H2O -0.176 0.105
(K+,HCO3 )
-
A H2O 6.250 1.215
+ -
A (K ,HCO3 ) H2O -3.728 0.966
+ 2-
B H2O (K ,CO3 ) -250.640 76.407
B (K+,CO32-) H2O -864.400 532.813
+ - b c
B H2O (K ,HCO3 ) 0 0
+ - b c
B (K ,HCO3 ) H2O 0 0
+ 2- b
C H2O (K ,CO3 ) 0 0c
C (K+,CO32-) H2O 0b 0c
+ - b c
C H2O (K ,HCO3 ) 0 0
C (K+,HCO3-) H2O 0
b
0
c

a b c
The parameter values are in SI units. Parameter value was fixed at zero. Parameter value was fixed so no
standard deviation was determined.

53
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

1.0E+01

20 wt%
30 wt%

Predicted CO2 Partial Pressure (bar)


40 wt%
1.0E+00

1.0E-01

1.0E-02

1.0E-03
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01
Experimental CO2 Partial Pressure (bar)

Figure 3.2.1: Comparison between the Electrolyte NRTL predictions and the experimental data. The
points represent the CO2 partial pressures over 20, 30 and 40 wt% equivalent K2CO3 solutions. The
dashed lines (---) represent the ± 10% lines.

1.0E+01
CO2 Partial Pressure (bar)

1.0E+00

1.0E-01

70°C 70°C 70°C


1.0E-02
90°C 90°C 90°C
110°C 110°C 110°C
130°C 130°C 130°C
1.0E-03
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

CO2 Loading CO2 Loading CO2 Loading

(a) (b) (c)


Figure 3.2.2: CO2 partial pressure over K2CO3 solution as a function of CO2 loading. The points represent
the data from Tosh and co-workers (1959) for (a) a 20 wt% equivalent K2CO3 solution, (b) a 30 wt%
equivalent K2CO3 solution and (c) a 40 wt% equivalent K2CO3 solution. The lines indicate the
corresponding Electrolyte NRTL predicted values. Note: The CO2 loading is not strictly dimensionless;
-1
its units are actually (mol CO2)·(mol K2CO3) .

54
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

3.2.2 The CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O System


In the second set of DRS regression runs, adjustable parameter values for the H2O-(K+,HS-) and H2O-
(K+,S2-) binary interaction pairs were obtained using Tosh and co-workers’ (1960) vapour-liquid
equilibrium data for the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system and the parameter values
given in Table 3.2.2.

Tosh and co-workers (1960) measured the partial pressures of H2S and CO2 over aqueous potassium
carbonate-potassium bicarbonate solutions as a function of the equivalent H2S content 1, the fractional
conversion of K2CO3 to KHCO3 and KHS, the solution K2CO3 equivalent weight percent 2 and
temperature. 20, 30 and 40 wt% equivalent K2CO3 solutions were studied over a temperature range
of 70° to 130°C, an equivalent H2S content range of 10 to 1500 mol/m3 and fractional conversions
between 0.073 and 0.68. It should be noted that the equivalent H2S content for the Moomba CO2
trains is typically between 0.5 and 2 mol/m3; however, no literature data are available for such low
concentrations of H2S in potassium carbonate-potassium bicarbonate solutions.

As for the previous set of partial pressure data, the apparent component approach was used to
determine the liquid phase composition. The apparent liquid phase components for the CO2-H2S-
K2CO3-KHCO3-KHS-K2S-H2O system were taken to be H2S, CO2, K2CO3 and H2O. The following
equations were used to determine the apparent liquid phase composition:
wf K 2CO3
n K 2CO3 (3.1.6)
MWK 2CO3

1  wf K 2CO3
n H2 O (3.1.7)
MWH2O

Ĉ H2S
n H2 S (3.1.10)
U solution ˜ 1000

n CO2 n K 2CO3 ˜ FC  n H2S (3.1.11)

nj
xj for j = H2S, CO2, K2CO3 and H2O (3.1.12)
¦nj
j

where n is the number of moles, wf K 2CO3 is the equivalent weight fraction of K2CO3 of the solution,

MW is the molecular weight, Ĉ H2S is the equivalent H2S content, is the density, Fc is the fractional

conversion of K2CO3 to KHCO3 and KHS, and x is the liquid phase mole fraction.

1 3
The equivalent H2S content is defined as the number of moles of H2S reacting with 1 m of solution to form
KHCO3 and KHS.
2
The equivalent weight percent of K2CO3 is defined here as the total concentration of K2CO3 in the solution if all
the KHCO3 and KHS in the solution are converted back into K2CO3.

55
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

Following advice from Aspen Technology (Popolizio, 2005) and the results of several preliminary
regressions, the vapour phase constraint was taken to be that only H2S was present.
Correspondingly, the system pressure was taken to be the partial pressure of H2S. The standard
errors associated with the temperature, pressure and liquid component mole fractions were set as
0.1°C, 5% and 0.1%, respectively. Since the vapour phase was assumed to be pure H2S, the
standard errors for the vapour component mole fractions were set as 0%.

A series of regression runs were performed, and the resulting optimal set of adjustable parameter
values is presented in Table 3.2.3. Appendix C outlines the method used to determine the optimal
parameter set. Figures 3.2.3 and 3.2.4 compare the predicted H2S partial pressure values against the
literature data. The average absolute deviation between the predicted and literature values was
16.5% with the regressed parameter values, compared to 81.9% when the Aspen Properties®
databank parameter values were used.

Table 3.2.3: Electrolyte NRTL parameters for the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system.


a
Parameter Component j Component k Parameter Value Standard Deviation
+ -
A H2O (K ,HS ) 3.076 1.028
+ -
A (K ,HS ) H2O -3.253 0.337
+ 2-
A H2O (K ,S ) 3.438 3.079
+ 2-
A (K , S ) H2O -6.305 1.311
+ - b
B H2O (K , HS ) 0 0c
B (K+, HS-) H2O -226.148 91.333
+ 2-
B H2O (K , S ) 4206.772 1127.216
B (K+, S2-) H2O -267.739 180.213
+ - b
C H2O (K , HS ) 0 0c
+ - b c
C (K , HS ) H2O 0 0
C H2O (K+, S2-) 0b 0c
+ 2- b c
C (K , S ) H2O 0 0
a b c
The parameter values are in SI units. Parameter value was fixed at zero. Parameter value was fixed so no
standard deviation was determined.

56
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

1.0E+00

20 wt%

Predicted H2S Partial Pressure (bar)


30 wt%
1.0E-01 40 wt%

1.0E-02

1.0E-03

1.0E-04
1.0E-04 1.0E-03 1.0E-02 1.0E-01 1.0E+00
Experimental H2S Partial Pressure (bar)

Figure 3.2.3: Comparison between the Electrolyte NRTL predictions and the experimental data. The
points represent the H2S partial pressures over 20, 30 and 40 wt% equivalent K2CO3 solutions. The
dashed lines (---) represent the ± 20% lines.

Equivalent H2S Content (mol/m 3)


0 100 200 300 400 500 600
1.0E+01 1.0E+00

66% Conversion
1.0E+00 1.0E-01
H2S Partial Pressure (bar)

H2S Partial Pressure (bar)


66% Conversion 1.0E-02
1.0E-01

33% Conversion
1.0E-02 1.0E-03

33% Conversion

70°C 70°C 70°C 1.0E-04


1.0E-03
90°C 90°C 90°C
110°C 110°C 110°C
130°C 130°C 20% Conversion
130°C
1.0E-04 1.0E-05
0 500 1000 1500 2000 0 200 400 600 0 20 40 60 80 100 120

Equivalent H2S Content (mol/m3) Equivalent H2S Content (mol/m3) Equivalent H2S Content (mol/m 3)

(a) (b) (c)


Figure 3.2.4: H2S partial pressure over K2CO3 solution as a function of equivalent H2S content. The points
represent the data from Tosh and co-workers (1960) for (a) a 30 wt% equivalent K2CO3 solution converted
only by H2S, (b) a 30 wt% equivalent K2CO3 solution converted by H2S and CO2 and (c) a 40 wt%
equivalent K2CO3 solution converted by H2S and CO2. The lines indicate the corresponding Electrolyte
NRTL predicted values.

57
Chapter 3: Thermodynamic and Physical Properties of the Hot Potassium Carbonate System

3.3 Summary
In summary, this chapter provided a brief description of the physical and thermodynamic property
models that were applied in this work to model the properties of the hot potassium carbonate system.
Particular attention was given to the Electrolyte NRTL model, while more detailed discussions of the
other property models were provided in Appendix B.

To ensure an accurate representation of the liquid phase thermodynamics for the hot potassium
carbonate system, the adjustable parameters for the Electrolyte NRTL model were regressed from
literature data. Comparison between the regressed parameter values and the default values provided
in the Aspen Properties® databanks showed that a significant improvement in model predictions was
achieved with the regressed values. Similar improvements in the model predictions for some of the
other physical and thermodynamic properties were obtained by replacing the Aspen Properties®
default model parameter values with values regressed from literature data or by developing empirical
correlations from the literature data.

The property models described in this chapter were used to perform the physical and thermodynamic
calculations for the non-equilibrium rate-based model that was developed for the hot potassium
carbonate process in Aspen Custom Modeler®. The following chapter discusses the development
and validation of this process model.

58
Chapter 4: Aspen Custom Modeler® Process Model Development

CHAPTER 4

ASPEN CUSTOM MODELER®


PROCESS MODEL
DEVELOPMENT

Having described the relevant physical and thermodynamic property models for the hot potassium
carbonate system in the previous chapter, this chapter continues the development of a new non-
equilibrium rate-based model for the hot potassium carbonate process by considering the model
equations and their implementation in Aspen Custom Modeler® for the Santos Moomba CO2 trains.

Three alternative stage models of varying complexity are presented in this chapter for the absorber
and regenerator columns for the hot potassium carbonate process, along with equations for several
simple ancillary unit models. The three stage models are evaluated via a series of preliminary Aspen
Custom Modeler® simulations to identify the optimal model in terms of model detail and complexity,
solution accuracy and computation time.

This chapter discusses the development of two model adjustments to correct the column models for
deviations between the model predictions and plant data from the Santos Moomba Processing
Facility. Also included is a sensitivity analysis to determine the effects of various model parameters on
the CO2 and H2S mass transfer rates predicted by the model.

Using the corrected column models and the ancillary unit models, CO2 train process models are
developed in Aspen Custom Modeler®, based on simplified configurations for the Moomba CO2 trains.
These process models are compared with the plant data used in the model development process, and
are also validated against an independent set of plant data.

59
Chapter 4: Aspen Custom Modeler® Process Model Development

4.1 Process Model Equations


4.1.1 The Column Model Equations
Mathematical equations for modelling packed bed absorber and regenerator columns can be derived
from either the equilibrium or the non-equilibrium rate-based stage approaches described in Section
2.2. In this work, three different sets of model equations were developed to identify the appropriate
model complexity and to facilitate simulation convergence:
1. An equilibrium stage model with chemical equilibrium (Model 1),
2. A non-equilibrium rate-based stage model with enhancement factors (Model 2), and
3. A non-equilibrium rate-based stage model with reaction rate expressions (Model 3).

The three models were developed based on the following key assumptions:
1. Adiabatic column operation
2. Vapour and liquid phase plug flow
3. Apparent composition basis for the material balances
4. Reactions in the liquid film and chemical equilibrium in the bulk liquid phase (Models 2 and 3)

The assumption of adiabatic operation was considered to be valid due to the use of insulation in the
Moomba CO2 trains to prevent heat losses, whereas the plug flow conditions were assumed to simplify
the modelling process. In reality, the vapour and liquid phase flows deviate from plug flow, with
significantly more mal-distribution occurring in the liquid phase (Stichlmair and Fair, 1998). In the CO2
trains, liquid re-distributors have been installed to reduce the degree of liquid mal-distribution;
however, it is expected that some mal-distribution still occurs. While it was possible to include the
effects of this mal-distribution in the column model equations, large-scale pilot plant studies would
have been required to investigate the mal-distribution in the CO2 trains, and these were beyond the
scope of this work.

Following the approach used by Mayer (2001), Brettschneider and co-workers (2004) and Thiele
(2007) to successfully develop chemical absorption and desorption models for various aqueous
electrolyte systems, the model material balances were based on the apparent liquid and vapour phase
compositions. For the liquid phase, the apparent composition ignores the solution chemistry, unlike
the true composition, and only consists of molecular or undissociated electrolyte species. The
apparent vapour phase composition, however, is identical to the true vapour phase composition due to
the absence of electrolyte and ionic species in the vapour phase.

For the CO2 trains, the apparent liquid phase composition consists of the acid gases (CO2 and H2S),
the solvent (H2O and K2CO3) and the inert vapour phase components (N2, CH4, C2H6, C3H8, C4H10,
C5H12, C6H14 and C7H16). In contrast, the true composition for this system is comprised of the vapour
phase components (CO2, H2S, H2O, N2, CH4, C2H6, C3H8, C4H10, C5H12, C6H14 and C7H16) and the
ionic species (H3O+, OH-, K+, CO32-, HCO3-, HS- and S-2) which participate in the chemical equilibria

60
Chapter 4: Aspen Custom Modeler® Process Model Development

(R2.1.28) to (R2.1.32). As depicted in Figures 4.1.1 to 4.1.3, the true and apparent compositions are
related by speciation relations, thereby enabling the computation of the true composition for the phase
equilibria calculations.

The fourth model assumption was established after an investigation into the Hatta numbers for the
rate-determining reactions (R2.1.6) and (R2.1.15) for the hot potassium carbonate process. It was
found that the Hatta numbers for the two reactions were greater than 3 under the typical operating
conditions for the CO2 trains. Consequently, the reactions were considered to be sufficiently fast that
they were completed within the liquid film and that the bulk liquid phase was at chemical equilibrium
(Charpentier, 1981). However, the two reactions could not be considered as instantaneous, unlike
proton transfer reactions, necessitating the inclusion of the liquid film kinetics in the two non-
equilibrium stage column models.

The model equations for the three column models are summarised on the following pages. It should
be noted that these equations only apply to the packed sections within the columns.

61
Chapter 4: Aspen Custom Modeler® Process Model Development

4.1.1.1 Model 1: Equilibrium Stage Model

NOTE:
This figure is included on page 62 of the print copy of
the thesis held in the University of Adelaide Library.

Figure 4.1.1: Equilibrium stage for Model 1 (adapted from Thiele (2007)).

The model equations for Model 1, the equilibrium stage model with chemical equilibrium, are:

Material balances: Equilibrium stage MESH equation (2.2.1)


Equilibrium relations: Equilibrium stage MESH equation (2.2.2)
Summation equations: Equilibrium stage MESH equation (2.2.3)
Enthalpy balances: Equilibrium stage MESH equation (2.2.4)
Adiabatic condition Qi = 0
Mass transfer relations: None required
Energy transfer relations: None required
Reaction kinetics: None required
Phase equilibria relations: Vapour-liquid equilibrium expressions (2.1.22) to (2.1.24) and (2.1.26)
Speciation relations: Chemical equilibrium relations (2.1.11) to (2.1.15)
Liquid phase relations (2.1.16) to (2.1.21) and (2.1.25)
Pressure drop relations: Pressure drop equations (2.2.8) and (2.2.44)
Hydrodynamic equations (2.2.38) to (2.2.42)
Static vapour head equation (2.2.43)

62
Chapter 4: Aspen Custom Modeler® Process Model Development

4.1.1.2 Model 2: Non-Equilibrium Stage Model with Enhancement Factors

NOTE:
This figure is included on page 63 of the print copy of
the thesis held in the University of Adelaide Library.

Figure 4.1.2: Non-equilibrium stage for Model 2 (adapted from Thiele (2007)).

The model equations for Model 2, the non-equilibrium rate-based stage model with enhancement
factors, are:

Material balances: Non-equilibrium stage MERQ equation (2.2.1)


Equilibrium relations: Non-equilibrium stage MERQ equation (2.2.2)
Summation equations: Non-equilibrium stage MERQ equation (2.2.3)
Enthalpy balances: Non-equilibrium stage MERQ equation (2.2.4)
Adiabatic condition QGi = QLi = 0
Mass transfer relations: Non-equilibrium stage MERQ equations (2.2.12) and (2.2.28)
Mass transfer correlations (2.2.31) to (2.2.33)
Energy transfer relations: Non-equilibrium stage MERQ equations (2.2.20) and (2.2.21)
Energy transfer correlations (2.2.35) to (2.2.37)
Reaction kinetics: Enhancement factor expression (2.2.29)
Reaction rate constant expression (2.1.3)
Phase equilibria relations: Vapour-liquid equilibrium expressions (2.1.22) to (2.1.24) and (2.1.26)
Speciation relations: Chemical equilibrium relations (2.1.11) to (2.1.15)
Liquid phase relations (2.1.16) to (2.1.21) and (2.1.25)
Pressure drop relations: Pressure drop equations (2.2.8) and (2.2.44)
Hydrodynamic equations (2.2.38) to (2.2.42)
Static vapour head equation (2.2.43)

63
Chapter 4: Aspen Custom Modeler® Process Model Development

4.1.1.3 Model 3: Non-Equilibrium Stage Model with Reaction Rate Expressions

NOTE:
This figure is included on page 64 of the print copy of
the thesis held in the University of Adelaide Library.

Figure 4.1.3: Non-equilibrium stage for Model 3 (adapted from Thiele (2007)).

The model equations for Model 3, the non-equilibrium rateb-based stage model with reaction rate
expressions, are:

Material balances: Non-equilibrium stage MERQ equations (2.2.1) and (2.2.26)


Equilibrium relations: Non-equilibrium stage MERQ equation (2.2.2)
Summation equations: Non-equilibrium stage MERQ equation (2.2.3)
Enthalpy balances: Non-equilibrium stage MERQ equation (2.2.4)
Adiabatic condition QGi = QLi = 0
Mass transfer relations: Non-equilibrium stage MERQ equations (2.2.12) and (2.2.17)
Mass transfer correlations (2.2.31) to (2.2.33)
Energy transfer relations: Non-equilibrium stage MERQ equations (2.2.20) and (2.2.21)
Energy transfer correlations (2.2.35) to (2.2.37)
Reaction kinetics: Reaction rate equation (2.2.22)
Chemical equilibrium constant expression (2.2.23)
Reaction rate constant expression (2.1.3)
Phase equilibria relations: Vapour-liquid equilibrium expressions (2.1.22) to (2.1.24) and (2.1.26)
Speciation relations: Chemical equilibrium relations (2.1.11) to (2.1.15)
Liquid phase relations (2.1.16) to (2.1.21) and (2.1.25)
Pressure drop relations: Pressure drop equations (2.2.8) and (2.2.44)
Hydrodynamic equations (2.2.38) to (2.2.42)
Static vapour head equation (2.2.43)

As suggested by Schneider and Górak (2001), the liquid film reactions were represented in Model 3 by
assuming a linear liquid film concentration profile and applying the averaged reaction kinetics. This
simplification eliminated the need to discretise the liquid film, thereby reducing the number of model
equations and hence, the required computation time.

64
Chapter 4: Aspen Custom Modeler® Process Model Development

4.1.2 The Regenerator Wash Section Model Equations


Besides the packed beds that are represented by the above model equations, the regenerator
columns also contain wash sections, as illustrated in Figure 2.1.2. In CO2 trains #5 to #7, the wash
sections are comprised of two valve trays. In contrast, the wash sections in CO2 trains #1 to #4
consist of a cooling water circuit in which the vapour entering the packed bed in the wash section is
cooled and partially condensed by a constant recirculating flow of water.

To simplify the Aspen Custom Modeler® simulations, the valve tray wash sections were modelled as
single equilibrium stages with a pressure drop of 1.0 kPa 1, while the more complex cooling water
circuits were treated as full reflux condensers with a pressure drop of 1.38 kPa 2 and water as the duty
fluid. Following the example of Krishnamurthy and Taylor (1985), these units were also represented
by single equilibrium stages. Consequently, the general wash section model equations are similar to
those for Model 1:

Material balances: Equilibrium stage MESH equation (2.2.1)


Equilibrium relations: Equilibrium stage MESH equation (2.2.2)
Summation equations: Equilibrium stage MESH equation (2.2.3)
Enthalpy balances: Equilibrium stage MESH equation (2.2.4)
Mass transfer relations: None required
Energy transfer relations: None required
Reaction kinetics: None required
Phase equilibria relations: Vapour-liquid equilibrium expressions (2.1.22) to (2.1.24) and (2.1.26)
Speciation relations: Chemical equilibrium relations (2.1.11) to (2.1.15)
Liquid phase relations (2.1.16) to (2.1.21) and (2.1.25)
Pressure drop relations: Pressure drop equation (2.2.8)

For the valve tray-type wash sections, there is no energy input so the adiabatic condition Qi = 0 was
applied. In the case of the cooling water circuits, the cooler duties QCooler were related to the water
flow rate Fwater and the water enthalpies hwater,in and hwater,out at the inlet and outlet conditions:
Q Cooler Fwater ˜ h water,out  h water ,in (4.1.1)

and the outlet water temperature was set equal to the temperature of the liquid stream leaving the
wash section.

1
This equates to a pressure drop of 4 inches of H2O per valve tray, as recommended by Kister (1992).
2
This equates to a pressure drop of 0.5 – 0.6 inches of H2O per foot of wash section packing, as recommended
by Kister (1992).

65
Chapter 4: Aspen Custom Modeler® Process Model Development

4.1.3 The Reboiler and Condenser Model Equations


The remaining major components of the regenerator columns are the solution reboilers, the overhead
condensers and the overhead catchpots. Like the regenerator wash sections, the solution reboilers
for the regenerator columns were treated as equilibrium stages, as recommended by Krishnamurthy
and Taylor (1985). Similarly, the overhead condenser and overhead catchpot for each regenerator
column were modelled as a combined equilibrium stage. As a result, the reboiler and condenser
model equations are similar to those for the regenerator wash sections:

Material balances: Equilibrium stage MESH equation (2.2.1)


Equilibrium relations: Equilibrium stage MESH equation (2.2.2)
Summation equations: Equilibrium stage MESH equation (2.2.3)
Enthalpy balances: Equilibrium stage MESH equation (2.2.4)
Mass transfer relations: None required
Energy transfer relations: None required
Reaction kinetics: None required
Phase equilibria relations: Vapour-liquid equilibrium expressions (2.1.22) to (2.1.24) and (2.1.26)
Speciation relations: Chemical equilibrium relations (2.1.11) to (2.1.15)
Liquid phase relations (2.1.16) to (2.1.21) and (2.1.25)
Pressure drop relations: Pressure drop equation (2.2.8)

The duty fluid for the solution reboilers is low pressure steam, which completely condenses in the
reboilers. Given the conservation of energy and mass, the reboiler duties QReb were therefore
calculated from:

Q Re b Fsteam ˜ h condensate  h steam Fsteam ˜ 'h HvapO (4.1.2)


2

where Fsteam is the steam flow rate, hcondensate and hsteam are the condensate and steam enthalpies, and
'hHvapO is the enthalpy of vaporisation of water. For the condensers, the duty fluid is air and the
2

condenser duties QCond were related to the air flow rate Fair and the enthalpies of air hair,in and hair,out at
the inlet and outlet conditions, respectively:
Q Cond Fair ˜ h air,out  h air,in (4.1.3)

The pressure drops 'P across the reboilers were set equal to the pressure drops due to the static
liquid head 'Pliq in the reboilers:
'P 'Pliq U L ˜ g ˜ Hliq (4.1.4)

66
Chapter 4: Aspen Custom Modeler® Process Model Development

where Hliq is the liquid height; while the pressure drops 'P across the combined condenser and
catchpot stages were determined from the condenser pressure drops 'PCond (which were set at 13.8
kPa 1) and the pressure drops due to the static liquid head 'Pliq in the overhead catchpots:
'P 'PCond  'Pliq 'PCond  U L ˜ g ˜ Hliq (4.1.5)

The effect of the static vapour head was neglected (Strigle, 1994) due to the low operating pressures
for the reboilers and condensers.

4.1.4 Other Ancillary Unit Model Equations


To simplify the Aspen Custom Modeler® simulations, the pressure drop between the absorber and
regenerator columns was represented by a valve model while a pump model was used to represent
the pressure increase between the two columns. These two ancillary units were treated as equilibrium
stages with the following general model equations:

Material balances: Equilibrium stage MESH equation (2.2.1)


Equilibrium relations: Equilibrium stage MESH equation (2.2.2)
Summation equations: Equilibrium stage MESH equation (2.2.3)
Enthalpy balances: Equilibrium stage MESH equation (2.2.4)
Mass transfer relations: None required
Energy transfer relations: None required
Reaction kinetics: None required
Phase equilibria relations: Vapour-liquid equilibrium expressions (2.1.22) to (2.1.24) and (2.1.26)
Speciation relations: Chemical equilibrium relations (2.1.11) to (2.1.15)
Liquid phase relations (2.1.16) to (2.1.21) and (2.1.25)
Pressure drop relations: Pressure drop equation (2.2.8)

For the valve model, the pressure drop was assumed to occur adiabatically and the valve duty QValve
was set as zero. Since some of the dissolved CO2 and H2S are flashed off from the depressurisation,
the full set of equations given above was implemented in the valve model.

In contrast, no phase change takes place in the pump model, so the phase equilibria and speciation
relations were not required and the vapour phase terms in the remaining relations were eliminated.
The pressure rise across the pump was assumed to occur isothermally so the pump outlet
temperature Tout was set equal to the inlet temperature Tin:
Tout Tin (4.1.6)

1
This equates to a pressure drop of 2 psi which is typical of air cooler-type condensers (Hudson Products
Corporation, 2002).

67
Chapter 4: Aspen Custom Modeler® Process Model Development

4.1.5 Numerical Solution


In order to solve the mathematical absorber and regenerator column models developed for the CO2
trains, the above model equations were implemented in Aspen Custom Modeler® using its object-
oriented modelling language. The resulting set of differential and algebraic equations were reduced to
a set of non-linear algebraic equations by the elimination of the time derivatives under steady-state
conditions.

Group decomposition was then applied to divide this large set of equations into a series of smaller
equation sub-sets, which were solved more efficiently than if the entire equation set was solved
simultaneously. These equation sub-sets were solved using the slow but very robust non-linear
Newton method, which calculates a new Jacobian matrix at each iteration. On average, the
computation time required to solve the individual steady-state column models varied between 1 to 3
minutes, depending on the model complexity, the column size and the initial values for the model
variables.

The convergence criterion was based on the equation residuals with an absolute equation tolerance of
1×10-5. Convergence was achieved when the residual for each equation was reduced to below the set
tolerance. To facilitate convergence, the model equations were implemented such that Model 1, being
the simplest, was used to generate initial values for Model 2. Likewise, Model 2 was used to initialise
the variables for the more complex Model 3.

68
Chapter 4: Aspen Custom Modeler® Process Model Development

4.2 Preliminary CO2 Train Simulations


Based on the first set of typical operating data provided in Table 2.1.2, a series of preliminary
simulations were conducted to investigate the performance of the individual Aspen Custom Modeler®
absorber and regenerator column models. The corresponding column model configurations are
shown in Figure 4.2.1. The only notable difference between the absorber column models that is not
depicted in the diagram below is the number of packed beds in each column. This varied between
one and three, depending on the CO2 train.

Acid Gas
Flash Steam CONDENSER Acid Gas
CONDENSER

Flash Steam Sour Water


Sweet Gas (Trains #5 & #6)
Sour Water WASH
(Trains #3 & #4) SECTION WASH Makeup Water
SECTION
Lean Solution
ABSORBER PRESSURE
DROP
PRESSURE
DROP

Rich Solution
REGENERATOR
Rich Solution
Raw Gas
REGENERATOR
th
(4 bed not in
Trains #1 &#2)
Makeup Water
Flash Steam
REBOILER
Rich Solution (Train #7)
REBOILER

Regenerated Solution
Regenerated Solution

(a) (b) (c)


Figure 4.2.1: Preliminary Aspen Custom Modeler® simulation column configurations. (a) The absorber
model for Trains #1 to #7. (b) The regenerator model for Trains #1 to #4. (c) The regenerator model for
Trains #5 to #7.

4.2.1 Number of Model Stages


In order to apply the model equations to the CO2 trains, the packed heights of the absorber and
regenerator columns had to be first discretised into a number of stages. To minimise the number of
model equations, and therefore the computation time, a series of simulation runs using Models 2 and
3 were performed on the absorber and regenerator in CO2 train #1 (the absorber and regenerator with
the shortest packed heights) to evaluate the discretisation of these two columns into different numbers
of stages.

The predicted sweet gas CO2 and H2S content and the overall predicted CO2 and H2S absorption
rates for the absorber are shown in Figure 4.2.2. Figure 4.2.3 compares the predicted regenerated
solution CO2 and H2S loading and the overall predicted CO2 and H2S desorption rates for the
regenerator.

For the absorber, negligible differences were observed between the values corresponding to 15
stages and the final asymptotic values. Similarly, there were negligible differences between the

69
Chapter 4: Aspen Custom Modeler® Process Model Development

values corresponding to 20 stages and the final asymptotic values for the regenerator. The optimal
number of stages was therefore taken to be 15 for the absorber and 20 for the regenerator, which
correspond to absorber and regenerator stage heights of approximately 0.49 and 0.96 m, respectively.
These optimal stage heights were then used to discretise the packed heights of the taller absorber and
regenerator columns in the CO2 trains.

CO2 H2S

6.0 4.0 650 0.080

CO2 Absorption Rate (kmol/h)

H2S Absorption Rate (kmol/h)


5.0 3.2
Sweet Gas CO2 (mol%)

Sweet Gas H2S (ppm)


600 0.076

4.0 2.4
550 0.072
3.0 1.6
500 0.068
2.0 0.8

450 0.064
1.0 0.0

0.0 -0.8 400 0.060


0 10 20 30 0 10 20 30

Number of Stages Number of Stages

Figure 4.2.2: Results of the absorber discretisation simulation runs for CO2 train #1. (a) Sweet gas
composition and (b) total absorption rates as a function of the number of stages.

CO2 H2S

0.46 3.6 620 0.074


CO2 Desorption Rate (kmol/h)

H2S Desorption Rate (kmol/h)


600 0.072
H2S Loading (×106)

0.44 2.4
CO2 Loading

580 0.070
0.42 1.2
560 0.068

0.40 0.0
540 0.066

0.38 -1.2 520 0.064


0 10 20 30 0 10 20 30

Number of Stages Number of Stages

(a) (b)
Figure 4.2.3: Results of the regenerator discretisation simulation runs for CO2 train #1. (a) CO2 and H2S
loading in the regenerated solution and (b) total desorption rates as a function of the number of stages.
Note: The CO2 and H2S loadings are not strictly dimensionless; their units are actually (mol CO2)·(mol
-1 -1
K2CO3) and (mol H2S)·(mol K2CO3) .

70
Chapter 4: Aspen Custom Modeler® Process Model Development

4.2.2 The Different Modelling Approaches


Once the optimal stage heights were established, a series of simulations were run to compare the
performances of Models 1, 2 and 3 for each absorber and regenerator, in order to determine the
appropriate model complexity.

The CO2 and H2S vapour phase profiles predicted by the three different mathematical models for the
absorbers in CO2 trains #1 and #7 are compared in Figure 4.2.4, which also compares the liquid
phase CO2 and H2S loading profiles predicted for the corresponding regenerators. The absorber and
regenerator profiles for the other five CO2 trains were very similar in shape to those for CO2 train #1,
and are included in Figures D.1.1 and D.1.2 in Appendix D.

The general shapes of the absorber and regenerator profiles were as anticipated. The vapour phase
CO2 and H2S concentrations decreased up the absorber columns, while the liquid phase CO2 and H2S
loading decreased down the regenerator columns. A slight initial increase in the liquid phase CO2
loading was observed at the top of the regenerator columns. This phenomenon was caused by
contact between the flashed-off CO2 entering the columns and the condensed steam leaving the wash
sections. A portion of the CO2 was absorbed by the condensed steam, resulting in the observed
increase in CO2 loading at the top of the regenerator columns.

As expected, the absorber CO2 and H2S vapour phase profiles obtained from the equilibrium stage
model (Model 1) differed considerably from those produced by the two non-equilibrium stage models
(Models 2 and 3), which agreed closely. Similarly, much closer agreement was observed between the
regenerator loading profiles predicted by Models 2 and 3 than with those predicted by Model 1.
However, all three models did predict very similar compositions for the sweet gas leaving the top of
the absorbers, as well as very similar acid gas loadings for the solution exiting the bottom of the
regenerators. Despite this agreement with the more complex models, Model 1 was considered to be
only suitable for initialisation runs due to the significant deviations observed for the absorber profiles.

As the simplest mathematical model, it was not surprising that Model 1 had the shortest computation
times. The computation times associated with Model 2 were typically between 50 to 150% greater
than that for Model 1, and similar increments in computation time were observed for Model 3
compared to Model 2. Despite the increased computation time and model complexity, the close
agreement (within ±2%) between the profiles produced by Models 2 and 3 indicated that no significant
improvement in the simulation results was achieved by using rigorous rate expressions instead of
enhancement factors. This confirmed the earlier findings of de Leye and Froment (1986a) and Mayer
(2001). It was therefore decided to employ Model 2 for subsequent simulations.

71
Chapter 4: Aspen Custom Modeler® Process Model Development

Model 1 Model 2 Model 3 Data

1.2
Cond
WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4

0.4
0.2

0.2 0.0
Reb
0.0 -0.2
0 5 10 15 20 0 5 10 15 20 25 0.3 0.4 0.5 0.6 0.7 0 10 20 30

CO2 (mol%) H2S (ppm) CO2 Loading H2S Loading (×106)

(a) (c)
1.2
Cond
WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4
0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 0 5 10 15 20 25 0.3 0.4 0.5 0.6 0.7 0 10 20 30 40

CO2 (mol%) H2S (ppm) CO2 Loading H2S Loading (×106)

(b) (d)
Figure 4.2.4: Results of the different modelling approaches for CO2 trains #1 and #7. Absorber CO2 and
H2S vapour phase profiles for (a) CO2 train #1 and (b) CO2 train #7. Regenerator liquid phase CO2 and H2S
loading profiles for (c) CO2 train #1 and (d) CO2 train #7. (Cond = condenser, WS = wash section, Reb =
reboiler)

72
Chapter 4: Aspen Custom Modeler® Process Model Development

4.3 Column Model Adjustments


It was observed in the previous section that the sweet gas CO2 content predicted by the three
mathematical models for the absorbers in CO2 trains #1 to #6 deviated significantly from the data
values. In contrast, a much smaller difference was observed for the absorber in CO2 train #7. Small
deviations from the data values were also observed for the predicted CO2 loading at the bottom of the
seven regenerators. The following sections investigate the reasons for these deviations and discuss
the model adjustments required to correct them.

4.3.1 Liquid Phase Enthalpy Correction


Given that the chemical absorption and desorption are temperature-dependent processes, the
temperature profiles along the absorber and regenerator columns were examined to determine if the
calculated temperatures had contributed to the above-mentioned deviations. For comparison, the
temperature profiles predicted by Model 2 for CO2 trains #1 and #7 are provided in Figure 4.3.1. The
corresponding temperature profiles for CO2 trains #2 to #6 are given in Figures D.2.1 and D.2.2 in
Appendix D.
Distinctive temperature bulges were observed in the absorber temperature profiles. These were
caused by an initial rise in the solution temperature as it flowed down the columns (due to the
exothermic nature of CO2 absorption), followed by a temperature drop near the bottom of absorbers
when the solution came into contact with the cooler raw gas. While the shapes of the absorber
temperature profiles were as expected, the predicted absorber bottom temperatures were
approximately 8°C higher than the data values. The overly high temperatures in the lower part of the
absorbers were thought to have led to the predicted over-absorption of CO2. However, they were not
considered to be the sole contributory factor since the absorber in CO2 train #7 had the highest
predicted temperature profile but the smallest deviation for the sweet gas CO2 content.

For the regenerators, the liquid phase temperatures were observed to remain relatively constant in the
top two-thirds of the columns and to increase in the lower third of the columns, reaching a maximum at
the reboilers. This correlated well with the corresponding CO2 loading profiles in Figure 4.2.4. The
solution CO2 loading also remained relatively constant in the top two-thirds of the columns before
decreasing to a minimum at the reboilers. These observed phenomena resulted from the rich solution
coming into contact with the hot stripping steam generated from the solution reboilers. The rich
solution experienced a temperature rise while the steam stripped CO2 from the solution.

Close agreement was observed between the vapour and liquid phase temperature profiles within the
regenerator columns while above the regenerator columns, the vapour phase temperatures decreased
to a minimum at the condensers. In contrast to the absorbers, there was closer agreement between
the temperatures predicted for the regenerators and the data values (within ±3°C). This suggested the
regenerator temperatures were not the reason for the deviation between the predicted CO2 loading for
the regenerated solution and the data values.

73
Chapter 4: Aspen Custom Modeler® Process Model Development

Vapour Phase Liquid Phase Vapour Data Liquid Data

Cond
1.2 1.2
Cond
WS WS

Fraction of Total Packed Height


Fraction of Total Packed Height
1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
30 50 70 90 110 130 60 70 80 90 100 110 120 80 90 100 110 120 130 140 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(a) (b)
Figure 4.3.1: Temperature profiles for CO2 trains #1 and #7. Temperature profiles predicted by Model 2 for
(a) CO2 train #1 and (b) CO2 train #7. (Cond = condenser, WS = wash section, Reb = reboiler)

The column temperatures were determined from the enthalpy balances, making the temperatures
dependent on the enthalpy values calculated from the thermodynamic property models. The vapour
phase enthalpies were calculated from the SRK equation of state, which has been shown to give very
good predictions for vapour phase enthalpies for mixtures of hydrocarbons and light gases (Poling et
al., 2001).

The liquid phase enthalpies were calculated using the Electrolyte NRTL model and were based on an
approximation (equation (B.1.7)) which neglects the effect of pressure on the liquid phase partial molar
enthalpy. The close agreement between the predicted temperatures and the data values for the
regenerator confirmed the validity of this approximation for atmospheric pressures. However, the
disparity observed for the absorber temperatures indicated that this approximation was not valid at the
much higher absorber operating pressures. Consequently, the effect of adjustments to the liquid
phase enthalpy on the vapour and liquid phase temperatures and the CO2 absorption rate in the seven
absorbers was investigated.

A series of absorber simulation runs were performed, in which the liquid phase enthalpy hL was
corrected via the following linear expression:
hL a ˜ hENRTL
L b (4.3.1)

where a and b are adjustable constants, and hENRTL


L is the liquid phase enthalpy calculated from the
Electrolyte NRTL model. It was observed that as the liquid phase enthalpy became more negative,
the absorber temperatures decreased, along with the CO2 and H2S absorption rates, albeit
insignificantly.

The values for the constants a and b were adjusted iteratively to achieve the best agreement with the
design temperatures for the seven absorbers. The resulting values of 0.7201 and -0.0933 for a and b,

74
Chapter 4: Aspen Custom Modeler® Process Model Development

respectively, were found to give an absolute average deviation of 0.5% and a maximum absolute
deviation of 0.9%. The corrected and uncorrected absorber temperature profiles for CO2 trains #1 and
#7 are compared in Figure 4.3.2. The corresponding temperature profiles for the remaining five CO2
trains are given in Figures D.2.1 and D.2.2 in Appendix D. The considerable improvement observed
for the corrected absorber temperature profiles indicated that the simple linear form for equation
(4.3.1) was sufficient and the use of a more complex expression was unnecessary.

Unlike the absorber temperatures, the CO2 absorption rate was not significantly affected by the liquid
phase enthalpy correction since the resulting CO2 vapour phase profiles closely resembled those
obtained with the uncorrected liquid phase enthalpies, as seen in Figure 4.3.2. The H2S absorption
rate and vapour phase profiles were similarly unaffected by the liquid phase enthalpy correction.
Consequently, as observed for the regenerators, the calculated temperatures did not contribute to the
deviations in the CO2 vapour phase profiles observed in the absorbers. This suggested that a model
parameter sensitive to CO2 mass transfer was instead the cause for the observed deviations in the
absorber CO2 vapour phase profiles and the regenerator CO2 loading profiles. The sensitivity of the
absorber and regenerator column models to the key mass transfer parameters is investigated in the
next section.

Uncorrected enthalpy Corrected enthalpy Data


1.0
1.0
Fraction of Total Packed Height

Fraction of Total Packed Height

0.8
0.8

0.6
0.6

0.4
0.4

0.2
0.2

0.0 0.0
30 50 70 90 110 130 100 110 120 130 0 5 10 15 20 0 5 10 15 20 25
Vapour Phase Temperature (°C) Liquid Phase Temperature (°C) CO2 (mol%) H2S (ppm)

(a) (c)
1.0
1.0
Fraction of Total Packed Height

Fraction of Total Packed Height

0.8
0.8

0.6
0.6

0.4
0.4

0.2
0.2

0.0 0.0
90 100 110 120 130 140 100 110 120 130 140 0 5 10 15 20 0 5 10 15 20 25
Vapour Phase Temperature (°C) Liquid Phase Temperature (°C) CO2 (mol%) H2S (ppm)

(b) (d)
Figure 4.3.2: Effect of the liquid phase enthalpy correction on the absorber profiles. Temperature profiles
for (a) CO2 train #1 and (b) CO2 train #7. CO2 and H2S vapour phase profiles for (c) CO2 train #1 and (d)
CO2 train #7.

75
Chapter 4: Aspen Custom Modeler® Process Model Development

4.3.2 Sensitivity to Model Parameters


The four key model parameters that are sensitive to CO2 mass transfer are the effective interfacial
area aI, the vapour phase mass transfer coefficient kG, the liquid phase mass transfer coefficient kL,
and the reaction rate constant k OH . These were calculated using the correlations presented by Onda

and co-workers (1968ab) (equations (2.2.31) to (2.2.33)) and Astarita and co-workers (1983) (equation
(2.1.3)).

To determine the variation associated with the parameter values, the above correlations were
compared against the alternative mass transfer and reaction rate constant correlations given in
Section 2.1.3 and Table D.2.1 in Appendix D, using the average CO2 train phase properties. The
resulting parameter values and corresponding percentage deviations are given in Table 4.3.1.

Given the wide variations in the model parameter values observed in Table 4.3.1, it was considered
highly likely that the observed deviations in the column composition profiles were caused by the errors
associated with calculated parameter values. To investigate the sensitivity of the two column models
to the variations in these four model parameters, a series of simulation runs were performed for the
absorbers and regenerators in CO2 trains #1 and #7, in which the model parameters were changed by
±50%. The absorbers in CO2 trains #2 to #6 were not included in the sensitivity studies since they
were observed to behave quite similarly to the absorber in CO2 train #1 from their temperature and
composition profiles. The regenerators in these five trains were likewise excluded due to their similar
behaviour to the regenerators in CO2 trains #1 and #7.

To assess the sensitivity of the absorber column model to the four model parameters, the relative
changes in the total CO2 and H2S absorption rates and the sweet gas CO2 and H2S content were
compared, as depicted in Figure 4.3.3. From these plots, it was evident that the absorber column
model was most sensitive to the effective interfacial area. The changes in the acid gas absorption
rates and sweet gas content associated with the effective interfacial area were significantly greater
than the corresponding changes for the reaction rate constant and the vapour and liquid phase mass
transfer coefficients. These findings agree with the results presented by Schneider and Górak (2001),
who performed a similar sensitivity analysis on their non-equilibrium rate-based absorber model for an
amine gas treatment process.

The magnitude and direction of the changes in the sweet gas CO2 content indicated that the predicted
over-absorption of CO2 observed in Figure 4.2.4 could be corrected by the introduction of an
adjustment factor to reduce the effective interfacial area. Given the different order of magnitude for
these changes for the absorbers in CO2 trains #1 and #7, it was expected that this factor would not be
constant for the seven CO2 trains, but would instead have to be fitted for each absorber. The inclusion
of the effective interfacial area adjustment factor was the second model adjustment applied to the
absorber column model (the first being the liquid phase enthalpy correction term), and is discussed in
the next section.

76
Chapter 4: Aspen Custom Modeler® Process Model Development
Table 4.3.1: The average variation associated with the model parameter values for the CO2 train absorbers and regenerators. The bracketed values are the
percentage deviations from the correlations used in this work.

CO2 Train Absorbers CO2 Train Regenerators


#1 #2 #3 #4 #5 #6 #7 #1 #2 #3 #4 #5 #6 #7
2 3
aI (m /m )
Onda and co- 97 83 99 100 59 59 58 89 84 135 138 79 79 135
workers (1968ab) a
Puranik and 113 98 121 129 87 76 79 93 89 137 144 87 86 139
Vogelpohl (1974) (16%) (18%) (22%) (29%) (47%) (47%) (35%) (4%) (5%) (2%) (5%) (9%) (9%) (3%)
Kolev (1976) 52 44 57 61 50 50 44 42 31 51 54 38 38 52
(-47%) (-47%) (-43%) (-39%) (-15%) (-15%) (-24%) (-53%) (-63%) (-62%) (-61%) (-52%) (-52%) (-62%)
Billet and Schultes 159 161 174 190 209 207 178 132 134 132 140 142 142 134
(1999) (63%) (94%) (77%) (90%) (255%) (252%) (206%) (49%) (58%) (-2%) (2%) (79%) (80%) (-1%)
2
k G,CO (cm /s)
2

Onda and co- 0.08 0.11 0.12 0.13 0.16 0.19 0.20 2.41 2.61 1.40 1.58 2.36 2.43 1.59
77

workers (1968ab) a
Van Krevelen and 0.05 0.05 0.07 0.07 0.08 0.10 0.10 0.86 0.87 0.67 0.77 0.89 0.92 0.78
Hoftijzer (1948) (-46%) (-57%) (-43%) (-42%) (-50%) (-49%) (-49%) (-64%) (-67%) (-52%) (-51%) (-62%) (-62%) (-51%)
Shulman and co- 0.28 0.23 0.38 0.42 0.34 0.39 0.40 5.24 5.03 6.25 7.03 5.43 5.56 7.04
workers (1955) (232%) (112%) (223%) (220%) (107%) (104%) (103%) (118%) (93%) (348%) (343%) (130%) (129%) (342%)
2
k L,CO (cm /s)
2

Onda and co- 0.20 0.20 0.23 0.26 0.39 0.40 0.36 0.10 0.11 0.09 0.10 0.13 0.13 0.09
a
workers (1968ab)
Van Krevelen and 0.14 0.16 0.17 0.19 0.30 0.31 0.29 0.07 0.08 0.06 0.07 0.10 0.10 0.06
Hoftijzer (1948) (-28%) (-22%) (-28%) (-28%) (-23%) (-22%) (-21%) (-26%) (-24%) (-35%) (-35%) (-26%) (-27%) (-36%)
Shulman and co- 0.57 0.62 0.64 0.70 0.70 0.70 0.64 0.40 0.55 0.46 0.49 0.48 0.47 0.46
workers (1955) (188%) (206%) (174%) (165%) (78%) (76%) (78%) (290%) (420%) (400%) (389%) (260%) (265%) (405%)
6 3
k OH (×10 m /kmol·s)
Astarita and co- 3.4 3.4 3.8 3.9 3.8 3.8 3.5 3.4 3.4 3.7 3.7 3.8 3.8 3.7
a
workers (1983)
Pohorecki and 2.4 2.4 2.8 2.9 3.0 2.9 2.4 380 390 43 43 45 44 44
Moniuk (1988) (-29%) (-28%) (-25%) (-26%) (-21%) (-24%) (-30%) (-99%) (-99%) (-88%) (-88%) (-88%) (-88%) (-88%)
a
These are the mass transfer and reaction rate constant correlations used in the Aspen Custom Modeler® column models.
Chapter 4: Aspen Custom Modeler® Process Model Development

To evaluate the sensitivity of the regenerator column model to the model parameters, the relative
changes in the total CO2 and H2S desorption rates and the regenerated solution CO2 and H2S loading
were compared in Figure 4.3.4. It should be noted that the similarity between the sensitivity analysis
results for the regenerators in CO2 trains #1 and #7 was not surprising since the two regenerators
were expected to behave very similarly, given their temperature and CO2 loading profiles. As for the
absorber column model, the effective interfacial area was found to be the most significant model
parameter for the regenerator column model.

However, in contrast to the absorber column model, the relative changes for all four model parameters
were observed to be insignificant, especially those associated with the H2S desorption rate, which had
order of magnitudes around 10-5 to 10-7 and are not clearly depicted in Figure 4.3.4. Although the
relative changes in the regenerated solution H2S loading were large, the actual H2S loading was
extremely low (around the order of 10-13), and therefore changes of a few orders of magnitude to the
H2S loading were still relatively inconsequential.

These insignificant relative changes indicated that an effective interfacial area adjustment factor for
the regenerator column model would be inadequate for correcting the deviations observed in Figure
4.2.4 for the regenerated solution CO2 loading. Adjustment of the steam flow to the solution reboilers
was considered more appropriate, given the regenerator column model’s significant sensitivity to this
key operating parameter. This is illustrated in Figure 4.3.5, which compares the relative changes in
the CO2 desorption rate and regenerated solution loading for changes of ±15% to the reboiler steam
flow. Consequently, unlike the absorber column model, no adjustment factors or correction terms
were applied to the regenerator column model.

It should be noted from Figure 4.3.3 that the changes in the liquid phase mass transfer coefficient had
a stronger effect on the CO2 absorption rate and the sweet gas CO2 content than the corresponding
changes in the vapour phase mass transfer coefficient. In contrast, Figure 4.3.4 shows the vapour
and liquid phase mass transfer coefficients having similar levels of influence on the CO2 desorption
rate and the regenerated solution CO2 loading. This indicated that the main resistance to CO2 mass
transfer lay in the liquid phase for the CO2 train absorbers, whereas in the regenerators, neither phase
resistance dominated.

Figures 4.3.3 and 4.3.4 also indicate that H2S mass transfer was vapour-phase limited in both the
absorbers and regenerators. The vapour phase mass transfer coefficient was observed to have a
stronger effect on the H2S absorption rate and the sweet gas content compared to the liquid phase
mass transfer coefficient. A similar trend for the H2S desorption rate and the solution loading was
observed for the regenerators. These results regarding the CO2 and H2S mass transfer limitations are
in agreement with the empirical findings of Yih and Lai (1987), thereby validating the mass transfer
relations utilised in this work.

78
Chapter 4: Aspen Custom Modeler® Process Model Development

CO2
CO2 -(CO
Absorber
2 train #1)
1 CO2
CO2 -(CO
Absorber
2 train #7)
7 H2S - Absorber
H2S (CO2 train #1)
1 H2S - Absorber
H2S (CO2 train #7)
7

k OH-

Parameter
kL

kG

aI

-0.05 0.00 0.05 0.10 0.15 0.20 0.25


Relative Change in Total Absorption Rate (%)

(a)

k OH-
Parameter

kL

kG

aI

-3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5


Relative Change in Total Absorption Rate (%)

(b)

CO2
+50%- (CO
Absorber 1
2 train #1)
CO2
+50%- (CO
Absorber 7
2 train #7)
H2S
-50%- (CO
Absorber 1
2 train #1)
H2S
-50%- (CO
Absorber 7
2 train #7)

k OH-
Parameter

kL

kG

aI

-50 0 50 100 150 200


Relative Change in Sweet Gas CO2 Content (%)

(c)
k OH-
Parameter

kL

kG

aI

-500 0 500 1000 1500 2000 2500


Relative Change in Sweet Gas H2S Content (%)

(d)
Figure 4.3.3: Sensitivity analysis results for the absorber column model (Model 2). Relative changes in
the total CO2 and H2S absorption rates for the CO2 trains #1 and #7 after (a) a 50% increase and (b) a 50%
decrease in the model parameters, and the corresponding changes in (c) the sweet gas CO2 content and
(d) the sweet gas H2S content.

79
Chapter 4: Aspen Custom Modeler® Process Model Development

CO2
CO2 -(CO
Absorber
2 train #1)
1 CO2
CO2 -(CO
Absorber
2 train #7)
7 H2S - Absorber
H2S (CO2 train #1)
1 H2S - Absorber
H2S (CO2 train #7)
7

k OH-

Parameter
kL

kG

aI

-0.006 -0.004 -0.002 0.000 0.002 0.004 0.006 0.008 0.010 0.012 0.014
Relative Change in Total Desorption Rate (%)

(a)
k OH-
Parameter

kL

kG

aI

-0.030 -0.025 -0.020 -0.015 -0.010 -0.005 0.000 0.005 0.010 0.015
Relative Change in Total Desorption Rate (%)

(b)

CO2
+50%- (CO
Absorber 1
2 train #1)
CO2
+50%- (CO
Absorber 7
2 train #7)
H2S
-50%- (CO
Absorber 1
2 train #1)
H2S
-50%- (CO
Absorber 7
2 train #7)

k OH-
Parameter

kL

kG

aI

-0.020 -0.015 -0.010 -0.005 0.000 0.005 0.010 0.015 0.020 0.025 0.030
Relative Change in Regenerated Solution CO2 Loading (%)

(c)
k OH-
Parameter

kL

kG

aI

-800 -600 -400 -200 0 200 400 600 800 1000 1200 1400
Relative Change in Regenerated Solution H2S Loading (%)

(d)
Figure 4.3.4: Sensitivity analysis results for the regenerator column model (Model 2). Relative changes in
the total CO2 and H2S desorption rates for the CO2 trains #1 and #7 after (a) a 50% increase and (b) a 50%
decrease in the model parameters, and the corresponding changes in (c) the regenerated solution CO2
loading and (d) the regenerated solution H2S loading.

80
Chapter 4: Aspen Custom Modeler® Process Model Development

CO2
CO2desorption
desorptionrate
rate(co2
(CO2train
train#1)
#1) solution co2 loading
CO2 desorption rate(co2
(COtrain #1)
2 train #7)
Solutionco2
solution COloading
2 loading (COtrain
(co2 2 train #1)
#1) Solution
co2 CO2 loading
desorption (COtrain
rate (co2 2 train
#7)#7)

15%

Steam Flow

-15%

-15 -10 -5 0 5 10 15
Relative Change (%)

Figure 4.3.5: Effect of the solution reboiler steam flow on the regenerator column model (Model 2).
Relative changes in the total CO2 desorption rate and the regenerated solution CO2 loading for changes of
±15% to the steam flow.

4.3.3 Effective Interfacial Area Adjustment Factor


In the preceding sections, it was determined that a model parameter sensitive to CO2 mass transfer
was the most likely cause for the sweet gas CO2 content deviations observed in Figure 4.2.4. Of
these model parameters, the effective interfacial area was identified as having the most significant
effect on the rate of CO2 mass transfer for the absorber column model.

Consequently, it was decided to apply an adjustment factor to the effective interfacial area in order to
fit the absorber column model to the first set of CO2 train data in Table 2.1.2. This follows a similar
approach taken by Mayer (2002) to correct his non-equilibrium rate-based absorber model for an
amine gas treatment process. Like Mayer (2002), who selected his effective interfacial area
adjustment factor such that his model more closely agreed with his experimental results, the
adjustment factors were adjusted iteratively in this work until optimal agreement was achieved
between the predicted sweet gas CO2 content and the data values.

Table 4.3.2 lists the individual adjustment factors obtained for each absorber, and compares the
deviations between the predicted and data values for the sweet gas CO2 content for the adjusted and
unadjusted absorber column models. The effect of the adjustment factor on the CO2 and H2S vapour
phase profiles for the absorbers in CO2 trains #1 and #7 is shown in Figure 4.3.6. The corresponding
vapour profiles for the other five CO2 trains are included in Appendix D as Figure D.2.3.

A marked improvement in the absorber column model performance was evident for the first six CO2
trains through the inclusion of the adjustment factors. A less obvious improvement in the predicted
sweet gas CO2 content was observed for CO2 train #7 due to the smaller deviation between the
unadjusted model’s results and the data. However, notable changes in the shape of CO2 and H2S
vapour phase profiles were observed after the inclusion of the adjustment factor.

81
Chapter 4: Aspen Custom Modeler® Process Model Development

Table 4.3.2: Effective interfacial area adjustment factor values and their effect on the CO2 train absorbers.

Effective Interfacial Area Sweet Gas CO2 Content Deviation (%)


Absorber
Adjustment Factor Unadjusted Model Adjusted Model
#1 0.300 84.95 0.21
#2 0.218 86.47 0.31
#3 0.312 83.34 0.14
#4 0.257 85.00 -0.05
#5 0.394 80.33 0.04
#6 0.295 84.74 -0.31
#7 0.525 4.27 -0.08

Unadjusted Model Adjusted Model Data

1.0 1.0

Fraction of Total Packed Height


Fraction of Total Packed Height

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
0 5 10 15 20 0 5 10 15 20 25 0 5 10 15 20 0 5 10 15 20 25

CO2 (mol%) H2S (ppm) CO2 (mol%) H2S (ppm)

(a) (b)
Figure 4.3.6: Effect of the effective interfacial area adjustment factor on the absorber CO2 and H2S vapour
phase profiles. (a) CO2 train #1. (b) CO2 train #7.

82
Chapter 4: Aspen Custom Modeler® Process Model Development

4.4 CO2 Train Model Validation


Using the model adjustments discussed in the previous sections, simplified models of the Santos
Moomba CO2 trains were constructed in Aspen Custom Modeler®, and their performances were
compared against the two sets of typical operating data in Table 2.1.2. It should be noted that the first
data set was used in the development of the absorber and regenerator column models, while the
second set of data was used solely for model validation purposes.

Each CO2 train model was configured according to the flow sheets given in Figure 4.4.1, which are
simplified versions of the detailed process flow diagrams in Figure 2.1.2. The only notable difference
between the models that is not shown in the diagram below is the number of packed beds in the
absorber columns, which varied between one and three, depending on the CO2 train.

Acid Gas Acid Gas


Flash Steam (Trains #5 &#6)
Flash Steam
CONDENSER CONDENSER

WASH
Sour Water (Trains #3 & #4) SECTION
WASH
SECTION Sweet Gas
Sweet Gas
Makeup Water

ABSORBER
ABSORBER REGENERATOR

PRESSURE
PRESSURE DROP
DROP
Raw Gas
REGENERATOR
th Flash Steam
(4 bed not in Raw Gas (Train #7)
Trains #1 &#2)

Makeup Water Rich Solution


Rich Solution

REBOILER
REBOILER
Lean Solution
Lean Solution PRESSURE
PRESSURE INCREASE
INCREASE

(a) (b)
Figure 4.4.1: Simplified CO2 train configurations for the Aspen Custom Modeler® simulations. (a) CO2
trains #1 to #4. (b) CO2 trains #5 to #7.

The resulting vapour and liquid phase composition profiles for the absorber and regenerator columns
in CO2 trains #1 and #7 are given in Figures 4.4.2 and 4.4.3, and the column temperature profiles are
shown in Figure 4.4.4. The corresponding profiles for the other five CO2 trains are very similar to
those for CO2 train #1, and are included as Figures D.3.1 to D.3.6 in Appendix D. Values for the
operating parameters calculated by the CO2 train models are provided in Tables 4.4.1 and 4.4.2, along
with their percentage deviation from the data values listed in Table 2.1.2.

The absorber composition profiles in Figures 4.4.2 and 4.4.3 showed the CO2 and H2S vapour phase
concentrations decreasing with the vapour flow up the columns while the acid gas loading in the liquid
phase increased down the columns. For CO2 trains #1 to #6, the CO2 and H2S vapour phase profiles
indicated a gradual reduction in the vapour phase acid gas content along the absorbers, with the acid

83
Chapter 4: Aspen Custom Modeler® Process Model Development

gas content decreasing slightly more rapidly towards the bottom of the columns. This corresponded
well with the liquid phase loading profiles, which showed the acid gas loading increasing at a
marginally faster rate near the bottom of the absorbers.

For CO2 train #7, the absorber H2S vapour and liquid phase profiles resembled those for the other six
CO2 trains. In contrast, the absorber CO2 profiles differed quite markedly. In the top and central
regions of the column, the CO2 profiles followed a similar trend to the profiles for the other six CO2
trains. However, in the bottom section, there was a sharp decline in the rates at which the vapour
phase CO2 content decreased and the liquid phase CO2 loading increased, giving rise to sigmoid
curve-like profiles. This phenomenon was caused by the very high CO2 loading in the lower part of the
absorber (compared to the other six absorbers), which significantly reduced the capacity of the
solution to further absorb CO2 from the vapour phase in this region. The rich solution CO2 loading was
much higher for CO2 train #7 than for the other six CO2 trains due to its much larger CO2 removal
capacity per unit volume of solution, as indicated by the greater raw gas to lean solution flow ratio to
the absorber.

In contrast to the absorbers, the CO2 and H2S vapour phase concentrations in the regenerators
increased with the vapour flow up the columns while the acid gas loading in the liquid phase
decreased down the columns. At the top of the regenerator columns, the condensed steam leaving
the wash sections absorbed some of the flashed-off CO2 entering the regenerators, resulting in a slight
increase in the liquid phase CO2 loading and a corresponding decrease in the vapour phase CO2
concentration.

From a comparison of the liquid phase CO2 and H2S loading profiles in Figures 4.4.2 and 4.4.3, it was
evident that a substantial portion of the acid gases absorbed in the rich solution leaving the absorbers
was flashed off during the depressurisation between the absorber and regenerator columns. For all
seven CO2 trains, the remaining H2S in the liquid phase was rapidly removed near the top of the
regenerators, as indicated by the sharp increase in vapour phase H2S content and the corresponding
decrease in liquid phase H2S loading in this region. The opposite behaviour was observed for the
removal of CO2, in that the vapour phase CO2 content rapidly increased near the bottom of the
regenerators while the liquid phase CO2 loading decreased in this region. This indicated that the
removal of the residual CO2 in the solution was greatly facilitated by contact with the hot stripping
steam generated from the solution reboilers.

The absorber temperature profiles provided in Figure 4.4.4 showed distinctive bulges due to the
solution temperature rise from the exothermic absorption reactions and the temperature drop upon
contact between the hot solution and the cooler raw gas near the bottom of the columns. A
comparison of the liquid phase temperatures at the bottom of the absorbers and at the top of the
regenerators indicated that the depressurisation between the two columns was accompanied by a
temperature drop between 7 to 15°C, depending on the rich solution temperature.

84
Chapter 4: Aspen Custom Modeler® Process Model Development

In the regenerators, the solution temperature rapidly increased towards the bottom of the columns due
to contact between the solution and the hot stripping steam from the solution reboilers. Within the
regenerator columns, the vapour phase temperature profiles were observed to closely agree with the
liquid phase profiles. Above the regenerator columns, the vapour temperatures decreased up through
the wash sections, reaching a minimum at the condensers.

Good agreement was observed between the model results and both sets of plant data for all seven
CO2 trains. The calculated and data values for most of the operating parameters were within ±5%.
Larger deviations of up to ±13.5% were observed for the reboiler steam flows, which had to be
adjusted to give the desired regenerated solution CO2 loading.

In the case of parameters for which only a maximum value was specified in Table 2.1.2 (i.e. the sweet
gas H2S content, the rich solution H2S loading and the column pressure drops), the model calculated
values were found to be within the given range. Percentage deviations were not determined for the
lean solution H2S loading since the data values were zero. However, the calculated values were of
the order of 10-12 to 10-19 and were therefore viewed to be effectively zero.

The most significant deviations between the model results and the plant data were associated with the
acid gas temperature for CO2 train #4. For the two sets of plant data, the predicted temperatures were
approximately 30% lower than the data values. However, these deviations were considered to be
acceptable since the data values were unexpectedly high. According to the CO2 train technical
manual (Santos Ltd, 1998), the temperature of the acid gas leaving the regenerator wash section in
CO2 train #4 is typically around 80°C, which corresponds well with the model results.

The agreement between the model results and both sets of plant data was regarded as sufficiently
close that no further model tuning was required to improve the model performances. The CO2 train
models developed in Aspen Custom Modeler® were therefore considered suitable for developing
corresponding CO2 train models in HYSYS® according to the novel approach described in Section
2.4. To obtain the data required to develop the HYSYS® models, a series of simulations were
performed to investigate the effect of various operating parameters on the performance of the Aspen
Custom Modeler® process models. These parametric studies are discussed in the next chapter.

85
Chapter 4: Aspen Custom Modeler® Process Model Development

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
Absorber H2S (ppm) Absorber H2S Loading 1.2
Cond
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05 WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4

0.4
0.2

0.2 0.0
Reb
0.0 -0.2
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 10 30 50 70 90 0.2 0.4 0.6 0.8

Absorber CO2 (mol%) Absorber CO2 Loading Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05 4.E-05
Absorber H2S (ppm) Absorber H2S Loading 1.2
Cond
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05 8.E-05 WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4
0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 10 30 50 70 90 0.2 0.4 0.6 0.8

Absorber CO2 (mol%) Absorber CO2 Loading Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Figure 4.4.2: CO2 and H2S vapour and liquid phase profiles for the first set of plant data. (a) CO2 train #1.
(b) CO2 train #7. (Cond = condenser, WS = wash section, Reb = reboiler)

86
Chapter 4: Aspen Custom Modeler® Process Model Development

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2
S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05
Absorber H2S (ppm) Absorber H2S Loading Cond
1.2
0 5 10 15 20 0.E+00 1.E-05 2.E-05 3.E-05 4.E-05 WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4

0.4
0.2

0.2 0.0
Reb
0.0 -0.2
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 10 30 50 70 90 0.2 0.4 0.6 0.8

Absorber CO2 (mol%) Absorber CO2 Loading Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
0 20 40 60 80 0.E+00 1.E-05 2.E-05 3.E-05
Absorber H2S (ppm) Absorber H2S Loading 1.2
Cond
0 5 10 15 20 0.E+00 2.E-05 4.E-05 6.E-05 WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4
0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 10 30 50 70 90 0.2 0.4 0.6 0.8

Absorber CO2 (mol%) Absorber CO2 Loading Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Figure 4.4.3: CO2 and H2S vapour and liquid phase profiles for the second set of plant data. (a) CO2 train
#1. (b) CO2 train #7. (Cond = condenser, WS = wash section, Reb = reboiler)

87
Chapter 4: Aspen Custom Modeler® Process Model Development

Vapour Phase Liquid Phase Vapour Data Liquid Data

1.2
Cond 1.2
Cond
WS WS

Fraction of Total Packed Height


Fraction of Total Packed Height
1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
30 50 70 90 110 130 60 70 80 90 100 110 120 30 50 70 90 110 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(a) (c)
1.2
Cond 1.2
Cond
WS WS

Fraction of Total Packed Height


Fraction of Total Packed Height

1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
90 100 110 120 130 60 70 80 90 100 110 120 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(b) (d)
Figure 4.4.4: Vapour and liquid phase temperature profiles for the two sets of plant data. (a) CO2 train #1
and (b) CO2 train #7 profiles for the first set of data. (c) CO2 train #1 and (d) CO2 train #7 profiles for the
second set of data. (Cond = condenser, WS = wash section, Reb = reboiler)

88
Chapter 4: Aspen Custom Modeler® Process Model Development
Table 4.4.1: CO2 train simulation results for the first set of plant data in Table 2.1.2. The bracketed values are the percentage deviations from the data.

CO2 Train #1 #2 #3 #4 #5 #6 #7
Sweet Gas
a
mol% CO2 2.49 (-0.4%) 2.49 (-0.4%) 2.51 (0.4%) 2.51 (0.4%) 2.51 (0.4%) 2.50 (0.0%) 2.53 (1.2%)
f f f f f f f
ppm H2S 1.91 (- ) 2.03 (- ) 1.70 (- ) 1.82 (- ) 2.00 (- ) 1.83 (- ) 0.58 (- )
Temperature (°C) 110.8 (0.7%) 110.7 (0.6%) 110.9 (0.8%) 110.9 (0.8%) 110.7 (0.6%) 110.9 (0.8%) 110.9 (0.8%)
Acid Gas
6 3 b
CO2 (10 sm /d) 0.319 (-0.3%) 0.269 (-0.4%) 0.478 (-0.4%) 0.548 (-0.4%) 0.548 (-0.4%) 0.667 (-0.4%) 0.906 (-0.4%)
c
Temperature (°C) 102.3 (-0.7%) 103.6 (0.6%) 79.1 (-1.1%) 76.3 (-25.9%) 94.5 (-0.5%) 94.4 (0.4%) 93.9 (-0.1%)
Lean Solution
d
CO2 loading 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%)
-16 d g g g g g g g
H2S loading (×10 ) 278.5 (- ) 30.1 (- ) 2.2 (- ) 2.2 (- ) 3.1 (- ) 0.9 (- ) 0.9 (- )
e
wt% K2CO3 27.0 (0.0%) 27.0 (0.0%) 30.0 (0.0%) 30.0 (0.0%) 30.0 (0.0%) 30.0 (0.0%) 30.0 (0.0%)
Flow (m3/h) 550.0 (0.0%) 550.0 (0.0%) 670.0 (0.0%) 812.0 (0.0%) 1000.0 (0.0%) 1000.0 (0.0%) 1068.0 (0.0%)
Temperature (°C) 110.6 (0.5%) 110.5 (0.5%) 110.7 (0.6%) 110.7 (0.6%) 110.7 (0.6%) 110.7 (0.6%) 110.7 (0.6%)
89

Rich Solution
CO2 loading d 0.815 (-0.6%) 0.748 (-0.3%) 0.864 (0.5%) 0.837 (-0.4%) 0.751 (0.1%) 0.832 (0.2%) 0.955 (-0.5%)
H2S loading (×10-5) d 4.9 (- f) 4.1 (- f) 5.5 (- f) 5.2 (- f) 4.2 (- f) 5.1 (- f) 6.8 (- f)
3
Flow (m /h) 569.5 (-0.1%) 566.7 (-0.1%) 699.0 (-0.1%) 845.3 (0.0%) 1033.6 (0.0%) 1040.7 (0.1%) 1122.4 (0.0%)
Temperature (°C) 109.7 (-0.5%) 109.9 (-0.1%) 118.4 (1.2%) 118.1 (0.9%) 117.0 (0.0%) 118.1 (0.9%) 117.1 (0.1%)
h
Makeup Water
3 f f f f f f f
Flow (m /h) 1.27 (- ) 0.96 (- ) 2.39 (- ) 2.66 (- ) 2.39 (- ) 3.22 (- ) 5.25 (- )
Absorber
f f f f f f f
Pressure drop (bar) 0.04 (- ) 0.04 (- ) 0.05 (- ) 0.06 (- ) 0.05 (- ) 0.15 (- ) 0.14 (- )
Regenerator
f f f f f f f
Pressure drop (bar) 0.01 (- ) 0.01 (- ) 0.02 (- ) 0.03 (- ) 0.01 (- ) 0.02 (- ) 0.02 (- )
Solution Reboiler
Steam flow (t/h) 30.6 (5.5%) 27.7 (13.5%) 36.5 (4.9%) 42.9 (1.4%) 47.3 (6.8%) 52.5 (6.5%) 60.3 (-7.1%)
a b 3 3 c
This refers to the CO2 content in the dry gas. sm /d refers to m /d at standard conditions (dry gas at 15°C and 1 atm). This is the vapour temperature entering the regenerator wash
d
sections in CO2 trains #1 and #2 and leaving the wash sections in trains #3 to #7. The CO2 and H2S loading refer to the moles of CO2 and H2S per equivalent mole of K2CO3 in the solution.
e
The equivalent moles of K2CO3 are defined as the total number of moles of K2CO3 in the solution if all the KHCO3, KHS and K2S in the solution are converted back into K2CO3. This refers to
f g
the equivalent weight percent of K2CO3 in the solution. No percentage deviation was calculated as the corresponding data value was not given. No percentage deviation was calculated as
h
the data value was zero. Includes the sour water recovered from the saturated sweet gas.
Chapter 4: Aspen Custom Modeler® Process Model Development
Table 4.4.2: CO2 train simulation results for the second set of plant data in Table 2.1.2. The bracketed values are the percentage deviations from the data.
CO2 Train #1 #2 #3 #4 #5 #6 #7
Sweet Gas
a
mol% CO2 2.79 (-0.4%) 2.76 (-1.4%) 2.93 (1.0%) 2.95 (1.7%) 3.14 (1.3%) 2.49 (-0.4%) 1.22 (1.7%)
f f f f f f f
ppm H2S 1.45 (- ) 1.54 (- ) 1.27 (- ) 1.36 (- ) 1.52 (- ) 1.34 (- ) 0.27 (- )
Temperature (°C) 112.2 (0.2%) 112.1 (0.1%) 112.5 (-0.4%) 112.4 (0.4%) 113.1 (1.0%) 113.5 (0.4%) 113.6 (0.5%)
Acid Gas
6 3 b
CO2 (10 sm /d) 0.303 (1.0%) 0.261 (0.4%) 0.466 (-0.9%) 0.531 (0.2%) 0.539 (-0.2%) 0.681 (0.1%) 0.827 (-0.4%)
c
Temperature (°C) 101.6 (-0.4%) 103.1 (-0.1%) 74.6 (-0.5%) 73.3 (-28.8%) 94.6 (-0.4%) 94.1 (0.1%) 94.3 (0.3%)
Lean Solution
d
CO2 loading 0.400 (0.0%) 0.400 (0.0%) 0.400 (0.0%) 0.400 (0.0%) 0.440 (0.0%) 0.350 (0.0%) 0.350 (0.0%)
-16 d g g g g g g g
H2S loading (×10 ) 1013.3 (- ) 177.0 (- ) 27.7 (- ) 47.9 (- ) 26759.0 (- ) 0.005 (- ) 0.021 (- )
e
wt% K2CO3 28.0 (0.0%) 28.0 (0.0%) 30.0 (0.0%) 29.0 (0.0%) 30.0 (0.0%) 31.0 (0.0%) 30.0 (0.0%)
Flow (m3/h) 481.0 (0.0%) 475.0 (0.0%) 638.0 (0.0%) 758.0 (0.0%) 923.0 (0.0%) 940.0 (0.0%) 946.0 (0.0%)
Temperature (°C) 111.4 (-0.5%) 111.3 (0.6%) 111.8 (-0.3%) 111.7 (-0.3%) 112.5 (0.4%) 112.9 (-0.1%) 112.6 (0.5%)
90

Rich Solution
CO2 loading d 0.857 (0.8%) 0.799 (-0.1%) 0.896 (0.7%) 0.891 (0.1%) 0.838 (-0.2%) 0.825 (0.6%) 0.941 (0.1%)
H2S loading (×10-5) d 3.7 (- f) 3.2 (- f) 4.1 (- f) 4.1 (- f) 3.3 (- f) 3.9 (- f) 4.8 (- f)
3
Flow (m /h) 499.4 (0.1%) 491.0 (0.2%) 666.1 (-0.1%) 790.0 (0.1%) 955.8 (0.1%) 981.4 (0.1%) 995.9 (0.1%)
Temperature (°C) 111.2 (-0.7%) 111.9 (-0.1%) 119.6 (1.4%) 119.2 (0.2%) 118.9 (0.8%) 120.5 (0.4%) 119.1 (-0.8%)
h
Makeup Water
3 f f f f f f f
Flow (m /h) 1.35 (- ) 1.10 (- ) 2.66 (- ) 3.05 (- ) 3.15 (- ) 3.37 (- ) 4.58 (- )
Absorber
f f f f f f f
Pressure drop (bar) 0.04 (- ) 0.04 (- ) 0.05 (- ) 0.06 (- ) 0.05 (- ) 0.15 (- ) 0.13 (- )
Regenerator
f f f f f f f
Pressure drop (bar) 0.01 (- ) 0.01 (- ) 0.01 (- ) 0.02 (- ) 0.01 (- ) 0.02 (- ) 0.02 (- )
Solution Reboiler
Steam flow (t/h) 26.2 (11.0%) 23.9 (10.6%) 32.1 (-8.3%) 37.0 (2.8%) 35.1 (-5.4%) 53.6 (8.1%) 55.2 (1.7%)
a b 3 3 c
This refers to the CO2 content in the dry gas. sm /d refers to m /d at standard conditions (dry gas at 15°C and 1 atm). This is the vapour temperature entering the regenerator wash
d
sections in CO2 trains #1 and #2 and leaving the wash sections in trains #3 to #7. The CO2 and H2S loading refer to the moles of CO2 and H2S per equivalent mole of K2CO3 in the solution.
e
The equivalent moles of K2CO3 are defined as the total number of moles of K2CO3 in the solution if all the KHCO3, KHS and K2S in the solution are converted back into K2CO3. This refers to
f g
the equivalent weight percent of K2CO3 in the solution. No percentage deviation was calculated as the corresponding data value was not given. No percentage deviation was calculated as
h
the data value was zero. Includes the sour water recovered from the saturated sweet gas.
Chapter 4: Aspen Custom Modeler® Process Model Development

4.5 Summary
In summary, the development of a new non-equilibrium rate-based model for the hot potassium
carbonate process was continued in this chapter. The relevant model equations were presented and
their implementation in Aspen Custom Modeler® and application to the Moomba CO2 trains were
discussed.

Three alternative stage models were developed for the absorber and regenerator columns: an
equilibrium model with chemical equilibrium (Model 1), a non-equilibrium rate-based model with
enhancement factors (Model 2), and a non-equilibrium rate-based model with reaction rate
expressions (Model 3). Following a series of preliminary Aspen Custom Modeler® simulations, Model
2 was identified as the optimal stage model for this work. The results produced by Model 2 were in
close agreement with those for the more complex and computationally demanding Model 3. This
indicated that the inclusion of the rigorous reaction rate expressions provided minimal improvement
compared to the enhancement factor approach.

Significant deviations were observed between some of the absorber column model predictions and
plant data from the Santos Moomba Processing Facility. An enthalpy correction term was developed
to correct the absorber temperature profile, while the rate of CO2 absorption was corrected via an
adjustment factor for the absorber effective interfacial area. This parameter was identified as having
the most significant effect on the CO2 mass transfer rate by a sensitivity analysis of the various model
parameters. In contrast, no model adjustments were required for the regenerator column model.

Using the above column models, along with simple models of the various ancillary units, process
models of the CO2 trains were developed in Aspen Custom Modeler®, based on simplified
configurations for the CO2 trains. The results from these models were found to agree well with the
plant data used in the model development process, as expected. Good agreement was also observed
with an independent set of data, thereby validating the Aspen Custom Modeler® process models.

The following chapter examines the effect of various operating parameters on the performance of the
Aspen Custom Modeler® process models, in order to develop corresponding CO2 train models in the
HYSYS® simulation environment.

91
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

CHAPTER 5

ASPEN CUSTOM MODELER®


CO2 REMOVAL TRAIN
PARAMETRIC STUDIES

The performance of the Santos Moomba CO2 trains is influenced by many key factors. While some
are constraints set by the design of the CO2 trains, others are operating parameters that can vary or
be manipulated to some degree. An investigation into the effects of the main operating parameters on
the performance of the CO2 trains was therefore considered vital to the development of reliable
process models of the CO2 trains in HYSYS®.

Following their successful development and validation in the previous chapter, the Aspen Custom
Modeler® CO2 train process models are studied in this chapter to determine the response of the CO2
trains to changes in several key operating parameters. The results of a series of simulations, in which
a specific operating parameter was adjusted while all other parameters were held constant, are
presented and discussed. Particular attention is given to the effects of the operating parameters on
the absorption and desorption of CO2 in the CO2 trains.

The parametric studies described in this chapter represent the final part of the Aspen Custom
Modeler® simulation work performed for this work.

92
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

5.1 Solution Operating Parameters


Unsurprisingly, the effectiveness of the Moomba CO2 trains is affected by the properties of the
potassium carbonate solution as it is integral to the function of the CO2 trains. This section
investigates the influence of three key operating parameters associated with the potassium carbonate
solution: flow rate, strength and acid gas loading.

5.1.1 Solution Flow Rate


The solution flow rate establishes the contact time between the vapour and liquid phases in the
column packing, as well as determining the quantity of solution available for CO2 and H2S absorption
in the absorbers and the quantity of rich solution exposed to the stripping steam in the regenerators.

The effect of the solution flow rate on the performance of the CO2 trains is illustrated in Figure 5.1.1.
The first two plots show the sweet gas CO2 and H2S content decreasing with increasing lean solution
flow to the absorbers. This enhancement of the acid gas absorption processes resulted from the
greater volumes of solution available per unit volume of acid gas. Although the increased flow rate
also reduced the length of contact between the two phases, this did not have a noticeable detrimental
effect on the gas absorption due to the rapid nature of acid gas mass transfer processes.

The differences observed in Chapter 4 between the performance of the absorbers in CO2 trains #1 to
#6 and the absorber in CO2 train #7 were highlighted in Figure 5.1.1. The lean solution flow rate had a
significantly greater effect on the sweet gas CO2 content for CO2 train #7, which also had a
considerably lower sweet gas H2S content than the other CO2 trains. These trends in the sweet gas
composition for CO2 train #7 were also observed for the other operating parameters studied in this
chapter. The two phenomena were attributed to the larger acid gas removal capacity of CO2 train #7,
which had its absorber operating far closer to the solution capacity limit, as indicated by the extremely
high CO2 loading in the rich solution. As a result, the performance of the CO2 train #7 absorber was
more sensitive to changes in the operating parameters.

The last plot of Figure 5.1.1 depicts the relationship between the rich solution flow rate and the CO2
loading of the regenerated solution. Increasing the rich solution flow rate was found to increase the
level of CO2 remaining in the regenerated solution. This reduction in the extent of solution
regeneration resulted from the lower volumes of stripping steam produced per unit volume of solution
for the given reboiler duty. As for the absorbers, increasing the rich solution flow rate also shortened
the phase contact time. However, this was expected to have a relatively insignificant effect due to the
rapid nature of acid gas desorption.

A similar trend was observed between the regenerated solution H2S loading and the rich solution flow
rate. However, this was considered inconsequential due to the H2S loading being of a very low order

93
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

of magnitude (between 10-10 and 10-22). The H2S loading therefore remained effectively zero despite
spanning several orders of magnitude over the examined range of rich solution flow rates.

Although a wide range of solution flow rates have been included in Figure 5.1.1, it should be noted
that during normal operation, the solution flow rates to the absorbers are usually maintained at a
minimum of 90% of the first set of flow rates given in Table 2.1.2 (Santos Ltd, 1998). This operating
constraint ensures optimal wetting of the absorber packing, which maximises the effective interfacial
area between the vapour and liquid phases and hence, the mass transfer between the two phases.
Keeping the solution flow rates above this minimum level also prevents corrosion of the absorber walls
by ensuring they are adequately wetted.

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

7.0 2.5 0.50

Regenerated Solution CO2 Loading


Sweet Gas CO2 Content (mol%)

6.0
Sweet Gas H2S Content (ppm)

2.0 0.45

5.0

1.5 0.40
4.0

3.0
1.0 0.35

2.0

0.5 0.30
1.0

0.0 0.0 0.25


350 850 1350 350 850 1350 350 850 1350

Lean Solution Flow Rate Lean Solution Flow Rate Rich Solution Flow Rate
(m3/h) (m3/h) (m3/h)

(a) (b) (c)


Figure 5.1.1: Effect of the solution flow rate on the performance of the CO2 trains. The influence of the
lean solution on (a) the sweet gas CO2 content and (b) the sweet gas H2S content; and the influence of the
rich solution flow rate on (c) the regenerated solution CO2 loading.

5.1.2 Solution Strength


As alluded to in the previous section, the performance of the CO2 trains is limited by the capacity of
the potassium carbonate solution to absorb acid gas. This capacity is dictated by two properties, one
of which is the solution strength. The solution strength represents the total K2CO3 content of the
solution and is usually expressed as the equivalent weight percent of K2CO3. This quantity is
calculated by converting all the KHCO3, KHS and K2S present in the solution back into K2CO3.

The sensitivity of the performance of the CO2 trains to the solution strength is presented in Figure
5.1.2. The first two plots show the concentrations of CO2 and H2S in the sweet gas decreasing with

94
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

increasing lean solution strength. This enhancement of the acid gas absorption was due to the
greater solution absorption capacity arising from the higher total K2CO3 content of the solution and
therefore, larger quantities of K2CO3 for reaction with CO2 and H2S. However, the ensuing higher
levels of acid gas in the rich solution reduced the extent of solution regeneration for the given reboiler
duty, and thereby increased the acid gas loading in the regenerated solution, as illustrated in the third
plot of Figure 5.1.2. A similar relationship between the rich solution strength and the H2S loading of
the regenerated solution was also observed, but was regarded as insignificant since the H2S loading
remained essentially zero over the examined range of solution strengths.

It should be noted that the likelihood of solution crystallisation increases with solution strength. For
this reason, the solution strength for the CO2 trains is usually maintained between 25 and 30 wt%
K2CO3 (Santos Ltd, 1998).

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

5.0 2.5 0.44

Regenerated Solution CO2 Loading


Sweet Gas CO2 Content (mol%)

4.5
Sweet Gas H2S Content (ppm)

0.42
2.0

4.0
0.40
1.5
3.5
0.38
3.0
1.0
0.36
2.5

0.5
0.34
2.0

1.5 0.0 0.32


24 26 28 30 32 34 24 26 28 30 32 34 24 26 28 30 32 34

Lean Solution Strength Lean Solution Strength Rich Solution Strength


(equivalent wt% K2CO3) (equivalent wt% K2CO3) (equivalent wt% K2CO3)
(a) (b) (c)
Figure 5.1.2: Effect of the solution strength on the performance of the CO2 trains. The influence of the
lean solution strength on (a) the sweet gas CO2 content and (b) the sweet gas H2S content; and the
influence of the rich solution strength on (c) the regenerated solution CO2 loading.

5.1.3 Solution Acid Gas Loading


The other property affecting the acid gas absorption capacity of the potassium carbonate solution is
the solution acid gas loading. This property represents the amount of K2CO3 that has been converted
to KHCO3, KHS and K2S through the absorption of acid gas. It is therefore is a measure of the actual
amount of K2CO3 available for reaction with CO2 and H2S, and can be considered to be the combined
CO2 and H2S loading in the solution. However, as the solution H2S loading for the CO2 trains is

95
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

negligible in comparison to the CO2 loading, only the effect of the solution CO2 loading on the
performance of the CO2 trains was considered here.

The effect of the solution CO2 loading on the performance of the CO2 trains is depicted in Figure 5.1.3.
The first two plots illustrate the relationship between the acid gas content of the sweet gas and the
lean solution CO2 loading. Raising the lean solution CO2 loading was observed to increase the level
of acid gas in the sweet gas. This reduction in the extent of acid gas absorption was attributed to the
lower solution absorption capacity due to the smaller quantity of K2CO3 available for reaction with CO2
and H2S.

In regenerators, the CO2 loading of the regenerated solution was found to increase with the rich
solution CO2 loading, as shown in the last plot of Figure 5.1.3. This reduction in the extent of solution
regeneration resulted from the larger quantities of CO2 present in the rich solution, which reduced the
volume of stripping steam available per unit volume of CO2 for the given reboiler duty. A similar trend
was observed for the H2S loading of the regenerated solution, but was disregarded since the H2S
loading remained effectively zero due to its very low order of magnitude.

To ensure optimal removal of acid gas in the absorbers, the CO2 trains are normally operated to
maintain a lean solution CO2 loading of less than 0.45 (Santos Ltd, 1998). This typically corresponds
to a maximum CO2 loading between 0.90 and 0.95 for the rich solution leaving the absorbers.

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

5.0 2.5 0.50


Regenerated Solution CO2 Loading

4.5
Sweet Gas CO2 Content (mol%)

Sweet Gas H2S Content (ppm)

2.0 0.45
4.0

3.5
1.5 0.40

3.0

1.0 0.35
2.5

2.0
0.5 0.30
1.5

1.0 0.0 0.25


0.33 0.38 0.43 0.48 0.33 0.38 0.43 0.48 0.6 0.7 0.8 0.9 1.0

Lean Solution Lean Solution Rich Solution


CO2 Loading CO2 Loading CO2 Loading

(a) (b) (c)


Figure 5.1.3: Effect of the solution CO2 loading on the performance of the CO2 trains. The influence of the
lean solution CO2 loading on (a) the sweet gas CO2 content and (b) the sweet gas H2S content; and the
influence of the rich solution CO2 loading on (c) the regenerated solution CO2 loading.

96
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

5.2 Raw Gas Operating Parameters


Like the potassium carbonate solution, the raw gas plays an essential role in the function of the
Moomba CO2 trains. In this section, two key operating parameters associated with the raw gas, i.e.
flow rate and acid gas content, are examined to determine their effect on the performance of the CO2
trains.

5.2.1 Raw Gas Flow Rate


The flow rate of the raw gas to the CO2 trains establishes the quantity of raw gas entering the
absorbers and, therefore, the flow rate of the sweet gas. Furthermore, like the lean solution flow rate,
the raw gas flow rate also determines the contact time between the vapour and liquid phases in the
absorber packing.

The effect of the raw gas flow rate on the performance of the CO2 trains is depicted in Figure 5.2.1. In
the first two plots, the acid gas content of the sweet gas was observed to increase with the raw gas
flow rate. This was primarily attributed to the larger volumes of raw gas to be treated per unit volume
of solution, which decreased the volume of solution available per unit volume of acid gas and thereby
constricted the acid gas absorption process. As for the solution flow rate, the reduced phase contact
times arising from the higher gas flow rates were expected to have a relatively minor effect.

The higher raw gas flow rates also pushed the CO2 train absorbers closer to the limits of the solution
capacity, which led to higher levels of acid gas in the rich solution, the effect of which has been
studied in the previous section. The relationship between the raw gas flow rate and the rich solution
CO2 loading is illustrated in the last plot in Figure 5.2.1. A similar trend was observed for the rich
solution H2S loading, but is not included here as the H2S loading was negligible in comparison with the
CO2 loading.

During the normal operation of the CO2 trains, the raw gas flow rate to the absorbers is dictated by the
demand for sales gas. To maintain the desired sweet gas composition, changes in the raw gas flow
rate are typically compensated for by corresponding changes in the solution flow rate (Santos Ltd,
1998).

97
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

4.5 2.5 1.0

4.0
Sweet Gas CO2 Content (mol%)

Sweet Gas H2S Content (ppm)


2.0 0.9

Rich Solution CO2 Loading


3.5

3.0
1.5 0.8
2.5

2.0
1.0 0.7
1.5

1.0
0.5 0.6

0.5

0.0 0.0 0.5


0.E+00 1.E+05 2.E+05 3.E+05 0.E+00 1.E+05 2.E+05 3.E+05 0.E+00 1.E+05 2.E+05 3.E+05

Raw Gas Flow Rate (sm /h)


3
Raw Gas Flow Rate (sm3/h) Raw Gas Flow Rate (sm3/h)
(a) (b) (c)
Figure 5.2.1: Effect of the raw gas flow rate on the performance of the CO2 trains. Its influence on (a) the
sweet gas CO2 content, (b) the sweet gas H2S content and (c) the rich solution CO2 loading.

5.2.2 Raw Gas Acid Gas Content


Like the raw gas flow rate, the acid gas content of the raw gas also determines the quantity of acid gas
entering the absorbers. The acid gas content of the raw gas is the combined CO2 and H2S content of
the raw gas, the latter of which is negligible in comparison to the former. Consequently, only the effect
of the raw gas CO2 content on the performance of the CO2 trains was considered here.

The sensitivity of the performance of the CO2 trains to the CO2 content of the raw gas is illustrated in
Figure 5.2.2. The first two plots show the relationship between the acid gas content of the sweet gas
and the CO2 content of the raw gas. Raising the raw gas CO2 content was observed to increase the
acid gas content in the sweet gas. This was attributed to the greater volumes of CO2 entering the
absorbers per unit volume of available solution, which therefore constricted the acid gas absorption
process.

As for the raw gas flow rate, increasing the raw gas CO2 content also pushed the CO2 train absorbers
closer to the limits of the solution capacity, which led to higher levels of acid gas in the rich solution.
The last plot in Figure 5.2.2 shows the effect of the raw gas CO2 content on the rich solution CO2
loading. The effect on the rich solution H2S loading was negligible in comparison, and has therefore
not been presented.

98
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

The acid gas content of the raw gas is dependent on the acid gas content of the raw gas supplied from
the various gas fields. As for the raw gas flow rate, changes in the raw gas acid gas content are
balanced by corresponding changes in the solution flow rate in order to maintain the desired sweet
gas composition (Santos Ltd, 1998).

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

5.0 2.5 1.0

4.5
Sweet Gas CO2 Content (mol%)

Sweet Gas H2S Content (ppm)

Rich Solution CO2 Loading


2.0
4.0 0.9

3.5
1.5

3.0 0.8

1.0
2.5

2.0 0.7
0.5
1.5

1.0 0.0 0.6


16 17 18 19 20 21 16 17 18 19 20 21 16 17 18 19 20 21
Raw Gas CO2 Content Raw Gas CO2 Content Raw Gas CO2 Content
(mol%) (mol%) (mol%)
(a) (b) (c)
Figure 5.2.2: Effect of the raw gas CO2 content on the performance of the CO2 trains. Its influence on (a)
the sweet gas CO2 content, (b) the sweet gas H2S content and (c) the rich solution CO2 loading.

99
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

5.3 Column Operating Parameters


Apart from the solution and gas properties examined in the previous sections, the performance of the
CO2 trains is also dependent on a number of operating parameters associated with the operation of
the absorber and regenerator columns. Four key column operating parameters are studied in this
section: column temperature, column pressure, steam flow rate to the regenerator solution reboilers
and makeup water flow rate to the regenerator overhead catchpots.

5.3.1 Column Temperatures


Due to the temperature dependency of the reaction rates and the equilibrium constants for the hot
potassium carbonate system, the temperatures of the absorber and regenerator columns influence the
rate at which the acid gas absorption and desorption processes take place.

In the absorbers, the column temperature profiles were observed to be primarily dependent on the
temperature of the lean solution entering the columns. In comparison, the raw gas temperature had a
relatively minor effect, as illustrated in the left and centre plots of Figure 5.3.1. As expected, the
temperatures of the rich solution and sweet gas increased with the temperatures of the lean solution
and raw gas.

Higher temperatures for the lean solution and raw gas also led to greater concentrations of CO2 in the
sweet gas, as shown in the right plots in Figure 5.3.1. This was the result of the higher temperatures
in the absorber, which reduced the solubility of CO2 in the potassium carbonate solution and retarded
the slightly exothermic CO2 absorption process. The influence of the raw gas and lean solution
temperatures on the sweet gas H2S content was negligible in comparison, and has therefore not been
presented.

In the regenerators, the temperature of the rich solution was found to establish the column
temperature profile, whereas the temperature of regenerator overhead condensers had little to no
effect. These trends are presented in the left and centre plots of Figure 5.3.2. As expected, the
temperatures of the regenerated solution and regenerator overheads increased with the rich solution
temperature.

In contrast to the absorbers, increases in the column temperatures enhanced the performance of the
regenerators due to the reduced acid gas in the solution. This enabled a greater fraction of acid gas
to flash off during the pressure reduction between the absorber and regenerator columns, along with
facilitating the steam stripping of the acid gas in the regenerators. Consequently, the CO2 loading of
the regenerated solution decreased as the regenerator temperature was increased, as shown by the
two plots on the right in Figure 5.3.2. Similar trends were observed for the regenerated solution H2S
loading, but were considered inconsequential since the H2S loading remained effectively zero due to
its very low order of magnitude.

100
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

130 125 3.0

2.9

Sweet Gas CO2 Content (mol%)


Rich Solution Temperature (°C)

125
120

Sweet GasTemperature (°C)


2.8

120 2.7
115
2.6
115
110 2.5
110
2.4
105
105 2.3

2.2
100
100
2.1

95 95 2.0
95 105 115 125 95 105 115 125 95 105 115 125
Lean Solution Lean Solution Lean Solution
Temperature (°C) Temperature (°C) Temperature (°C)
(a) (b) (c)
120 111.00 2.520

Sweet Gas CO2 Content (mol%)


110.95
Rich Solution Temperature (°C)

118 2.515
Sweet Gas Temperature (°C)

116 110.90 2.510

114 110.85 2.505

112 110.80 2.500

110 110.75 2.495

108 110.70 2.490

106 110.65 2.485


20 45 70 95 120 20 45 70 95 120 20 45 70 95 120

Raw Gas Temperature Raw Gas Temperature Raw Gas Temperature


(°C) (°C) (°C)
(d) (e) (f)
Figure 5.3.1: Effect of the absorber temperature on the performance of the CO2 trains. The influence of
the lean solution temperature on (a) the rich solution temperature, (b) the sweet gas temperature and (c)
the sweet gas CO2 content; and the influence of the raw gas temperature on (d) the rich solution
temperature, (e) the sweet gas temperature and (f) the sweet gas CO2 content.

101
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

114 108 0.47

Regenerator Overheads Temperature (°C)


Regenerated Solution Temperature (°C)

Regenerated Solution CO2 Loading


104 0.45
113
100 0.43

112 96 0.41

92 0.39
111
88 0.37
110
84 0.35

109 80 0.33

76 0.31
108
72 0.29

107 68 0.27
100 110 120 130 100 110 120 130 100 110 120 130

Rich Solution Rich Solution Rich Solution


Temperature (°C) Temperature (°C) Temperature (°C)
(a) (b) (c)
110.80 105 0.383
Regenerator Overheads Temperature (°C)
Regenerated Solution Temperature (°C)

Regenerated Solution CO2 Loading


100
0.382
110.75

95
0.381
110.70
90
0.380
85
110.65
0.379
80

110.60
0.378
75

110.55 70 0.377
45 55 65 75 45 55 65 75 45 55 65 75

Condenser Temperature Condenser Temperature Condenser Temperature


(°C) (°C) (°C)
(d) (e) (f)
Figure 5.3.2: Effect of the regenerator temperature on the performance of the CO2 trains. The influence of
the rich solution temperature on (a) the regenerated solution temperature, (b) the regenerator overheads
temperature and (c) the regenerated solution CO2 loading; and the influence of the overhead condenser
temperature on (d) the regenerated solution temperature, (e) the regenerator overheads temperature and
(f) the regenerated solution CO2 loading. (The regenerator overheads temperature is the temperature of
the vapour leaving the regenerator wash sections.)

102
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

The temperatures of the lean (or regenerated) solution and rich solution are interdependent, as
observed from Figures 5.3.1 and 5.3.2, and are usually between 100 and 120°C (Santos Ltd, 1998).
For CO2 trains #3 to #7, the raw gas to the absorbers is preheated to between 90 and 100°C by the
sweet gas. In contrast, in CO2 trains #1 and #2, the raw gas temperature is set by the temperature of
the raw gas supplied from the gas fields and varies between 25 and 45°C, depending on the ambient
conditions (Santos Ltd, 1998).

The overhead condenser temperatures are typically maintained between 60 and 70°C, which prevents
the condensation of any hydrocarbons present in the regenerator overheads (Santos Ltd, 1998). This
enables the hydrocarbons to be removed with the vented acid gas, instead of being returned to the
regenerator where they may cause foaming.

5.3.2 Column Pressures


Like the column temperatures, the absorber and regenerator pressures affect the rate at which acid
gas absorption and desorption take place. The column pressures determine the partial pressures of
CO2 and H2S over the potassium carbonate solution and, therefore, the driving force for the mass
transfer of acid gas between the vapour and liquid phases. The higher the pressure, the greater the
acid gas partial pressures and hence, the larger the driving force for acid gas mass transfer into the
liquid phase. Conversely, the lower the pressure, the smaller the acid gas partial pressures and
therefore, the larger the driving force for acid gas mass transfer into the vapour phase.

As a result, the absorbers in the Moomba CO2 trains are kept under high pressure to ensure
sufficiently high partial pressures for acid gas absorption to take place. The relationships between the
absorber pressure and the sweet gas CO2 and H2S concentrations are shown in the first two plots in
Figure 5.3.3. As expected, it was observed that the rate of acid gas absorption increased with the
absorber pressure.

In contrast, the regenerators are kept at low pressure to facilitate acid gas desorption. The last plot in
Figure 5.3.3 presents the effect of the regenerator pressure on the CO2 loading of the regenerated
solution. Unsurprisingly, the extent of solution regeneration increased with decreasing regenerator
pressure. A similar trend was observed for the regenerated solution H2S loading; however, this was
deemed insignificant since the H2S loading remained essentially zero due to its very low order of
magnitude.

The absorber pressure is set by the pressure of the raw gas, which is in turn dependent on the
pressure of the raw gas supplied from the various gas fields (Santos Ltd, 1998). Typically, the
absorber pressure is between 68 and 72 bar. The regenerator pressure is set by the head losses
between the absorber and regenerator columns, and is normally between 1.2 and 1.4 bar.

103
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

2.65 2.5 0.40

Regenerated Solution CO2 Loading


Sweet Gas CO2 Content (mol%)

2.60

Sweet Gas H2S Content (ppm)


2.0
0.39

2.55
1.5

2.50 0.38

1.0
2.45

0.37
0.5
2.40

2.35 0.0 0.36


67 69 71 73 67 69 71 73 1.1 1.2 1.3 1.4 1.5

Absorber Pressure (bar) Absorber Pressure (bar) Regenerator Pressure (bar)


(a) (b) (c)
Figure 5.3.3: Effect of pressure on the performance of the CO2 trains. The influence of the absorber
pressure on (a) the sweet gas CO2 content, (b) the sweet gas H2S content; and the influence of the
regenerator pressure on (c) the regenerated solution CO2 loading.

5.3.3 Regenerator Solution Reboiler Steam Flow Rate


As discussed previously, the performance of the CO2 trains is affected by the acid gas loading of the
lean solution entering the absorbers. This, in turn, is dependent on the ability of the regenerators to
strip the acid gas from the rich solution.

To facilitate the removal of acid gas from the rich solution, the regenerator solution reboilers boil the
solution to generate stripping steam, which travels up the regenerator columns to strip CO2 and H2S
from the downward flowing rich solution, and to further remove the CO2 and H2S remaining in the
solution. The volume of stripping steam produced and the fraction of acid gas removed in the solution
reboilers depend on the reboiler duties. Since the solution reboilers are heated by low pressure
steam, the reboiler duties are set by the reboiler steam flow rates: the higher steam flow rates, the
greater the reboiler duties.

The relationship between the steam flow rate and the CO2 loading of the regenerated solution is
shown in the first plot of Figure 5.3.4. Increasing the steam flow rate was found to enhance the extent
of solution regeneration. This was attributed to the greater volumes of stripping steam generated per
unit volume of rich solution and the larger fractions of acid gas boiled off in the reboilers. A similar
trend was observed for the regenerated solution H2S loading, but was regarded as inconsequential
since the H2S loading remained effectively zero due to its very low order of magnitude.

104
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

It was noted that the steam flow rates to the solution reboilers also affected the temperature of the
regenerated solution. From the second plot in Figure 5.3.4, it was observed that the regenerated
solution temperature increased with the steam flow, the effect of which was studied in the previous
section concerning the column temperature.

Given the observed effect of the steam flow rate, it is not surprising that during normal operation of the
CO2 trains, the steam flow rates to the regenerator solution reboilers are adjusted to ensure sufficient
solution regeneration to maintain the desired sweet gas composition (Santos Ltd, 1998).

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

0.48 113

Regenerated Solution Temperature (°C)


Regenerated Solution CO2 Loading

0.46
112

0.44

111
0.42

0.40 110

0.38
109

0.36

108
0.34

0.32 107
15 35 55 75 15 35 55 75

Steam Flow Rate (t/h) Steam Flow Rate (t/h)


(a) (b)
Figure 5.3.4: Effect of the steam flow rate to the regenerator solution reboilers on the performance of the
CO2 trains. Its influence on (a) the CO2 loading of the regenerated solution and (b) the regenerated
solution temperature.

5.3.4 Regenerator Makeup Water Flow Rate


The final column operating parameter of interest is the flow rate of makeup water to the regenerator
overhead catchpots. This makeup water compensates for the loss of water from the potassium
carbonate solution to the saturated sweet and acid gases leaving the absorbers and regenerator
columns, respectively.

The effect of the makeup water flow rate on the performance of the CO2 trains is presented in Figure
5.3.5. The first plot shows the CO2 loading of the regenerated solution increasing with the makeup
water flow rate. This was due to the increased solution flow rate, which decreased the volume of
stripping steam generated per unit volume of solution and thereby reduced the extent of solution
regeneration, as discussed in Section 5.1.1. A similar trend was observed for the H2S loading, but

105
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

was considered insignificant since the H2S loading remained effectively zero due to its very low order
of magnitude.

The second plot in Figure 5.3.5 depicts the relationship between the makeup water flow rate and the
solution strength of the regenerated solution. Increasing the makeup water flow rates was found to
dilute the potassium carbonate solution and thereby reduce the solution strength, the effect of which
was described in Section 5.1.2.

During normal operation, the makeup water flow rates are adjusted to maintain the solution flow rates
and the solution strength for the CO2 trains (Santos Ltd, 1998).

Train #1 Train #2 Train #3 Train #4 Train #5 Train #6 Train #7

0.385 30.5
Regenerated Solution CO2 Loading

0.384 30.0

Regenerated Solution Strength


(equivalent wt % K2CO3)
29.5
0.383

29.0
0.382
28.5
0.381
28.0

0.380
27.5

0.379
27.0

0.378 26.5
0 2 4 6 8 0 2 4 6 8
3 3
Makeup Water Flow (m /h) Makeup Water Flow (m /h)
(a) (b)
Figure 5.3.5: Effect of the makeup water flow rate to the regenerator overhead catchpots on the
performance of the CO2 trains. Its influence on (a) the CO2 loading of the regenerated solution and (b) the
regenerated solution strength.

106
Chapter 5: Aspen Custom Modeler® CO2 Removal Train Parametric Studies

5.4 Summary
In summary, this chapter presented and discussed the results of a series of simulations that were
performed using the CO2 train process models developed in Aspen Custom Modeler® in the previous
two chapters. These simulations investigated the influence of several key operating parameters on
the performance of the Moomba CO2 trains in order to obtain sufficient information for the
development of the HYSYS® CO2 train process models in the following chapters. The results of these
parametric studies will form the foundation of the novel modelling approach used in Chapters 6 and 7
to compensate for the limitations of HYSYS® regarding the simulation of electrolyte processes like the
hot potassium carbonate process.

The seven CO2 trains were found to behave quite similarly, with regard to the relationships observed
between the operating parameters and the performance of the absorber and regenerator columns, as
indicated by the acid gas content of the sweet gas and the regenerated solution. However, the
parametric studies did highlight differences between the performance of the absorber in CO2 train #7
and that of the other six absorbers. These dissimilarities were unsurprising given the column profiles
presented in Chapter 4, and were attributed to the larger acid gas removal capacity of CO2 train #7.
Accordingly, the absorber in CO2 train #7 operated far closer to the solution capacity limit, making it
more sensitive to the operating parameters, compared to the absorbers in the other six CO2 trains.

107
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

CHAPTER 6

MODELLING THE HOT


POTASSIUM CARBONATE
SYSTEM IN HYSYS®

As discussed in Chapter 2, HYSYS® is the primary process simulation package used by Santos, and
is therefore the desired simulation platform for modelling the Moomba CO2 trains. However, the
HYSYS® has restricted capabilities with regard to the simulation of electrolyte systems like the hot
potassium carbonate process. This chapter outlines a novel modelling approach which overcomes
these limitations of HYSYS® to enable the development of reliable process models for the CO2 trains.

The first part of this chapter focuses on the definition of hypothetical components to compensate for
the lack of electrolyte components in the standard HYSYS® component library. It also discusses the
selection of an appropriate property package and the application of column stage efficiencies in order
to accurately represent the behaviour of the hot potassium carbonate process in HYSYS®.

The second part of this chapter briefly describes the property models used in HYSYS® for the
thermodynamic and physical calculations for the hot potassium carbonate system. Particular attention
is given to the regression of model parameters for the Peng-Robinson equation of state, which was
used for the thermodynamic calculations.

This chapter begins the development of CO2 train process models in HYSYS®, which is continued in
the following three chapters.

108
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

6.1 The Modelling Approach


In this work, a novel approach was undertaken to model the hot potassium carbonate process in the
HYSYS® simulation environment. This approach was designed to overcome the limitations of
HYSYS® concerning such electrolyte systems, such that reliable process models for the CO2 trains
could be developed in HYSYS®. The following sections outline the three main aspects of this
modelling approach: the definition of hypothetical components, the selection and modification of an
appropriate property package and the application of column stage efficiencies.

6.1.1 Definition of Hypothetical Components


The first step in the development of any HYSYS® model is the definition of the components in the
system of interest. In this work, the system of interest is the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O
system, which forms the basis of the hot potassium carbonate process.

Unfortunately, none of the electrolyte species (i.e. K2CO3, KHCO3, KHS, K2S, K+, H3O+, HCO3-, CO32-,
HS-, S2- and OH-) associated with the above system are available as standard HYSYS® library
components. However, such species can be introduced into a HYSYS® simulation as hypothetical
HYSYS® components, which are denoted by asterisks: K2CO3*, KHCO3*, KHS*, K2S*, K+*, H3O+*,
HCO3-*, CO32-*, HS-*, S2-* and OH-*. This approach was followed for this work.

In HYSYS®, both hypothetical and library components are divided into a number of classes, such as
Hydrocarbon, Organic, Inorganic and Miscellaneous, and two types of components are available
within each class: liquid or solid. Given that components of the latter type do not participate in the
vapour-liquid equilibrium calculations and can only be used in steady-state simulations, the eleven
electrolyte species of interest were defined as liquid hypothetical components. Since the H2O library
component is a Miscellaneous-class component, the hypothetical electrolyte components were
likewise created as components of the Miscellaneous class.

A number of property estimation methods are included in HYSYS® to estimate the critical, physical
and thermodynamic properties of a hypothetical component. The default estimation methods for the
Miscellaneous class of components are listed in Table 6.1.1. In this work, the following properties
were specified for the hypothetical electrolyte components from literature data: molecular weight,
normal boiling point, ideal liquid density, vapour pressure, ideal gas enthalpy, ideal gas Gibbs energy
and heat of formation. The remaining properties were estimated via the HYSYS® default methods,
which are discussed in further detail in Appendix E.

109
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

Table 6.1.1: Property estimation methods for the Miscellaneous class of hypothetical components.

Property Default Estimation Methods References


Critical temperature Lee-Kesler Kesler and Lee (1976)
Bergman Bergman (1976)
Cavett Cavett (1964)
Critical pressure Lee-Kesler Kesler and Lee (1976)
Bergman Bergman (1976)
Cavett Cavett (1964)
Critical volume Pitzer Pitzer et al. (1955)
Acentricity Pitzer Pitzer et al. (1955)
Molecular weight Bergman Bergman et al. (1975)
Lee-Kesler Kesler and Lee (1976)
Normal boiling point Proprietary method Hyprotech Ltd (2001a)
Vapour pressure Riedel Riedel (1954)
Ideal liquid density Yen-Woods Yen and Woods (1966)
Ideal gas enthalpy Cavett Cavett (1964)
Heat of formation Joback Joback (1984)
Ratio with respect to octane Hyprotech Ltd (2001a)
Ideal gas Gibbs energy Proprietary method Hyprotech Ltd (2001a)
Dipole moment Set equal to zero Hyprotech Ltd (2001a)
Radius of gyration Proprietary method Hyprotech Ltd (2001a)

Preliminary simulations with the full component set of H2O, N2, the hydrocarbons CH4 to C7H16 and the
eleven hypothetical electrolyte components were performed to investigate the behaviour of the
hypothetical components. It should be noted that a property package had to be selected in order to
run these simulations, and this is discussed in the next section. It was found that the default values for
the property package’s adjustable parameters were inadequate for describing the interaction between
the hypothetical components and H2O. The hypothetical components tended to form a separate liquid
phase (with the hydrocarbon components) from the aqueous phase containing H2O, indicating that
values for these parameters had to be regressed from literature data to ensure more correct liquid
phase behaviour.

It was also discovered that the interaction parameters for the CO2-H2O and H2S-H2O component pairs
are functions of temperature and therefore could not be altered, unlike the other component pair
parameters with constant values. Since these two component pairs are critical to the behaviour of the
hot potassium carbonate system, it was decided to clone the H2O library component to create the
hypothetical H2O* component to enable the regression of these two significant component pair
parameters. The H2O* component was identical to H2O, except that its interaction parameters could
be specified and it no longer formed an aqueous phase. This resolved the issue with the formation of
two separate liquid phases, although regression of the interaction parameters was still required to
accurately represent the vapour-liquid equilibrium behaviour. The regression of the property package
interaction parameters is discussed later in this chapter.

110
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

The preliminary simulations also compared the true and apparent component approaches for the liquid
phase composition. The former method considers the liquid phase speciation reactions and resulted
in a more complex system due to the involvement of all eleven hypothetical electrolyte components. A
further complication was that the speciation reactions given in Table 2.1.6 and the kinetic reactions
(R2.1.6) and (R2.1.15) had to be incorporated into the simulations to describe the vapour-liquid phase
behaviour. The equilibrium reactions were easily specified in HYSYS®; however, the ionic strength
dependency of the reaction constants for the two kinetic reactions had to be excluded since HYSYS®
only considers these constants as functions of temperature, as defined by an extended form of the
Arrhenius equation.

Besides this issue with the kinetic reaction constants, another problem with using the speciation and
kinetic reactions in HYSYS® was that HYSYS® limits reactions to column, separator and reactor
operations. Consequently, modelling difficulties were encountered for other types of equipment,
namely valves, the power recovery turbines and the solution reboilers, in which CO2 and H2S are
flashed off due to a pressure drop or the application of heat. It was not possible to properly simulate
the desorption of CO2 and H2S in these operations without the use of chemical reactions as attempts
to do so resulted in either mass balance inconsistencies (observed for the valves and power recovery
turbines) or the absence of acid gas desorption (observed for the solution reboilers). Attempts to
model these pieces of equipment using separator operations with reactions were just as unsuccessful
as the separator operations could not converge to a solution.

Convergence difficulties were also experienced for the column operations, especially for the
regenerators. Both the absorber and regenerator columns converged rapidly (within a couple of
seconds) in the absence of reactions; however, problems were encountered with the inclusion of the
speciation and kinetic reactions. The absorbers were found to converge fairly quickly if the simulation
initial values were taken from a solved non-reactive absorption run; otherwise, convergence was not
achieved. The regenerators, on the other hand, either had not reached a solution after a lengthy
period of time (in excess of 10 minutes) or failed to converge, even when the solution to a non-reactive
regeneration run was used as the initial values. This critical lack of convergence for the regenerator
columns, along with the other above-mentioned modelling difficulties, rendered the true component
approach infeasible for this work.

Fortunately, the simulation difficulties associated with the true component approach were avoided by
the apparent component approach as this method ignores the solution chemistry. Consequently, only
the K2CO3* hypothetical electrolyte component was required and the liquid phase reactions were
excluded from the simulations. The effect of the solution chemistry on the performance of the
absorber and regenerator columns was instead taken into account through the use of column stage
efficiencies, which are discussed further on in this chapter. The physical and thermodynamic
properties of the K2CO3* hypothetical component are given in Appendix E.

111
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

6.1.2 Selection of Property Package


The next step in the development of a HYSYS® model is the selection of an appropriate property
package for the thermodynamic and physical property calculations associated with the HYSYS®
model. Since suitable electrolyte property packages are not included in HSYSY®, one of the standard
HYSYS® non-electrolyte property packages had to be used instead for this work.

In HYSYS®, the recommended default property package for oil, gas and petrochemical systems is the
Peng-Robinson (PR) property package, which is based on a proprietary enhanced version of the PR
equation of state (Peng and Robinson, 1976). The enhanced PR equation of state is also the basis for
the Sour PR property package, which is a modification of the PR property package to suit sour water
systems. The Sour PR property package combines the enhanced PR equation of state with the
Wilson API-Sour method (Wilson, 1980) to account for the ionisation of CO2 and H2S in the aqueous
phase. In the absence of an aqueous phase, the Sour PR property package reduces to the PR
property package.

As shown in Section A.3.2 in Appendix A, the PR equation of state is relatively simple and requires
only a single adjustable binary interaction parameter for every component pair in the system of
interest. It is also applicable to both vapour and liquid phases. In contrast, the activity coefficient
models available in HYSYS®, such as the NRTL model, are more complex. They involve a larger
number of adjustable parameters and require the use of a separate vapour phase thermodynamic
model. However, while activity coefficient models are more suitable for non-ideal systems, they
should be limited to pressures less than 5 atm whereas the enhanced PR equation of state is
applicable to pressures up to 1000 bar (Hyprotech Ltd, 2001a). Since the hot potassium carbonate
process involves pressures around 70 bar, the enhanced PR equation of state was considered more
appropriate for this work than the activity coefficient models available in HYSYS®.

An alternative to the enhanced PR equation of state, the Peng-Robinson Stryjek-Vera (PRSV)


equation of state (Stryjek and Vera, 1986ab) was also considered. This model is an extension of the
original PR equation of state and it forms the basis for the PRSV property package in HYSYS®.
Stryjek and Vera (1986ab) introduced two new adjustable parameters (a pure component vapour
pressure parameter and an additional component pair binary interaction parameter) to extend the
applicability of the original PR equation of state to moderately non-ideal (non-electrolyte) systems.

However, despite these modifications, the PRSV equation of state still has the same inherent
limitations as the enhanced PR equation of state with regard to the representation of electrolyte
systems. Like all traditional thermodynamic models, neither equation of state can accurately describe
the non-ideal behaviour associated with the electrostatic interactions observed in electrolyte systems.
Since the modelling approach for this work involved the use of column stage efficiencies to
compensate for the short-comings of the thermodynamic model, it was decided that the

112
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

thermodynamic model should be as simple as possible. Consequently, the enhanced PR equation of


state was selected as the thermodynamic model for the HYSYS® simulations performed for this work.

A modified form of the standard PR property package was selected as the primary property package
for the HYSYS® simulations. The modifications concerned the use of tabular physical property
models instead of the default HYSYS® models to accommodate the hypothetical H2O* and K2CO3*
components and to better represent the properties of the hot potassium carbonate system. However,
these changes rendered the modified PR property package unsuitable for sour water systems, in
which potassium carbonate was absent. As a result, a secondary property package, the standard
Sour PR property package, was required for the sour water sections of the HYSYS® models, and a
second component set was defined for these sections to exclude K2CO3* and to replace H2O* with the
HYSYS® library component H2O. The standard Sour PR property package was selected instead of
the standard PR property package due to its greater accuracy in handling sour water systems.

The thermodynamic and physical property models associated with the two selected property packages
are further discussed in Section 6.2 and in Appendix F.

6.1.3 Column Stage Efficiencies


As discussed in the preceding sections, a modified PR property package, based on an apparent
component liquid phase composition, was used to characterise the physical and thermodynamic
behaviour of the HYSYS® models developed in this work. This approach, however, did not facilitate
the direct inclusion of the solution chemistry associated with the hot potassium carbonate process.
Consequently, the effects of the speciation reactions and the kinetic reactions were instead indirectly
accounted for through the use of column stage efficiencies.

In HYSYS®, column operations consist of equilibrium stages, which require stage efficiencies to relate
the performance of these ideal stages to that of real stages in which thermodynamic equilibrium is not
achieved. For steady-state simulations, these stage efficiencies are represented by a modified
version of the Murphree stage efficiency, as defined in the first figure of Figure 6.1.1. This can be
specified for each stage in one of two ways: as a single overall value (i.e. the same value is applied to
all components on the stage) or as individual values for each component.

In contrast, in dynamic simulations, the stage efficiency is treated as a vapour bypass which diverts
part of the vapour stream entering a stage so that it mixes with the vapour leaving the stage instead of
mixing with the liquid entering the stage. This is illustrated in the second figure of Figure 6.1.1.
Consequently, only a single overall efficiency value can be specified for each stage in dynamic mode.
While the dynamic and steady-state stage efficiencies are not directly related mathematically, they
share the same general effect on the column performance: the lower the stage efficiency, the poorer
the column performance.

113
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

NOTE:
This figure is included on page 114 of the print copy of
the thesis held in the University of Adelaide Library.

Figure 6.1.1: The different definitions of the column stage efficiency K in HYSYS®. (a) Steady-state
column. (b) Dynamic column (adapted from Hyprotech Ltd (2001b)).

In this work, the column stage efficiencies were not only used to represent the non-equilibrium
behaviour of the absorber and regenerator columns in the Moomba CO2 trains, but were also used to
describe the effect of the liquid phase reactions on the column performance. To achieve this, stage
efficiency correlations were developed for each HYSYS® CO2 train model. These were based on the
parametric studies for the Aspen Custom Modeler® CO2 train models from Chapter 5, and defined the
overall column efficiencies as functions of temperature, pressure, composition and flow. A similar
approach is taken by the commercial add-in Amines property package for alkanolamine gas
sweetening processes, in which CO2 and H2S component stage efficiencies are calculated as
functions of the operating conditions, tray specifications and kinetic and mass transfer parameters
(Hyprotech Ltd, 2001a).

Chapter 7 discusses the development and implementation of the stage efficiency correlations in
greater detail.

114
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

6.2 Thermodynamic and Physical Property Models


6.2.1 Summary of Property Models
Two different property packages were used in the HYSYS® models to facilitate the more reliable
simulation of the CO2 trains: the PR and the Sour PR property packages. The standard HYSYS®
property models presented in Table 6.2.1 were used to perform the general thermodynamic and
physical property calculations for these two property packages. Table 6.2.1 also includes the
references from which the relevant model equations and/or model parameters were taken.

Where the standard models were unsuitable, tabular property models were used instead for the
property calculations. The model parameters for these tabular models, and for the enhanced PR
equation of state, were derived from the literature data listed in Table 3.1.3.

Table 6.2.1: Thermodynamic and physical property models.

Property PR Package Sour PR Package References


Fugacity Enhanced PR Enhanced PR Peng and Robinson (1976)
coefficients
Enthalpy Enhanced PR Enhanced PR Peng and Robinson (1976)
Heat capacity From Cv value From Cv value Hyprotech Ltd (2001a)
Molecular Weighted average Weighted average Hyprotech Ltd (2001a)
weight
Density Enhanced PR (V) Enhanced PR (V) Peng and Robinson (1976),
Tabular model (L) COSTALD (L) Hyprotech Ltd (2001a),
Hankinson and Thomson
(1979)
Dynamic Modified Ely-Hanley (V) Modified Ely-Hanley Ely and Hanley (1981),
viscosity Tabular model (L) Twu (L) Hyprotech Ltd (2001a), Twu
Modified Letsou-Stiel (L) (1985), Letsou and Stiel (1973)
Surface Tabular model (L) Modified Brock-Bird (L) Brock and Bird (1955),
tension Proprietary polynomial (L) Hyprotech Ltd (2001a)
Thermal Modified Ely-Hanley (V) Modified Ely-Hanley Ely and Hanley (1983),
conductivity Misic-Thodos (V) Misic-Thodos (V) Hyprotech Ltd (2001a), Misic
Chung (V) Chung (V) and Thodos (1961, 1963),
Tabular model (L) Proprietary polynomial (L) Chung and co-workers (1988),
Modified Missenard- Reid and co-workers (1977),
Riedel (L) Baroncini and co-workers
Latini (L) (1980)
Sato-Riedel (L)
Note: (V) refers to the vapour phase and (L) refers to the liquid phase. If no phase is indicated, the model applies to
both vapour and liquid phases.

The reliable simulation of the hot potassium carbonate process in HYSYS® requires an accurate
representation of the thermodynamic behaviour of the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system.
Consequently, the key properties of interest are the fugacity coefficients, which were calculated using
the enhanced PR equation of state. The next section discusses the regression of the adjustable
model parameters for this thermodynamic model. For the purpose of brevity, detailed descriptions of
the property models have not been included in the main body of this thesis. Instead, the equations for

115
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

the original PR equation of state are presented in Section A.3.2 in Appendix A, while the other
thermodynamic and physical property models are discussed in greater detail in Appendix F.

6.2.2 The Enhanced Peng-Robinson Equation of State


As discussed previously, the enhanced PR equation of state, a HYSYS® proprietary version of the
original PR equation of state, was used to determine the fugacity coefficients for this work. This
equation of state requires a binary interaction parameter kjk (where kjk = kkj and kjj = 0) for each
component pair j-k in the system of interest. To ensure the best representation of the system
thermodynamic behaviour, these interaction parameters should be determined from experimental
data.

Binary interaction parameter values for the full set of non-hypothetical component pairs of interest are
included in the HYSYS® property package library, and these default values were used in the Sour PR
property package. For the PR property package, the presence of the hypothetical water H2O* and
potassium carbonate K2CO3* components required the regression of vapour-liquid equilibrium data to
obtain the parameter values. It should be noted that the default value is 0 for component pairs that
are not included in the HYSYS® property package library.

For the hot potassium carbonate process, the system of interest is the CO2-H2S-K2CO3-KHCO3-KHS-
K2S-H2O system, which reduces to the CO2-H2S-K2CO3-H2O system based on the apparent
component approach. The five key component pairs in the PR property package were therefore H2O*-
K2CO3*, CO2-H2O*, CO2-K2CO3*, H2S-H2O* and H2S-K2CO3*. The binary interaction parameter values
for these five pairs were determined from two sets of regressions, which are described in the following
sections. The parameters for the remaining component pairs involving H2O* (i.e. N2-H2O*, CH4-H2O*,
etc.) were set equal to the default HYSYS® values for the corresponding component pairs with H2O.

Default HYSYS® parameter values were also used for the non-hypothetical component pairs. The
CO2-H2S component pair parameter was kept at the HYSYS® default value and was not re-regressed
(unlike the other component pairs in the CO2-H2S-K2CO3-H2O system) because preliminary
simulations showed that the parameter value sets that included the re-regressed CO2-H2S component
pair parameter value resulted in negligible H2S desorption in the regenerator.

It was expected that the enhanced PR equation of state would have inherent limitations in describing
accurately the thermodynamic behaviour of the CO2-H2S-K2CO3-H2O system. Consequently, the
model could not be assumed to be capable of representing the vapour-liquid equilibrium data without
any systematic deviation. Since this is one of the key assumptions in the maximum likelihood method,
the simpler and statistically less rigorous implicit least squares method was instead applied to the
parameter regressions.

116
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

The regressions were performed in a Microsoft® Excel spreadsheet, and the Gauss-Newton method
(with the bisection rule) was used to minimise the following objective function Q:
ND NC
Q ¦¦ ln f
i 1 j 1
L,i, j  ln f G,i, j 2 (6.2.1)

where ND is the number of data points, NC is the number of components, and f is the fugacity. The
subscripts G and L refer to the vapour and liquid phases, respectively, and the subscripts i and j
denote the data point i and the component j, respectively. The above objective function was
minimised via the manipulation of the model parameters, subject to the vapour-liquid equilibrium
constraints.

6.2.2.1 Parameters for the CO2-K2CO3-H2O system


The first set of regressions was performed to determine the binary interaction parameters for the
H2O*-K2CO3*, CO2-H2O* and CO2-K2CO3* component pairs from the vapour-liquid equilibrium data
presented by Tosh and co-workers (1959). Details of this data set were given in Section 3.2.1. The
liquid phase composition was determined as for the Aspen Properties® DRS regression runs, i.e. from
equations (3.2.6) to (3.2.9), and only CO2 and H2O* were assumed to be present in the vapour phase.

A series of regressions were performed and the resulting optimal set of parameter values is presented
in Table 6.2.2, while Figure 6.2.1 compares the predicted CO2 and H2O* partial pressures against the
literature data. The procedure used to obtain the optimal parameter set is described in Appendix G.
The absolute average deviations between the predicted and literature values were 135.9% and 15.8%
for CO2 and H2O*, respectively. In comparison, the average absolute deviations were substantially
larger (24500.7% and 126.7%, respectively) when the default HYSYS® parameter values were used.

The large and systematic deviation between the experimental and literature values for the CO2 partial
pressures illustrated the inherent limitations for the enhanced PR equation of state in representing the
CO2-K2CO3-H2O system, and confirmed the unsuitability of the maximum likelihood method for the
regression of the parameter values. The inadequacy of the enhanced PR equation of state required
the use of column stage efficiencies to more accurately represent the reactions occurring in the
absorber and regenerator units in the HYSYS® CO2 train models. This will be discussed further in the
next chapter.

Table 6.2.2: Enhanced PR parameter values for the CO2-K2CO3-H2O system.

Component Pair Parameter Value Standard Deviation


H2O*-K2CO3* -1.1152 0.0748
CO2-H2O* -0.5891 0.1019
CO2-K2CO3* -0.3836 0.2783

117
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

1.0E+01 1.0E+01

Predicted H2O* Partial Pressure (bar)


Predicted CO2 Partial Pressure (bar)
20 wt% 20 wt%
30 wt% 30 wt%
1.0E+00
40 wt% 40 wt%

1.0E-01 1.0E+00

1.0E-02

1.0E-03 1.0E-01
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E-01 1.0E+00 1.0E+01
Experimental CO2 Partial Pressure (bar) Experimental H2O* Partial Pressure (bar)

(a) (b)
Figure 6.2.1: Comparison between the enhanced PR predictions and the experimental data. Partial
pressure plots for (a) CO2 and (b) H2O*. The points represent the partial pressures over 20, 30 and 40 wt%
equivalent K2CO3 solutions. The dashed lines (---) represent the ± 20% lines.

6.2.2.2 Parameters for the CO2-H2S-K2CO3-H2O system


The second set of regressions was performed to determine the binary interaction parameters for the
H2S-H2O* and H2S-K2CO3* component pairs using the parameter values in Table 6.2.2 and the
vapour-liquid equilibrium data presented by Tosh and co-workers (1960). Details of this data set were
given in Section 3.2.2. The liquid phase composition was determined as for the Aspen Properties®
DRS regression runs, i.e. from equations (3.2.6), (3.2.7) and (3.2.10) to (3.2.12). Only H2S, CO2 and
H2O* were assumed to be present in the vapour phase.

A series of regressions were performed and the resulting optimal set of parameter values is presented
in Table 4.2.3, while Figure 4.2.2 compares the predicted CO2, H2S and H2O* partial pressures against
the literature data. The procedure used to obtain the optimal parameter set is described in detail in
Appendix G. The absolute average deviations between the predicted and literature values were
63.7%, 19.9% and 274.5% for CO2, H2S and H2O*, respectively. In comparison, the average absolute
deviations were significantly larger (11576.1%, 144.7% and 24137.7%, respectively) when the default
HYSYS® parameter values were used.

Most of the points on the plots in Figure 6.2.2 were found to be relatively evenly scattered about the
y = x diagonal. However, a distinct outlying cluster of 30 wt% points was observed in the lower right-
hand corner of the first plot in Figure 6.2.2. These points corresponded to the CO2 partial pressures
over a 30 wt% equivalent K2CO3 solution converted only by H2S, and their substantial underestimation
demonstrated the inherent limitations of the enhanced PR equation of state in representing the CO2-
H2S-K2CO3-H2O system. The enhanced PR equation of state could not adequately describe the
system equilibria in Table 2.1.6, and consequently was unable to accommodate the CO2-H2S-K2CO3-
H2O system in which only H2S is absorbed into the K2CO3 solution. However, this particular

118
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

inadequacy was not of concern for this work as the hot potassium carbonate process of interest
focuses primarily on the removal of CO2.

Table 6.2.3: Enhanced PR parameter values for the CO2-H2S-K2CO3-H2O system.

Component Pair Parameter Value Standard Deviation


H2S-H2O* -1.2366 0.5274
H2S-K2CO3* 0.0951 2.3660

1.0E+01 1.0E+00
Predicted CO2 Partial Pressure (bar)

Predicted H2S Partial Pressure (bar)


20 wt% 20 wt%
30 wt% 30 wt%
40 wt% 1.0E-01 40 wt%

1.0E-01

1.0E-02

1.0E-03
1.0E-03

1.0E-05 1.0E-04
1.0E-05 1.0E-03 1.0E-01 1.0E+01 1.0E-04 1.0E-03 1.0E-02 1.0E-01 1.0E+00
Experimental CO2 Partial Pressure (bar) Experimental H2S Partial Pressure (bar)

(a) (b)
1.0E+01
Predicted H2O* Partial Pressure (bar)

20 wt%
30 wt%
40 wt%

1.0E+00

1.0E-01
1.0E-01 1.0E+00 1.0E+01

Experimental H2O* Partial Pressure (bar)

(c)
Figure 6.2.2: Comparison between the enhanced PR predictions and the experimental data. Partial
pressure plots for (a) CO2, (b) H2S and (c) H2O*. The points represent the partial pressures over 20, 30
and 40 wt% equivalent K2CO3 solutions. The dashed lines (---) represent the ± 20% lines.

119
Chapter 6: Modelling the Hot Potassium Carbonate System in HYSYS®

6.3 Summary
In summary, this chapter outlined the novel modelling approach used in this work to overcome the
limitations associated with modelling the hot potassium carbonate process in HYSYS®. The three
main features of this approach were: the introduction of hypothetical electrolyte components, the
selection of a suitable HYSYS® property package, and the application of column stage efficiencies.

The hypothetical electrolyte components were created to compensate for the absence of electrolyte
components in the HYSYS® component library, which would have otherwise prevented the hot
potassium carbonate system from being modelled in this simulation environment. The behaviour of
the hypothetical components was investigated in a series of preliminary simulations, which determined
that the liquid phase composition was most feasibly represented by the apparent component
approach. However, by adopting the apparent component approach, the solution chemistry had to be
represented implicitly through the property package model parameters and the use of column stage
efficiencies.

Due to the lack of electrolyte property packages in HYSYS®, the PR property package was selected
for its simplicity and was modified to provide a more accurate representation of the hot potassium
carbonate system. Particular attention was given to the regression of the model parameters for the
enhanced PR equation of state, while the modifications to the other property models in the PR
property package were discussed in Appendix F. Significant improvements in the model predictions
were achieved with the modifications to the PR property package; however, large systematic
deviations were still observed for the vapour-liquid equilibria. These highlighted the inherent
limitations of the enhanced PR equation of state in describing the chemistry associated with the hot
potassium carbonate system.

To compensate for the above inadequacies, column stage efficiencies were applied to the absorber
and regenerator columns to better represent the effects of the solution chemistry. These efficiencies
are normally used to relate the ideal HYSYS® column stages to real non-equilibrium stages, and
consequently also facilitated a more accurate depiction of the non-equilibrium behaviour of these
columns. To enable a reasonable representation of the non-equilibrium column behaviour and the
effects of the solution chemistry, the stage efficiencies were correlated against the column operating
parameters based on the results of the Aspen Custom Modeler® parametric studies in Chapter 5. The
development of these stage efficiency correlations and their implementation in the absorber and
regenerator column models are further discussed in the following chapter.

120
Chapter 7: The Absorber and Regenerator Column Models

CHAPTER 7

THE ABSORBER AND


REGENERATOR COLUMN
MODELS

This chapter continues the discussion on the simulation work performed in HYSYS® to model the hot
potassium carbonate process in the form of the Moomba CO2 trains. In this chapter, the focus shifts to
the development of the absorber and regenerator column models, in particular, the correlation and
implementation of column stage efficiencies.

It was determined in the previous chapter that, despite the modifications to the PR property package,
the inherent limitations of the enhanced PR equation of state prevented the accurate representation of
the solution chemistry for the hot potassium carbonate process. To compensate for this inadequacy,
the application of column stage efficiencies was proposed to improve the performance of the absorber
and regenerator column models.

This chapter investigates the feasibility of the above suggested modelling approach. Detailed
absorber and regenerator column models for the CO2 trains are developed using the standard
HYSYS® operations. A series of preliminary simulations are performed to examine the performance
of these column models and their sensitivity to variations in the column stage efficiencies. Finally, the
Aspen Custom Modeler® parametric study results from Chapter 5 are used to correlate the column
stage efficiencies with respect to the column operating parameters, and the implementation of these
correlations in HYSYS® is discussed.

121
Chapter 7: The Absorber and Regenerator Column Models

7.1 Absorber Column Models


In HYSYS®, column operations are treated as a series of equilibrium stages which comprise the
column tray sections as well as any attached reboiler and condenser operations. In this work, the
HYSYS® absorber operation, consisting solely of a tray section, was used to model the absorber
columns in the Moomba CO2 trains.

This approach necessitated the discretisation of the packed heights of the CO2 train absorbers into a
number of stages, just as required for the Aspen Custom Modeler® absorber models in Chapter 4.
Consequently, the number of stages for each absorber operation was set equal to that for the
corresponding Aspen Custom Modeler® absorber model. This ensured that the HYSYS® and Aspen
Custom Modeler® absorber models were directly comparable. Other sizing information was obtained
from the operator and vendor manuals and the technical data sheets and drawings stored in the
Santos TIMS (The Information Management System) controlled documentation database.

An idiosyncrasy of the HYSYS® absorber operations is that they end with a tray and not a sump,
unlike real absorber columns. In steady-state simulations, the absence of a column sump had no
effect on the simulation results. However, in dynamic simulations, column sumps are essential to
ensure proper pressure driven flow, and they enable the modelling of level control at the bottom of the
column. Consequently, for this work, appropriately sized separator operations were used to represent
the absorber sumps in HYSYS®, resulting in the general absorber model layout in Figure 7.1.1. For
consistency with the real absorber columns, the vapour streams leaving the top of the modelled
sumps were recycled back to the bottom of the absorber tray sections so that the absorber models
only had two feed streams (RawGas and LeanSoln) and two product streams (SweetGas and
RichSoln).

Figure 7.1.1: Process flow diagram of an absorber column model in HYSYS®. The // denote a change in
property package.

122
Chapter 7: The Absorber and Regenerator Column Models

It should be noted that in steady-state simulations, HYSYS® can only solve material recycle loops if
they contain a recycle operation. For this reason, a recycle operation RCY-SUMP was included in the
vapour return circuit at the bottom of the absorber models, as shown in Figure 7.1.1. The HYSYS®
recycle operation acts as a theoretical tear in a material stream, in which the inlet stream properties
are calculated based on assumed values for the outlet stream. The outlet stream conditions are
modified iteratively until the two streams agree within a specified tolerance. In this work, the default
variable tolerances were used in the recycle operations, except for the flow and composition
tolerances. These were tightened by a factor of 10 to accommodate the extremely low concentration
of H2S in the system.

The physical and thermodynamic property calculations for the absorber models were performed using
both the PR and the Sour PR property packages described in Chapter 6. The PR property package
was the primary property package and was applied to the entire absorber model, except for the vapour
feed and product streams. The property calculations for these two streams were instead performed by
the Sour PR property package due to the absence of the hypothetical K2CO3* component.

To accommodate this change in property package between the two vapour streams and the rest of the
absorber model, a transfer basis had to be specified. The T-P (Temperature-Pressure) flash transfer
basis was selected, since it can be used between different property packages, unlike the default P-H
(Pressure-Enthalpy) flash transfer basis (Aspen Technology Inc., 2006). For a material stream
undergoing a change in property package, the T-P flash transfer basis transfers the composition, flow,
temperature and pressure between the two property packages and recalculates the enthalpy and
vapour fraction. In contrast, the P-H flash transfer basis transfers the composition, flow, pressure and
enthalpy and recalculates the temperature and vapour fraction.

123
Chapter 7: The Absorber and Regenerator Column Models

7.2 Regenerator Column Models


In contrast to the relatively simple absorber models described above, the regenerator column models
were considerably more complex and involved different configurations for the seven CO2 trains. This
increased model complexity can be observed in Figure 7.2.1, which depicts the general layout of the
two most dissimilar regenerator models: the regenerator models for CO2 trains #1 and #7. For clarity,
parallel items such as the solution reboilers have been represented by single units in Figure 7.2.1.

Besides the presence of the cooling water circuit in the model for CO2 train #1, the other significant
difference between the two models in Figure 7.2.1 concerned the solution reboilers. Kettle reboilers
are used in CO2 train #1 (and in CO2 trains #2 to #4), whereas the solution reboilers in CO2 train #7
(and in CO2 trains #5 and #6) are semi-thermosyphon-type reboilers. These two types of reboilers
operate in quite different ways. In kettle reboilers, the generated steam is fed back into the bottom of
the regenerator while the heated solution leaves as the regenerator bottoms product. In contrast, the
steam and solution from the semi-thermosyphon reboilers flow back together into the bottom of the
regenerator. The two different reboiler operations are reflected in the dissimilar reboiler and sump
arrangements shown in Figure 7.2.1.

Like the absorber column models, the regenerator column models for the CO2 trains were developed
around the HYSYS® absorber operation. The regenerator main packed sections were modelled by
absorber operations with the same number of stages as for the corresponding Aspen Custom
Modeler® regenerator models. Likewise, the regenerator wash sections were represented by
absorber operations. For the valve tray-type wash sections, the number of stages was taken as the
number of trays. For the packed wash sections, the HETP of the corresponding main packed section
was used to determine the number of stages.

As for the absorber models, separator operations were used to model the regenerator sumps. These
operations were also used to represent the overhead catchpots and the steam condensate receivers.
The solution reboilers were modelled by heat exchanger operations, while the overhead condensers
and water coolers were modelled as cooler operations. Pump operations were used to model the
solution, cooling water, steam condensate and overhead condensate pumps, while valve operations
were used to represent the pressure drops across the solution filters. The actual function of the filters
was not considered in the HYSYS® simulations due to the exclusion of solid components from the
simulation system. All these various components of the regenerator models were sized according to
information obtained from the Santos TIMS database.

Up to five material recycle loops were associated with each regenerator model: the reboiler circuit, the
solution filter circuit, the wash section circuit, the overhead condensate circuit and the cooling water
circuit (only for CO2 trains #1 to #4). The location of the recycle operations was carefully considered
to maximise the model stability and to minimise the number required, so as to reduce the simulation
computation time. Special consideration was given to ensure the assigned recycle operations could

124
Chapter 7: The Absorber and Regenerator Column Models

also accommodate the solution recycle and sour water circuits between the absorber and regenerator
models, thereby minimising the number of recycle operations required for the completed CO2 train
process models. Due to the configuration of the regenerator recycle loops, a minimum of three
recycle operations were necessary for each regenerator model, as shown in Figure 7.2.1. The recycle
operation configurations were identical for the regenerator models for CO2 trains #1 to #4 and for the
regenerator models for CO2 trains #5 to #7.

Like the absorber column models, the PR and the Sour PR property packages from Chapter 6 were
used to perform the physical and thermodynamic property calculations for the regenerator models.
The PR property package was applied to the sections of the models in which the hypothetical K2CO3*
component was present, namely the main packed sections, the solution reboilers and the regenerator
sumps. Where K2CO3* was assumed to be absent (such as in the wash sections, the overhead
condenser system, the cooling water circuits and the reboiler steam circuit), the property calculations
were performed using the Sour PR property package. As for the absorber models, the T-P flash
transfer basis was applied to all changes between the two property packages.

It should be noted that although the above assumption regarding the absence of K2CO3* in the
regenerator wash sections and overhead condenser system is theoretically valid, it is not always true
in practice (Willcocks, 2008). The regenerator condensate streams have been found to contain very
small quantities of potassium carbonate due to some entrainment of the potassium carbonate solution
at the top of the regenerator columns. Likewise, it is possible that, in practice, some potassium
carbonate solution is also entrained at the top of the absorber columns, leading to the presence of
potassium carbonate solution in the saturated sweet gas streams, and therefore the sour water
streams.

Entrainment can be accounted for in the HYSYS® column operations by reducing the column stage
efficiencies such that non-ideal separation of the various components occurs. However, this approach
was unsuccessful for the hypothetical K2CO3* component as negligible carryover of K2CO3* was
obtained for the absorber and regenerator column models. This lack of entrainment was attributed to
the high boiling point 1 for K2CO3* as the use of lower values for the boiling point was found to increase
the concentration of K2CO3* in the vapour phase. However, changes to the boiling point had
undesirable effects on the other physical and thermodynamic properties for K2CO3* and on the phase
equilibrium calculations for the potassium carbonate system. Consequently, given the relatively low
levels of potassium carbonate reported in the sour water and condensate streams, it was considered
reasonable to neglect the entrainment of potassium carbonate solution in this work.

1
In actuality, the melting point of K2CO3 was used as the boiling point of K2CO3* since K2CO3 decomposes at
temperatures near the melting point.

125
Chapter 7: The Absorber and Regenerator Column Models
126

(a) (b)
Figure 7.2.1: Process flow diagrams of the two most dissimilar regenerator column models in HYSYS®. (a) The regenerator model for CO2 train #1. (b) The
regenerator model for CO2 train #7. The // denote a change in property package.
Chapter 7: The Absorber and Regenerator Column Models

7.3 Preliminary Column Model Simulations


A series of preliminary simulations were undertaken to investigate the performance of the individual
HYSYS® absorber and regenerator column models. The results obtained from these column model
simulations are discussed below.

7.3.1 Equilibrium Stage Model Performance


Using the first set of plant operating data provided in Table 2.1.2 and the default column stage
efficiencies of 100% (i.e. equilibrium stages, K = 1), steady-state simulations of the absorber and
regenerator columns in the Moomba CO2 trains were performed. Figure 7.3.1 shows the acid gas
composition (in the vapour phase for the absorbers and in the liquid phase for the regenerators) and
temperature profiles predicted for the columns in CO2 trains #1 and #7. The corresponding profiles for
the remaining CO2 trains were very similar to these, and are provided in Figures H.1.1 and H.1.2 in
Appendix H.

Excessively high levels of acid gas absorption were predicted in all seven absorbers, with an average
deviation of -82.0% between the predicted sweet gas CO2 content and the plant data. The vapour
phase composition profiles indicated extremely rapid removal of CO2 and H2S near the bottom of the
absorbers, considerably more so than observed for the corresponding Aspen Custom Modeler
simulations® in Chapter 4. This was reflected in the column temperature profiles, which featured
sharper temperature bulges than observed for the Aspen Custom Modeler simulations®.

In contrast, much better agreement was observed between the plant data and the predicted profiles
for the equilibrium stage HYSYS® regenerator models. The predicted CO2 loading values for the
regenerated solution were found to be within ±7% of the corresponding data values. The liquid phase
composition profiles and the phase temperature profiles were also found to agree well with those from
the corresponding Aspen Custom Modeler simulations® in Chapter 4.

The results of these equilibrium stage simulations indicated that the equilibrium stage approach was
suitable for the CO2 train regenerators. However, the equilibrium stage approach was inadequate for
modelling the absorber columns since it greatly over-predicted the level of acid gas absorption. This
suggested that decreasing the absorber stage efficiencies to below 100% (i.e. non-equilibrium stages,
K < 1) would reduce the predicted acid gas absorption and therefore more accurately reflect the actual
performance of the CO2 train absorbers.

127
Chapter 7: The Absorber and Regenerator Column Models

CO2
CO2 H2S
H2S VapourPhase
Vapour Phase LiquidPhase
Liquid Phase
CO2 Data
CO2 Data H2S Data
H2S Data VapourData
Vapour Data LiquidData
Liquid Data

Regenerator H2S Loading


0.E+00 1.E-05 2.E-05 3.E-05
Absorber H2S (ppm) Cond
1.2
0 10 20 30 WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4
0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 30 50 70 90 110 130 0.0 0.2 0.4 0.6 0.8 60 70 80 90 100 110 120

Absorber CO2 (mol%) Absorber Temperature (°C) Regenerator CO2 Loading Regenerator Temperature (°C)

(a)
Regenerator H2S Loading
0.E+00 1.E-05 2.E-05 3.E-05
Absorber H2S (ppm) Cond
1.2
0 10 20 30 WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4
0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 90 100 110 120 130 0.0 0.2 0.4 0.6 0.8 60 70 80 90 100 110 120

Absorber CO2 (mol%) Absorber Temperature (°C) Regenerator CO2 Loading Regenerator Temperature (°C)

(b)
Figure 7.3.1: Equilibrium stage simulation results for the absorber and regenerator columns. (a) CO2 train
#1. (b) CO2 train #7. (Cond = condenser, WS = wash section, Reb = reboiler) Note: The CO2 and H2S
-1
loadings are not strictly dimensionless; their units are actually (mol CO2)·(mol K2CO3) and (mol H2S)·(mol
-1
K2CO3) .

7.3.2 Sensitivity to Column Stage Efficiencies


To test the above hypothesis, a series of simulation runs were performed for the absorber columns in
CO2 trains #1 and #7 in which the column stage efficiencies were decreased by 25, 50, 75 and 95%.
Similar simulation runs were also undertaken for the corresponding regenerator columns to determine
if their performance could be further improved through the adjustment of the stage efficiencies. The
results of these sensitivity analyses are shown in Figures 7.3.2 to 7.3.5. CO2 trains #2 to #6 were not
included in this study as their absorber and regenerator columns were observed to behave very
similarly to those in the two tested trains.

From Figure 7.3.2, it was evident that the rate of acid gas absorption decreased with decreasing
absorber stage efficiencies. Correspondingly, the acid gas content of the sweet gas increased with
the decreasing stage efficiencies. However, the CO2 absorption rate and the sweet gas CO2 content

128
Chapter 7: The Absorber and Regenerator Column Models

were only observed to be significantly affected by stage efficiencies below 25%. In contrast, any
reduction in the stage efficiencies resulted in the sweet gas H2S content increasing by several orders
of magnitude. Nevertheless, the effect of the stage efficiencies on the sweet gas H2S content was still
relatively inconsequential since the sweet gas H2S content was extremely low ( 15 ppm).

Figure 7.3.3 compares the absorber column profiles for stage efficiencies of 5, 25 and 100% (K = 0.05,
0.25 and 1). It was observed that reducing the stage efficiencies from 100 to 25% resulted in
composition and temperature profiles that more closely resembled those obtained from the Aspen
Custom Modeler® simulations in Chapter 4. However, decreasing the stage efficiencies to 5% led to
the significant under-absorption of CO2. This indicated the performance of the HYSYS® absorber
models could be made to more accurately reflect that of the actual CO2 train absorbers if column stage
efficiencies between 5 and 25% were used.

Like the absorber models, it was found that the smaller the column stage efficiencies, the more
significant their effect on the regenerator models. However, the relative changes observed for the acid
gas desorption rate and on the regenerated solution CO2 loading were still relatively minor, being no
more than ±12%, as shown in Figure 7.3.4. The relative changes in the regenerated solution H2S
loading were of a few orders of magnitude, but were still relatively inconsequential since the H2S
loading was extremely low (around the order of 10-9).

Figure 7.3.5 shows how the regenerator column profiles became increasingly distorted as the stage
efficiencies decreased from 100 to 5%, resulting in poorer agreement with the profiles obtained for the
Aspen Custom Modeler® simulations in Chapter 4. This suggested that unlike the absorber models,
the performance of the regenerator models could not be improved through the use of the column
stage efficiencies.

As for the regenerator models developed in Aspen Custom Modeler®, adjustment of the steam flow to
the solution reboilers was considered a more appropriate method of improving the performance of the
HYSYS® regenerator models. Figure 7.3.6 illustrates the effect of ±15% changes on the performance
of the equilibrium stage HYSYS® regenerator models. The regenerator models were found to be
highly sensitive to changes in the reboiler steam flow; however, these changes did not cause any
distortion of the column profiles, unlike the changes to the column stage efficiencies. These findings
confirmed the suitability of using equilibrium stages for the regenerator column models. They also
reflected the ineffectiveness of the model adjustments on the Aspen Custom Modeler® regenerator
models, compared to the corresponding absorber models.

129
Chapter 7: The Absorber and Regenerator Column Models

CO2
CO2 -(CO
Absorber
2 train #1)
1 CO2
CO2 -(CO
Absorber
2 train #7)
7 H2S
H2S -(CO
Absorber
2 train #1)
1 H2S
H2S -(CO
Absorber
2 train #7)
7

-25

Change in Stage
Efficiency (%)
-50

-75

-95

-90 -80 -70 -60 -50 -40 -30 -20 -10 0

Relative Change in Total Absorption Rate (%)

(a)
Change in Stage

-25
Efficiency (%)

-50

-75

-95

0 500 1000 1500 2000 2500

Relative Change in Sweet Gas CO2 Content (%)

(b)
Change in Stage

-25
Efficiency (%)

-50

-75

-95

1.E+00 1.E+02 1.E+04 1.E+06 1.E+08 1.E+10 1.E+12 1.E+14 1.E+16 1.E+18 1.E+20 1.E+22 1.E+24

Relative Change in Sweet Gas H2S Content (%)

(c)
Figure 7.3.2: Sensitivity analysis results for the absorber models. Relative changes in (a) the total
absorption rates, (b) the sweet gas CO2 content and (c) the sweet gas H2S content.

130
Chapter 7: The Absorber and Regenerator Column Models

100% Efficiency 25% Efficiency 5% Efficiency Data

1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 30 50 70 90 110 130 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(a)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 130 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(b)
Figure 7.3.3: Effect of the column stage efficiencies on the absorber composition and temperature
profiles. (a) CO2 train #1. (b) CO2 train #7.

131
Chapter 7: The Absorber and Regenerator Column Models

CO2
CO2 -(CO
Absorber
2 train #1)
1 CO2
CO2 -(CO
Absorber
2 train #7)
7 H2S
H2S -(CO
Absorber
2 train #1)
1 H2S
H2S -(CO
Absorber
2 train #7)
7

-25

Change in Stage
Efficiency (%)
-50

-75

-95

-10 -8 -6 -4 -2 0 2 4

Relative Change in Total Desorption Rate (%)

(a)
Change in Stage

-25
Efficiency (%)

-50

-75

-95

-6 -4 -2 0 2 4 6 8 10 12

Relative Change in Regenerated Solution CO2 Loading (%)

(b)
Change in Stage

-25
Efficiency (%)

-50

-75

-95

1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07 1.E+08 1.E+09

Relative Change in Regenerated Solution H2S Loading (%)

(c)
Figure 7.3.4: Sensitivity analysis results for the regenerator models. Relative changes in (a) the total
desorption rates, (b) the regenerated solution CO2 loading and (c) the regenerated solution H2S loading.

132
Chapter 7: The Absorber and Regenerator Column Models

100% Efficiency 25% Efficiency 5% Efficiency Data

1.2
Cond
Fraction of Total Packed Height WS

1.0

0.8

0.6

0.4

0.2

0.0
Reb
-0.2
0.0 0.2 0.4 0.6 0.8 1.0 0.E+00 2.E-05 4.E-05 6.E-05 60 80 100 120 90 100 110 120 130
CO2 Loading H2S Loading Vapour Phase Temperature (°C) Regenerator Temperature (°C)

(a)
1.2
Cond
WS
Fraction of Total Packed Height

1.0

0.8

0.6

0.4

0.2

0.0
Reb
-0.2
0.0 0.2 0.4 0.6 0.8 1.0 0.E+00 2.E-05 4.E-05 6.E-05 60 80 100 120 140 90 100 110 120 130
CO2 Loading H2S Loading Vapour Phase Temperature (°C) Regenerator Temperature (°C)

(b)
Figure 7.3.5: Effect of the column stage efficiencies on the regenerator composition and temperature
profiles. (a) CO2 train #1. (b) CO2 train #7. (Cond = condenser, WS = wash section, Reb = reboiler)

133
Chapter 7: The Absorber and Regenerator Column Models

CO2
CO2desorption
desorptionrate
rate(co2
(CO2train
train#1)
#1) solution co2 loading
CO2 desorption rate(co2
(COtrain #1)
2 train #7)
Solutionco2
solution COloading
2 loading (COtrain
(co2 2 train #1)
#1) Solution
co2 CO2 loading
desorption (COtrain
rate (co2 2 train
#7)#7)

Change in Steam
15

Flow (%)
-15

-15 -10 -5 0 5 10 15

Relative Change (%)

(a)

Unadjusted Steam 15% Steam -15% Steam Data

1.2
Cond
WS
Fraction of Total Packed Height

1.0

0.8

0.6

0.4

0.2

0.0
Reb
-0.2
0.0 0.2 0.4 0.6 0.8 1.0 0.E+00 1.E-05 2.E-05 3.E-05 60 80 100 120 90 100 110 120 130
CO2 Loading H2S Loading Vapour Phase Temperature (°C) Regenerator Temperature (°C)

(b)
1.2
Cond
WS
Fraction of Total Packed Height

1.0

0.8

0.6

0.4

0.2

0.0
Reb
-0.2
0.0 0.2 0.4 0.6 0.8 1.0 0.E+00 1.E-05 2.E-05 3.E-05 60 80 100 120 90 100 110 120 130
CO2 Loading H2S Loading Vapour Phase Temperature (°C) Regenerator Temperature (°C)

(c)
Figure 7.3.6: Effect of the reboiler steam flow on the regenerator column performance. The effect on (a)
the total CO2 desorption rate and the regenerated solution CO2 loading, and the effect on the regenerator
composition and temperature profiles for (b) CO2 train #1 and (c) CO2 train #7. (Cond = condenser, WS =
wash section, Reb = reboiler)

134
Chapter 7: The Absorber and Regenerator Column Models

7.4 Column Stage Efficiency Correlations


Based on the findings of the above sensitivity study, it was decided that adjustment of the column
stage efficiencies was only required for the absorber column models. The results of the Aspen
Custom Modeler® parametric studies in Chapter 5 were used to develop stage efficiency correlations
of the form:
2 2
E stage C 0  C1 ˜ TRG  C 2 ˜ TRG  C 3 ˜ Ĝ RG  C 4 ˜ Ĝ RG  C 5 ˜ 100 ˜ y RG,CO2
2 2 2
 C 6 ˜ 100 ˜ y RG,CO2  C 7 ˜ TLS  C 8 ˜ TLS  C 9 ˜ L̂ LS  C10 ˜ L̂ LS
(7.4.1)
2
 C11 ˜ 100 ˜ wf LS,K 2CO3  C12 ˜ 100 ˜ wf LS,K 2CO3  C13 ˜ FLS,CO2
2 2
 C14 ˜ FLS,CO2  C15 ˜ PAbs  C16 ˜ PAbs

ˆ is the
where Estage is the overall stage efficiency (between 0 and 1), T is the temperature in °C, G

volumetric vapour phase flow in sm3/h, Lˆ is the volumetric liquid phase flow in m3/h, y CO2 is the dry

gas CO2 mole fraction, wf K 2CO3 is the equivalent weight fraction of K2CO3, F CO2 is the solution CO2

loading, and PAbs is the average absorber column pressure in bar. The subscripts RG and LS denote
respectively the raw gas and lean solution entering the absorber.

The overall stage efficiency approach was used instead of individual component stage efficiencies for
two reasons. Firstly, the component stage efficiencies are restricted to steady-state simulations,
whereas overall stage efficiencies are applicable to both steady-state and dynamic simulations.
Secondly and more importantly, unlike the component stage efficiencies, the overall stage efficiencies
can be specified via HYSYS® macros written in the WinWrap™ Basic programming language (which
is very similar to and is compatible with Visual Basic for Applications™). Since the stage efficiencies
were expected to change during the course of a simulation due to changes in the column conditions
as part of the convergence process, this feature was essential for the column stage efficiencies to be
automatically updated upon the calculation of new efficiency values.

To simplify the calculation and specification of the column stage efficiencies, the stages in each
absorber model were assumed to be of the same efficiency, instead of being assigned individual
overall stage efficiency values. This approach required only one correlation for each absorber model,
as opposed to one correlation per model stage, thereby simplifying the regression of the coefficient
values. Very good results were obtained with this assumption, as indicated in Tables 7.4.1 and 7.4.2
and Figures 7.4.1 and 7.4.1. For all seven absorbers, the sweet gas CO2 content predicted using the
correlated stage efficiencies were found to be within ±10% of the corresponding Aspen Custom
Modeler® values, with average absolute deviations between 0.01 and 1.60%. Consequently, the
additional complexity associated with the specification of individual overall stage efficiencies could not
be justified.

135
Chapter 7: The Absorber and Regenerator Column Models

Table 7.4.1: Coefficients for the steady-state column stage efficiency correlations.

Absorber #1 #2 #3 #4 #5 #6 #7
1
C0 ×10 2.456 2.154 -0.019 1.895 1.997 0.700 16.467
C1 ×106 -7.943 -2.514 -14.000 -5.608 -3.300 -5.100 -55.800
C2 0.000 0.000 0.000 0.000 0.000 0.000 0.000
6
C3 ×10 -1.572 -1.355 -0.529 -0.929 -0.953 -0.591 1.261
C4 0.000 0.000 0.000 0.000 0.000 0.000 0.000
3
C5 ×10 -4.783 -1.670 7.851 -5.762 -0.103 2.968 -46.452
C6 ×104 1.080 0.450 -2.675 1.265 0.000 -0.996 9.118
4
C7 ×10 1.322 6.150 -0.367 0.139 6.014 3.653 -4.162
C8 ×106 0.000 -1.871 0.000 0.000 -2.255 -1.650 0.000
C9 ×104 3.579 0.422 7.788 3.472 0.376 2.021 -1.670
C10 ×107 -2.197 0.321 -4.742 -1.655 0.000 -0.765 1.891
4
C11 ×10 5.539 7.625 12.871 7.044 3.054 5.775 -291.884
C12 ×106 0.000 -4.773 0.000 0.000 0.000 0.000 639.126
C13 ×101 -3.704 -2.150 -7.416 -4.382 -2.111 -3.331 -31.130
C14 ×101 2.446 1.158 6.074 3.305 1.140 2.458 32.500
5
C15 ×10 -8.875 -15.650 3.175 -2.900 -11.925 -3.100 79.825
C16 0.000 0.000 0.000 0.000 0.000 0.000 0.000

MAD (%) 0.24 0.10 1.57 0.70 0.57 0.58 9.63


AAD (%) 0.06 0.01 0.18 0.11 0.12 0.08 1.60
Note: MAD = Maximum absolute deviation for the predicted sweet gas CO2 content. AAD = Average absolute deviation for the
predicted sweet gas CO2 content.

Table 7.4.2: Coefficients for the dynamic column stage efficiency correlations.

Absorber #1 #2 #3 #4 #5 #6 #7
C0 ×102 7.187 5.040 -5.054 7.433 7.736 0.706 103.800
C1 ×105 9.928 9.542 4.300 3.972 4.600 3.150 0.240
C2 ×107 -3.958 -3.601 0.000 0.000 0.000 0.000 0.000
C3 ×107 6.042 8.061 6.635 1.494 1.516 0.971 15.472
12
C4 ×10 -3.080 -4.369 -3.383 -0.662 -0.345 -0.382 -5.304
C5 ×103 -2.527 -0.848 4.315 -3.735 -0.311 1.820 -32.629
5
C6 ×10 4.942 1.444 -15.626 7.713 0.000 -6.625 65.487
C7 ×105 14.744 34.338 9.425 5.662 47.186 32.353 -23.168
7
C8 ×10 0.000 -7.754 -2.645 0.000 -15.794 -12.969 0.000
C9 ×105 17.042 1.998 45.608 21.244 2.438 13.110 -13.524
7
C10 ×10 -1.020 0.143 -2.758 -1.004 0.000 -0.494 1.355
C11 ×104 2.346 3.473 7.607 4.138 1.713 3.608 -193.693
6
C12 ×10 0.000 -2.829 -0.278 0.000 0.000 0.000 421.163
C13 ×101 -1.848 -0.940 -4.544 -2.744 -1.327 -2.164 -20.549
1
C14 ×10 1.251 0.503 3.769 2.083 0.711 1.592 21.558
C15 ×104 -6.230 -6.185 -4.469 -4.114 -5.276 -3.276 1.903
C16 0.000 0.000 0.000 0.000 0.000 0.000 0.000

MAD (%) 0.27 0.22 1.57 0.75 0.58 0.61 9.95


AAD (%) 0.07 0.03 0.18 0.11 0.10 0.07 1.64
Note: MAD = Maximum absolute deviation for the predicted sweet gas CO2 content. AAD = Average absolute deviation for the
predicted sweet gas CO2 content.

136
Chapter 7: The Absorber and Regenerator Column Models

Default Efficiencies Correlated Efficiencies Data


1.0

Fraction of Total Packed Height


0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 30 50 70 90 110 130 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(a)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 130 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(b)
Figure 7.4.1: Effect of the correlated overall stage efficiencies on the steady-state absorber columns.
(a) CO2 train #1. (b) CO2 train #7.

Default Efficiencies Steady-State Efficiencies Dynamic Efficiencies Data


1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 30 50 70 90 110 130 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(a)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 130 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(b)
Figure 7.4.2: Effect of the correlated overall stage efficiencies on the steady-state behaviour of the
dynamic absorber columns. (a) CO2 train #1. (b) CO2 train #7.

137
Chapter 7: The Absorber and Regenerator Column Models

Values for the coefficients C0 to C16 for the steady-state absorber models are given in Table 7.4.1,
while Table 7.4.2 lists the coefficient values for the dynamic absorber models. The two different sets
of coefficients were required due to the different manner in which the overall stage efficiencies are
treated in the two simulation modes, as explained in Section 6.1.3.

Both sets of coefficient values were derived via the same procedure. First, a set of overall stage
efficiencies corresponding to the results of the Aspen Custom Modeler® parametric studies were
determined by iterative adjustment of the efficiencies until the steady-state CO2 content of the sweet
gas produced by the HYSYS® absorber models matched the Aspen Custom Modeler® results to
within ±0.001%. The coefficients for equation (7.4.1) were then regressed from the resulting set of
efficiencies using the Microsoft® Excel Regression Tool.

It should be noted that while this method of correlating the overall stage efficiencies produced
excellent results for the sweet gas CO2 content, it resulted in the significant over-prediction of the H2S
content. On average, the sweet gas H2S concentrations calculated by the HYSYS® absorber models
were 94.8% greater than the corresponding Aspen Custom Modeler® values. This limitation was
however deemed acceptable for this work since the primary focus of the CO2 trains is to remove CO2.

The overall stage efficiency correlations were implemented in HYSYS® spreadsheets, like the one
shown in Figure 7.4.3, which were linked to the HYSYS® absorber models. This enabled the
automatic re-calculation of the stage efficiencies upon any changes to the relevant model process
conditions. Macros, like the one provided in Table 7.4.3, were attached to the absorber models (in the
form of HYSYS® user variables) to automatically update the absorber stage efficiencies with the newly
calculated values. Separate macros (and therefore user variables) were set up for the steady-state
and dynamic simulations since the two simulation modes involved different subroutines.

The absorber composition and temperature profiles predicted with the correlated column overall stage
efficiencies for CO2 trains #1 and #7 are given in Figures 7.4.1 and 7.4.2. The corresponding column
profiles for the other five trains are included as Figures H.1.3 to H.1.6 in Appendix H. The need for
separate steady-state and dynamic stage efficiency correlations was evident in Figure 7.4.2, which
shows the under-prediction of the sweet gas CO2 content at steady-state when the steady-state
efficiencies were used for the dynamic absorber models.

It should also be noted that the steady-state temperature profiles for the dynamic absorbers differed in
shape from those for the steady-state absorbers, most noticeably for the liquid phase temperature
profiles. While the steady-state temperatures of the rich solution and sweet gas leaving the dynamic
absorbers closely agreed with the temperatures predicted by the steady-state absorber models (within
±1°C), the distinctive temperature bulges in the lower part of the columns were absent. The absence
of these characteristic bulges did not appear to adversely affect the dynamic simulation results, and
was most likely due to the different definition of the overall stage efficiencies in dynamic simulations.

138
Chapter 7: The Absorber and Regenerator Column Models

In dynamic simulations, the overall stage efficiency determines how much of the vapour entering a
stage actually participates in the flash calculation for that stage, whereas in steady-state simulations, it
represents the deviation of the vapour phase leaving a stage from the vapour that is in equilibrium with
the liquid phase leaving that stage. As these two definitions are not directly equivalent, and given the
low stage efficiencies of the absorber models (<0.25), it was not unexpected that there should be
some dissimilarity between the results produced by the steady-state and dynamic simulations.

Figure 7.4.3: An example HYSYS® spreadsheet for calculating the absorber overall stage efficiencies.

139
Chapter 7: The Absorber and Regenerator Column Models

Table 7.4.3: An example HYSYS® macro for updating the absorber overall stage efficiencies.
'Updates the column stage efficiency for ABS-1

Option Explicit

Sub PostSolve() ‘Sub DynCompositionPreStep() for dynamic simulations

On Error Resume Next

Dim hyCase As HYSYS.SimulationCase


Dim hySprdsht As HYSYS.SpreadsheetOp
Dim hyCol As HYSYS.ColumnOp
Dim hyColStage As HYSYS.ColumnStage

Dim dblOldEff As Double


Dim dblNewEff As Double
Dim tol As Double
Dim intCount As Integer
Dim intStages As Integer

Set hyCase = ActiveCase


Set hySprdsht = hyCase.Flowsheet.Operations.Item("ABS-1 SPRDSHT")
Set hyCol = hyCase.Flowsheet.Operations.Item("V-20-A")

intStages = hyCol.ColumnFlowsheet.ColumnStages.Count
dblNewEff = hySprdsht.Cell("G23").CellValue
dblOldEff = hySprdsht.Cell("G24").CellValue
tol = 1e-5

'Compare old and new values of calculated stage efficiency. If less than tol, don't update.
If Abs((dblNewEff-dblOldEff)/dblOldEff) <= tol Then
Exit Sub
Else
'Don't update if calculated stage efficiency is outside the the range (0,1)
If dblNewEff <= 0 Or dblNewEff > 1 Then
Exit Sub
End If
'Update stage efficiency
For intCount = 0 To (intStages-1)
Set hyColStage = hyCol.ColumnFlowsheet.ColumnStages(intCount)
hyColStage.SeparationStage.OverallEfficiencyValue = dblNewEff
Next
'Send update alert to Trace Window
hyCase.Application.Trace "ABS-1 Stage Efficiency modified from " & dblOldEff & " to " & dblNewEff, False
End If

'Clear objects
Set hyCase = Nothing
Set hySprdsht = Nothing
Set hyCol = Nothing
Set hyColStage = Nothing

End Sub

140
Chapter 7: The Absorber and Regenerator Column Models

7.5 Summary
In summary, this chapter focussed on the development of the absorber and regenerator column
models for the HYSYS® CO2 train process models. Particular attention was given to the correlation
and implementation of column stage efficiencies to better represent the effects of the solution
chemistry.

Detailed column models were constructed for the absorbers and regenerators in the CO2 trains. The
standard HYSYS® absorber operation formed the basis of both types of column models by
representing the packed and/or tray sections of these columns. In addition, a variety of other standard
HYSYS® operations, such as separator, pump and heat exchanger operations, were used to model
various auxiliary parts, including the column sumps, circulation pumps and solution reboilers. The two
property packages selected in Chapter 6, i.e. the PR and Sour PR property packages, were applied to
different sections of the column models. The former was used in sections where the hypothetical
K2CO3* component was present, while the latter was applied to sections in which it was absent.

Preliminary simulations were performed to investigate the performance of the column models using
the default equilibrium stages, and to determine the effect of varying the column stage efficiencies. It
was observed that the equilibrium stage approach was suitable for the regenerator column models,
but not for the absorber column models. Significant improvements in the performance of the absorber
column models were obtained by reducing the stage efficiencies. In contrast, the reboiler steam flow
was found to have a greater effect on the performance of the regenerator column models. This was
analogous to the Aspen Custom Modeler® column models, in that the model adjustments were only
useful for the absorber models.

Consequently, the column stage efficiency correlations were only applied to the HYSYS® absorber
column models. These correlations were developed from the results of the Aspen Custom Modeler®
parametric studies in Chapter 5, thereby linking the HYSYS® absorber models to the rigorous rate-
based non-equilibrium models from Chapter 4. Two sets of correlations were required, i.e. one for
steady-state simulations and another for dynamic simulations, due to the differing definitions for the
column stage efficiencies for the two simulation modes. These correlations were implemented via
spreadsheet operations, which were linked to the absorber column models through HYSYS® user
variables. This enabled the automatic updating of the stage efficiencies during the calculation
process. Very good results were obtained with the correlated stage efficiencies.

The above absorber and regenerator column models form the basis of the CO2 train process models
in HYSYS®. The next chapter considers the development of steady-state HYSYS® process models
for the Moomba CO2 trains.

141
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

CHAPTER 8

HYSYS® CO2 REMOVAL TRAIN


PROCESS MODEL
DEVELOPMENT

This chapter concerns the development of steady-state HYSYS® process models of the Moomba CO2
trains. Particular attention is paid to the extension of the absorber and regenerator column models
that were developed in the previous chapter.

The configurations of the Moomba CO2 trains can be divided into two key sections: the absorption
circuits and the regeneration circuits. In this chapter, the absorber and regenerator column models
from Chapter 7 are expanded in order to develop HYSYS® models of these absorption and
regeneration circuits. Particular focus is given to the inclusion of the various ancillary operations, such
as the solution pumpset, and the representation of the relevant piping and head losses.

Complete steady-state HYSYS® CO2 train process models are then created by linking the above
absorption and regeneration circuit models. As in the case of the Aspen Custom Modeler® CO2 train
models, these HYSYS® process models are compared with the plant data used in the development of
the Aspen Custom Modeler® models, and are also validated against an independent set of plant data.

142
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

8.1 Ancillary Operation Models


While the absorber and regenerator column models developed in the previous chapter formed the
basis of the HYSYS® CO2 train process models, a number of ancillary operations were also required
to complete the process models. The following sections describe the development of such ancillary
operation models.

8.1.1 Absorption Circuit Ancillary Models


The absorption circuits encompassing the absorber columns constitute a key part of the Moomba CO2
trains. In order to model these absorption circuits in HYSYS®, the absorber column models from
Chapter 7 were expanded to include a sweet gas separator and a gas-gas heat exchanger (only in
trains #3 to #7), as shown in Figure 8.1.1.

(a) (b)
Figure 8.1.1: HYSYS® process flow diagram of the CO2 train absorption circuits. (a) CO2 trains #1 and #2.
(b) CO2 trains #3 to #7. The // denote a change in property package.

HYSYS® separator operations were used to model the sweet gas separators, while heat exchanger
operations were used to represent the gas-gas heat exchangers. The relevant information required to
size the separator and heat exchanger operations was determined from the available equipment
datasheets and manuals in the Santos TIMS controlled documentation database.

Given that the gas-gas heat exchanger models did not involve the potassium carbonate solution, their
thermodynamic and physical property calculations were performed using the Sour PR property
package. This property package was also applied to the sweet gas separator models since it was
assumed that there was no entrainment of K2CO3* at the top of the absorber columns, as previously
discussed in Section 7.2.

143
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

8.1.2 Regeneration Circuit Ancillary Models


Another key section of the CO2 trains are the regeneration circuits encircling the regenerator columns.
In order to model these regeneration circuits in HYSYS®, the regenerator column models from
Chapter 7 were expanded to include a solution pumpset model, which consisted of a main solution
pump linked to a steam turbine and to a power recovery turbine (a hydraulic turbine) with a bypass
valve, as illustrated in Figure 8.1.2.

A HYSYS® pump operation was used to model the main solution pump, while the steam turbine was
represented by an expander operation. The pump and expander operations were both considered for
the power recovery turbine since hydraulic turbines are not a default unit operation in HYSYS®. The
pump and expander operations were found to give fairly similar results for the power recovery turbine,
with the expander operation predicting a higher power recovery (by  10%) and a lower output
temperature (by  -1%) than the pump operation. Warning messages were generated about the
presence of liquid in the inlet stream for the expander operation and the negative head for the pump
operation, but these were non-fatal and could be safely ignored (Stone, 2005).

In the end, it was considered more appropriate to use the pump operation to represent the power
recovery turbine for two reasons. First, hydraulic turbines behave like pumps in reverse and second,
the expander operation is specific to gas expansion which is an isentropic process, unlike liquid
expansion.

Figure 8.1.2: HYSYS® process flow diagram of the solution pumpset model.

The components of the solution pumpset model were sized according to the technical datasheets and
equipment manuals available in the Santos TIMS database. The performance curves for the steam

144
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

turbine were not in the form required for the expander operation calculations, so the power output and
steam input for the steam turbine were instead calculated via the HYSYS® spreadsheet shown in
Figure 8.1.3.

Figure 8.1.3: HYSYS® spreadsheet for the pumpset calculations.

This spreadsheet was also used for the power recovery turbine calculations in the steady-state
simulation mode. Unlike in the dynamic simulation mode, there is no option in the steady-state mode
to enable the pump operation to act as a hydraulic turbine. Consequently, the pump operation could
not calculate the appropriate pressure drops from the positive head values for the power recovery
turbine characteristic curves. Attempts to treat the turbine characteristic curves as having negative
head values failed since these negative values could not be handled by the pump operation
calculations. As a result, the “Use Curves” option was deactivated for the power recovery turbine
models in the steady-state simulation mode, and the relevant characteristic curves were instead
implemented in the spreadsheet.

In contrast, the spreadsheet calculations were not necessary for the power recovery turbine models in
the dynamic simulation mode. Instead, the option “Pump is acting as turbine” was activated, which
enabled the pump operation to calculate the pressure drop and recovered power from the
characteristic curves. For this approach to work, the turbine head values had to be entered as positive
values.

The physical and thermodynamic property calculations for the main solution pump and the power
recovery turbine were performed with the modified PR property package. The Sour PR property

145
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

package was used for the steam turbine calculations since the steam turbine did not involve the
potassium carbonate solution.

8.1.3 Control Valves and Piping Losses


A number of control valves were required for the HYSYS® CO2 train process models. These were
modelled using HYSYS® valve operations and included flow control valves for the raw gas and
reboiler steam, as well as level control valves for the various vessels in the absorption and
regeneration circuits. The relevant sizing information for these valves was obtained from the Santos
TIMS database.

Valve operations were also used to represent the piping pressure losses in various parts of the CO2
train process models, such as in the cooling water and overhead condensate circuits. Also included
were the head losses due to the distance the potassium carbonate solution had to travel to reach the
top of the absorber and regenerator columns. The sizing information required for these valve
operations was determined from the available plant specifications in the Santos TIMS database. It
should be noted that the piping losses could have instead been modelled by HYSYS® pipe segment
operations, but were not since the additional complexity associated with the pipe segment operations
was not considered necessary for this work.

146
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

8.2 Steady-State CO2 Train Models


Detailed steady-state models of the Moomba CO2 trains were developed in HYSYS® by linking the
previously described absorption and regeneration circuit process models with the various valve
operations. The resulting model configurations for CO2 trains #1 and #7 are shown in Figures 8.2.1
and 8.2.2, while the corresponding process flow sheets for the other five trains are included as Figures
H.2.1 to H.2.5 in Appendix H.

In HYSYS®, the only flow measurements that may be user-specified for a material stream are the
molar flow, mass flow, standard ideal liquid volume flow and liquid volume flow at standard conditions,
of which only one may be specified for each material stream to avoid specification conflicts. Other
flow measurements, such as the standard gas flow and the actual volume flow, can only be calculated
from the stream properties and the specified flow. However, the flow data for the CO2 trains were in
terms of the standard gas flow for the gas streams and the actual volume flow for the liquid streams.
Consequently, adjust logical operations were attached to the unheated raw gas streams entering the
absorber columns to manipulate the user-specified molar flows to obtain the desired standard gas
flows.

Adjust operations were similarly set up for the lean solution streams entering the absorbers so as to
achieve the required actual volume flows. To accommodate these adjust operations, a tee operation
was used in each CO2 train process model to split the solution stream leaving the main solution pump
into an absorber feed stream and a material balance stream. This enabled the molar flow of the lean
solution streams to be user-specified, instead of being calculated by the main solution pumps. Manual
adjustment of the makeup water flow to the regenerators was required to ensure negligible calculated
molar flows for the material balance streams.

This application of the tee operations caused the three streams connected to each tee operation to
share the same composition calculated by the preceding main solution pumps. Consequently,
additional adjust operations were required to obtain the desired CO2 loading for the lean solution by
manipulating the user-specified reboiler steam mass flows.

Two alternative methods to the above tee operations were also considered. The first method replaced
the tee operations with recycle operations, which made the lean solution stream properties specifiable
and eliminated the need for the material balance streams. However, this approach was not
undertaken as it was not recommended that recycle operations be attached to streams which have
variables determined by an adjust operation (Hyprotech Ltd, 2001c).

In the second option, the tee operations were replaced by mixer operations which combined the
pressurised solution and material balance streams to produce the lean solution streams. As with the
tee operations, the reboiler steam adjust operations were required to obtain the desired lean solution
CO2 loading, and manual adjustment of the makeup water flow was needed to ensure negligible flow

147
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

for the material balance streams. However with the mixer operations, multiple set operations were
also necessary in order to set the temperature and composition of each lean solution stream equal to
that for its corresponding pressurised solution stream. Due to its additional complexity, this approach
was rejected in favour of the one with the tee operations.

148
Chapter 8: HYSYS® CO2 Removal Train Process Model Development
149

Figure 8.2.1: Process flow diagram for the steady-state HYSYS® model of CO2 train #1.
Chapter 8: HYSYS® CO2 Removal Train Process Model Development
150

Figure 8.2.2: Process flow diagram for the steady-state HYSYS® model of CO2 train #7.
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

8.3 Model Validation


The performance of the above steady-state HYSYS® CO2 train process models was compared
against both sets of typical operating data in Table 2.1.2. It should be noted that the first data set was
used in the development of the Aspen Custom Modeler® absorber and regenerator column models,
while the second set of data was used solely for model validation purposes.

The resulting vapour and liquid phase composition profiles for the absorber and regenerator columns
in CO2 trains #1 and #7 are given in Figures 8.3.1 and 8.3.2, and the column temperature profiles are
shown in Figure 8.3.3. The corresponding profiles for the other five CO2 trains are very similar to
these, and are included as Figures H.3.1 to H.3.6 in Appendix H. Values for the operating parameters
calculated by the CO2 train models are provided in Tables 8.3.1 and 8.3.2, along with their percentage
deviation from the data values listed in Table 2.1.2.

For all seven CO2 trains, the CO2 and H2S vapour phase concentrations were seen to decrease up the
absorber columns, with the acid gas content initially declining at a slightly faster rate near the bottom
of the columns. The reverse trend was observed for the liquid phase acid gas loading. These
HYSYS® absorber composition profiles agreed well with their corresponding Aspen Custom Modeler®
profiles from Chapter 4, with the exception of the CO2 curves for CO2 train #7 and the extent of H2S
removal predicted for the seven CO2 trains.

In the Aspen Custom Modeler® model of CO2 train #7, there was a sharp decrease in the rates at
which the CO2 levels changed in the lower third of the absorber column due to the very high CO2
loading in this part of the column which reduced the solution capacity for further CO2 absorption. The
absence of this behaviour in the corresponding HYSYS® model was an indication of the limitations of
the stage efficiency approach in representing the solution chemistry in HYSYS®. This suggested that
caution must be taken when considering the results for simulations with high raw gas to lean solution
flow ratios. In particular, the predicted rich solution CO2 loading should be checked to ensure it does
not exceed 1 (mol CO2)·(mol K2CO3)-1.

Another sign of the limitations associated with the stage efficiency approach was that the sweet gas
H2S concentrations predicted by the HYSYS® models were found to exceed those for the Aspen
Custom Modeler® models by approximately 95%. This resulted from the stage efficiency correlations
being based solely on the sweet gas CO2 content, due to the inability to use separate component
efficiencies. However, it should be noted that the HYSYS® values for the sweet gas H2S content were
still within the range specified by the plant data.

The general shape of the HYSYS® regenerator composition profiles matched that of the
corresponding Aspen Custom Modeler® profiles. CO2 was rapidly removed from the rich solution near
the bottom of the regenerator columns due to contact with the stripping steam, whereas the desorption
of H2S occurred swiftly near the top of the columns. As for the Aspen Custom Modeler® profiles, a

151
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

comparison of the HYSYS® absorber and regenerator liquid phase profiles in Figures 8.3.1 and 8.3.2
showed that a sizeable fraction of the absorbed acid gases was flashed off due to the depressurisation
of the rich solution. The HYSYS® models predicted that up to 43% of the absorbed CO2 and up to
58% of the absorbed H2S were flashed off, which were considerably less than the Aspen Custom
Modeler® model predictions of up to 64% and 86%, respectively, for CO2 and H2S. However, both
predictions for the quantity of CO2 flashed off during the solution depressurisation were within the
literature range of one- to two-thirds of the absorbed CO2 (Kohl and Nielson, 1997).

The observed discrepancies for the quantities of flashed acid gas were most likely due to the poorer
thermodynamic representation of the CO2-K2CO3-H2O system by the enhanced PR equation of state
compared to the Electrolyte NRTL model. However, despite the higher levels of residual acid gas in
the low pressure rich solution for the HYSYS® models, there was good agreement (within ±10%)
between the quantities of reboiler steam required by the HYSYS® and Aspen Custom Modeler®
models to regenerate the solution.

The absorber and regenerator temperature profiles also corresponded well with the equivalent Aspen
Custom Modeler® profiles in Chapter 4. The absorber profiles showed characteristic temperature
bulges near the bottom of the columns that were similar in shape to those in the Aspen Custom
Modeler® profiles, except for CO2 train #7. The HYSYS® profiles for this CO2 train closely resembled
those for the other six CO2 trains, unlike the corresponding Aspen Custom Modeler® profiles. This
reflected the above-mentioned differences between the HYSYS® and Aspen Custom Modeler®
absorber composition profiles for CO2 train #7. As for the Aspen Custom Modeler® process models,
temperature drops between 7 and 15°C accompanied the depressurisation of the rich solution, and the
regenerator temperatures increased down the columns, reaching a maximum at the solution reboilers
and a minimum at the overhead condensers.

For all seven CO2 trains, good agreement was observed between the HYSYS® model results and the
two sets of plant data. The model and data values for most of the operating parameters were within
±6%, with larger deviations of up to ±10% for the reboiler steam flows. The model values for the
sweet gas H2S content, the rich solution H2S loading and the column pressure drops were found to be
within the given range, while the predicted lean solution H2S loading values were effectively zero like
the plant data, being of the order of 10-8 to 10-25. The most significant deviations were associated with
the predicted acid gas temperatures for CO2 train #4, which were up to 30% lower than the data
values. However, as mentioned in Section 4.4, the data values were unexpectedly high, and since the
HYSYS® temperature values were within ±1% of the values predicted by the Aspen Custom Modeler®
models, these deviations were considered to be acceptable.

In conclusion, the agreement between the model results and the plant data was deemed sufficiently
close that the HYSYS® process models of the Moomba CO2 trains were regarded as suitable for the
steady-state simulation of the actual CO2 trains, within the range of operating parameters considered

152
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

during the model development process. To demonstrate a potential application of these HYSYS®
CO2 train process models, the steady-state models for CO2 trains #1 and #7 are converted into
dynamic process models in the next chapter.

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05 Cond
1.2
1.0 WS

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
Cond
1.2
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05 8.E-05
WS
1.0
Fraction of Total Packed Height

1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Figure 8.3.1: CO2 and H2S vapour and liquid phase profiles for the first set of data. (a) CO2 train #1.
(b) CO2 train #7. (Cond = condenser, WS = wash section, Reb = reboiler)

153
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05
Cond
1.2
0 5 10 15 20 0.E+00 1.E-05 2.E-05 3.E-05 4.E-05
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8 0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 1.E-05 2.E-05 3.E-05
Cond
1.2
0 5 10 15 20 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Figure 8.3.2: CO2 and H2S vapour and liquid phase profiles for the second set of data. (a) CO2 train #1.
(b) CO2 train #7. (Cond = condenser, WS = wash section, Reb = reboiler)

154
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

Vapour Phase Liquid Phase Vapour Data Liquid Data

1.2
Cond 1.2
Cond
Fraction of Total Packed Height WS WS

Fraction of Total Packed Height


1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
30 50 70 90 110 130 60 70 80 90 100 110 120 30 50 70 90 110 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(a) (c)
1.2
Cond 1.2
Cond
WS WS
Fraction of Total Packed Height

Fraction of Total Packed Height


1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
80 90 100 110 120 130 60 70 80 90 100 110 120 80 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(b) (d)
Figure 8.3.3: Vapour and liquid phase temperature profiles for the two sets of plant data. (a) CO2 train #1
and (b) CO2 train #7 profiles for the first set of data. (c) CO2 train #1 and (d) CO2 train #7 profiles for the
second set of data. (Cond = condenser, WS = wash section, Reb = reboiler)

155
Chapter 8: HYSYS® CO2 Removal Train Process Model Development
Table 8.3.1: CO2 train simulation results for the first set of plant data in Table 2.1.2. The bracketed values are the percentage deviations from the data.

CO2 Train #1 #2 #3 #4 #5 #6 #7
Sweet Gas
a
mol% CO2 2.51 (0.4%) 2.51 (0.4%) 2.53 (1.2%) 2.53 (1.2%) 2.52 (0.8%) 2.53 (1.2%) 2.55 (2.0%)
f f f f f f f
ppm H2S 2.80 (- ) 2.70 (- ) 2.66 (- ) 2.63 (- ) 2.54 (- ) 2.62 (- ) 2.76 (- )
Temperature (°C) 110.9 (0.8%) 110.8 (0.7%) 111.9 (1.7%) 112.0 (1.8%) 112.3 (2.1%) 112.1 (1.9%) 112.2 (2.0%)
Acid Gas
6 3 b
CO2 (10 sm /d) 0.320 (0.0%) 0.270 (0.0%) 0.479 (-0.2%) 0.549 (-0.2%) 0.549 (-0.2%) 0.669 (-0.2%) 0.907 (-0.3%)
c
Temperature (°C) 104.5 (1.5%) 105.3 (2.2%) 78.8 (-1.5%) 76.9 (-25.4%) 95.1 (0.1%) 94.4 (0.4%) 93.3 (-0.7%)
Lean Solution
d
CO2 loading 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%) 0.380 (0.0%)
-16 d g g g g g g g
H2S loading (×10 ) 4572.2 (- ) 4835.1 (- ) 174.9 (- ) 184.7 (- ) 85.1 (- ) 6.0 (- ) 0.2 (- )
e
wt% K2CO3 27.0 (0.0%) 27.0 (0.0%) 30.0 (0.0%) 30.0 (0.0%) 30.0 (0.0%) 30.0 (0.0%) 30.0 (0.0%)
Flow (m3/h) 550.0 (0.0%) 550.0 (0.0%) 670.0 (0.0%) 812.0 (0.0%) 1000.0 (0.0%) 1000.0 (0.0%) 1068.0 (0.0%)
Temperature (°C) 110.8 (0.7%) 110.8 (0.7%) 111.9 (1.7%) 112.0 (1.8%) 112.2 (2.0%) 112.1 (1.9%) 112.1 (1.9%)
156

Rich Solution
CO2 loading d 0.825 (0.6%) 0.756 (0.7%) 0.861 (0.1%) 0.834 (-0.7%) 0.749 (-0.1%) 0.830 (-0.1%) 0.951 (-0.9%)
H2S loading (×10-5) d 4.8 (- f) 4.1 (- f) 5.2 (- f) 4.9 (- f) 4.0 (- f) 4.9 (- f) 6.2 (- f)
3
Flow (m /h) 564.6 (-0.9%) 562.4 (-0.8%) 693.3 (-1.0%) 838.8 (-0.7%) 1026.8 (-0.7%) 1032.6 (-0.7%) 1112.0 (-0.7%)
Temperature (°C) 111.9 (1.7%) 111.7 (1.6%) 118.8 (1.5%) 118.5 (1.3%) 117.5 (0.5%) 118.6 (1.3%) 120.2 (2.7%)
Makeup Water
3 f f f f f f f
Flow (m /h) 1.51 (- ) 1.18 (- ) 1.54 (- ) 1.72 (- ) 1.59 (- ) 2.14 (- ) 3.41 (- )
Sour Water
3 g g
Flow (m /h) 0.00 (- ) 0.00 (- ) 1.28 (-1.7%) 1.46 (-2.6%) 1.45 (-3.3%) 1.78 (-0.9%) 2.48 (-4.5%)
Absorber
f f f f f f f
Pressure drop (bar) 0.03 (- ) 0.02 (- ) 0.02 (- ) 0.03 (- ) 0.02 (- ) 0.04 (- ) 0.04 (- )
Regenerator
Pressure drop (bar) 0.04 (- f) 0.04 (- f) 0.04 (- f) 0.04 (- f) 0.04 (- f) 0.04 (- f) 0.04 (- f)
Solution Reboiler
Steam flow (t/h) 29.4 (1.3%) 26.3 (7.7%) 35.5 (2.0%) 41.5 (-2.0%) 43.7 (-1.2%) 49.8 (1.1%) 61.5 (-5.3%)
a b 3 3 c
This refers to the CO2 content in the dry gas. sm /d refers to m /d at standard conditions (dry gas at 15°C and 1 atm). This is the vapour temperature entering the regenerator wash
d
sections in CO2 trains #1 and #2 and leaving the wash sections in trains #3 to #7. The CO2 and H2S loading refer to the moles of CO2 and H2S per equivalent mole of K2CO3 in the solution.
e
The equivalent moles of K2CO3 are defined as the total number of moles of K2CO3 in the solution if all the KHCO3, KHS and K2S in the solution are converted back into K2CO3. This refers to
f g
the equivalent weight percent of K2CO3 in the solution. No percentage deviation was calculated as the corresponding data value was not given. No percentage deviation was calculated as
the data value was zero.
Chapter 8: HYSYS® CO2 Removal Train Process Model Development
Table 8.3.2: CO2 train simulation results for the second set of plant data in Table 2.1.2. The bracketed values are the percentage deviations from the data.

CO2 Train #1 #2 #3 #4 #5 #6 #7
Sweet Gas
a
mol% CO2 2.95 (5.3%) 2.82 (0.8%) 3.01 (3.7%) 3.00 (3.4%) 3.10 (-0.1%) 2.48 (-0.7%) 1.19 (-0.7%)
f f f f f f f
ppm H2S 2.33 (- ) 2.18 (- ) 2.31 (- ) 2.34 (- ) 2.28 (- ) 1.84 (- ) 0.79 (- )
Temperature (°C) 111.1 (-0.8%) 111.1 (-0.8%) 111.7 (-1.2%) 111.5 (-0.5%) 111.3 (-0.6%) 113.1 (0.1%) 112.8 (-0.2%)
Acid Gas
6 3 b
CO2 (10 sm /d) 0.301 (0.5%) 0.261 (0.4%) 0.466 (-0.9%) 0.531 (0.2%) 0.542 (0.3%) 0.684 (0.6%) 0.831 (0.1%)
c
Temperature (°C) 104.3 (2.2%) 104.9 (1.9%) 75.4 (0.6%) 72.6 (-29.5%) 94.0 (-1.0%) 94.4 (0.4%) 93.6 (-0.4%)
Lean Solution
d
CO2 loading 0.400 (0.0%) 0.400 (0.0%) 0.400 (0.0%) 0.400 (0.0%) 0.440 (0.0%) 0.350 (0.0%) 0.350 (0.0%)
-16 d g g g g g g g
H2S loading (×10 ) 3314.6 (- ) 3366.4 (- ) 117.7 (- ) 47.4 (- ) 343.2 (- ) 6.2 (- ) 2.1E-13 (- )
e
wt% K2CO3 28.0 (0.0%) 28.0 (0.0%) 30.0 (0.0%) 29.0 (0.0%) 30.0 (0.0%) 31.0 (0.0%) 30.0 (0.0%)
Flow (m3/h) 481.0 (0.0%) 475.0 (0.0%) 638.0 (0.0%) 758.0 (0.0%) 922.9 (0.0%) 940.1 (0.1%) 946.1 (0.0%)
Temperature (°C) 111.0 (-0.9%) 111.0 (-0.9%) 111.6 (-0.4%) 111.4 (-0.5%) 111.3 (-0.6%) 113.1 (1.9%) 112.8 (0.7%)
157

Rich Solution
CO2 loading d 0.859 (1.1%) 0.803 (0.4%) 0.891 (0.2%) 0.892 (0.2%) 0.836 (-0.5%) 0.819 (-0.2%) 0.939 (-0.1%)
H2S loading (×10-5) d 3.6 (- f) 3.2 (- f) 3.8 (- f) 3.8 (- f) 3.1 (- f) 3.7 (- f) 4.6 (- f)
3
Flow (m /h) 494.8 (-0.8%) 487.0 (-0.6%) 660.7 (-0.9%) 783.8 (-0.7%) 949.4 (-0.6%) 973.4 (-0.7%) 986.7 (-0.8%)
Temperature (°C) 112.0 (0.0%) 112.0 (0.0%) 118.5 (0.4%) 118.0 (-0.8%) 116.9 (-0.9%) 120.0 (0.0%) 121.2 (1.0%)
Makeup Water
3 f f f f f f f
Flow (m /h) 1.51 (- ) 1.24 (- ) 1.56 (- ) 1.72 (- ) 1.84 (- ) 2.15 (- ) 2.99 (- )
Sour Water
3 g g
Flow (m /h) 0.00 (- ) 0.00 (- ) 1.22 (1.3%) 1.40 (-0.1%) 1.38 (-1.4%) 1.79 (-0.8%) 2.06 (-1.9%)
Absorber
f f f f f f f
Pressure drop (bar) 0.03 (- ) 0.02 (- ) 0.02 (- ) 0.03 (- ) 0.03 (- ) 0.04 (- ) 0.04 (- )
Regenerator
Pressure drop (bar) 0.04 (- f) 0.04 (- f) 0.04 (- f) 0.04 (- f) 0.04 (- f) 0.04 (- f) 0.04 (- f)
Solution Reboiler
Steam flow (t/h) 25.9 (9.6%) 23.4 (8.1%) 33.2 (-5.0%) 38.7 (7.6%) 38.1 (2.7%) 51.3 (3.4%) 57.4 (5.8%)
a b 3 3 c
This refers to the CO2 content in the dry gas. sm /d refers to m /d at standard conditions (dry gas at 15°C and 1 atm). This is the vapour temperature entering the regenerator wash
d
sections in CO2 trains #1 and #2 and leaving the wash sections in trains #3 to #7. The CO2 and H2S loading refer to the moles of CO2 and H2S per equivalent mole of K2CO3 in the solution.
e
The equivalent moles of K2CO3 are defined as the total number of moles of K2CO3 in the solution if all the KHCO3, KHS and K2S in the solution are converted back into K2CO3. This refers to
f g
the equivalent weight percent of K2CO3 in the solution. No percentage deviation was calculated as the corresponding data value was not given. No percentage deviation was calculated as
the data value was zero.
Chapter 8: HYSYS® CO2 Removal Train Process Model Development

8.4 Summary
In summary, this chapter concluded the development of steady-state process models of the Moomba
CO2 trains in the HYSYS® simulation environment. Particular focus was given to the expansion of the
absorber and regenerator column models from Chapter 7 into process models of the absorption and
regeneration circuits that comprise the CO2 trains.

Since the absorber and regenerator columns form the central components of the absorption and
regeneration circuits for the CO2 trains, process models of these circuits were developed by extending
the column process models to include various ancillary operations. Of particular importance were the
solution pumpsets, which consisted of the main solution pumps, steam turbines and power recovery
turbines. Due to the lack of a hydraulic turbine operation in HYSYS®, pump operations were used to
represent the power recovery turbines, and a spreadsheet operation was used to perform the
necessary pumpset calculations for the steady-state simulations. This spreadsheet operation was not
necessary in the dynamic simulation mode due to the ability of the HYSYS® pump operations to act
as hydraulic turbines in dynamic simulations.

The steady-state HYSYS® process models of the CO2 trains were completed by linking the resulting
absorption and regeneration circuit models. This required the addition of a number of adjust logical
operations to facilitate the user-specification of the raw gas and lean solution streams entering the
absorber columns. Tee operations were included to split each lean solution stream so that their flow
rates could be user-specified. This modelling approach necessitated the manual adjustment of the
makeup water streams to ensure negligible flow for the resulting material balance streams.

The performance of the completed steady-state HYSYS® CO2 train models was generally found to
agree well with that of the corresponding Aspen Custom Modeler® CO2 train models and with the
plant data used in the development of the Aspen Custom Modeler® models. Some discrepancies,
primarily concerning the absorption of H2S, were observed due to the limitations of the HYSYS®
models in representing the solution chemistry for the potassium carbonate system. However, these
deviations were deemed to be acceptable since the primary focus of the CO2 trains is to remove CO2.

Good agreement was also observed between the HYSYS® process models and a second
independent set of plant data, thereby validating the HYSYS® models. The HYSYS® CO2 trains
process models were therefore considered to be suitable for the steady-state simulation of the
Moomba CO2 trains, albeit within the range of operating parameters considered during the model
development process.

In the following chapter, the steady-state models for CO2 trains #1 and #7 are converted into dynamic
process models in order to demonstrate a potential application of the HYSYS® CO2 train process
models.

158
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

CHAPTER 9

DYNAMIC HYSYS®
SIMULATIONS OF CO2 REMOVAL
TRAINS #1 AND #7

Detailed steady-state process models of the Moomba CO2 trains were successfully developed in the
HYSYS® simulation environment in the previous chapter. However, these process models do not
provide any information on the dynamic behaviour of the CO2 trains, which limits the potential
applications of the HYSYS® models. As a result, this chapter focuses on the conversion of the
steady-state models for the two most dissimilar CO2 trains, CO2 trains #1 and #7, into dynamic
process models.

Particular attention is given to the various process model modifications required when transitioning
between steady-state and dynamic simulations in HYSYS®. The fundamental differences between
the two simulation modes necessitate several key changes to the steady-state CO2 train models in
order to perform dynamic simulations for the CO2 trains.

The development of the dynamic CO2 train process models is followed by a series of process case
studies, in which the dynamic behaviour of CO2 trains #1 and #7 is examined. Simple FOPDT process
transfer functions are derived to describe the observed behaviour of the two CO2 trains. These will be
used in the next chapter to analyse the multivariable controllability of the two CO2 trains and to
develop suitable diagonal control structures.

159
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

9.1 Dynamic CO2 Train Models


The detailed HYSYS® process models developed in Chapter 8 enable the steady-state simulation of
the Moomba CO2 trains and can be used for a variety of purposes, such as to evaluate the effects of
different operating conditions for process optimisation. While these steady-state HYSYS® models are
particularly useful, they do not provide any insight into the dynamic behaviour of the CO2 trains. As a
result, the steady-state models can not be utilised in potentially useful applications like controllability
analyses or control strategy development.

In order to simulate the dynamic behaviour of the CO2 trains, the steady-state HYSYS® models have
to be converted into dynamic process models. The following sections concentrate on the development
of dynamic process models of the two most dissimilar CO2 trains: CO2 trains #1 and #7. Dynamic
models of the other five CO2 trains were not considered in this work, but are anticipated to exhibit
dynamic behaviour somewhere in between that of these two CO2 trains.

9.1.1 Model Configuration


Process model modifications are required when switching between steady-state and dynamic
simulations in HYSYS®. These necessary changes are due to the inherent differences between the
two simulation modes.

In the steady-state simulations, the unit operations are modular and a non-sequential calculation
algorithm is applied to the material, energy and composition balances (Hyprotech Ltd, 2001b).
Consequently, the results of any calculation are automatically propagated both backwards and
forwards throughout the steady-state process model. This approach enables the properties of internal
streams to be user-specified where necessary, as well as those of the boundary streams.

In contrast, the dynamic simulation mode utilises a sequential algorithm for the energy and
composition balances, while the material (or pressure-flow) balances are calculated simultaneously in
a pressure-flow matrix (Hyprotech Ltd, 2001b). The temperatures and compositions of the boundary
feed streams must therefore be specified so that the temperatures and compositions can be
calculated sequentially for each downstream unit operation and material stream. These calculations
are based on the HYSYS® holdup model, which is described in detail in the HYSYS® Dynamic
Modelling manual (Hyprotech Ltd, 2001b).

The pressures for the boundary material streams must also be specified in the dynamic simulation
mode to enable the calculation of the internal pressures and flows based on the sizes and resistances
of the model unit operations. The pressure gradients across the entire process model and across
each unit operation have to be carefully considered since they act as the driving force for flow in the
dynamic simulation mode.

160
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

The process flow diagrams for the dynamic process models of CO2 trains #1 and #7 are included as
Figures 9.1.1 and 9.1.2. These are based on simplified forms of the steady-state process models in
Figures 8.2.1 and 8.2.2, in which all parallel items were modelled by single units that were sized
accordingly. This model simplification was implemented in order to reduce the computation time for
the dynamic simulations since the calculation speed decreased with increasing model complexity.

To satisfy the above requirements for the dynamic simulation mode, all the pressure and flow
specifications for the internal material streams were removed, while pressure specifications were set
for the boundary material streams. Similarly, temperature and composition specifications were set for
the boundary feed streams, but not for the remaining material streams. Sufficient sizing information
had been provided in the steady-state process models to satisfy the holdup calculations.
Consequently, only the dynamic specifications recommended by the HYSYS® Dynamic Modelling
manual (Hyprotech Ltd, 2001b) were needed for the various unit operations.

For the valve operations, the recommended dynamic specifications entailed replacing the fixed
pressure drops with pressure-flow relations. Similarly, the “overall K value” specification was used for
the heat transfer operations instead of fixed pressure drops. For the mixer operations, the pressure
specification was set to the “Equalise All” option, while the “Remove Use Splits” option was activated
for the tee operations. As in the steady-state process models, characteristic pump curves were used
as the operating specifications for the various pump operations. For the hydraulic turbine models, the
“Pump is acting as turbine” and the “Use Curves“ options were activated, while the corresponding
spreadsheet calculations were deactivated.

Due to the sequential nature of the dynamic calculations, the recycle operations were redundant in
dynamic mode and were therefore removed. The recycle operations could have been retained in the
dynamic models since they had no effect on the simulation results: the recycle operations simply acted
as throughputs without any holdup or pressure gradient. The adjust operations similarly did not
function in dynamic mode. Being the steady-state equivalent of the dynamic HYSYS® control
operations, the adjust operations for the raw gas, lean solution and reboiler steam flows were instead
replaced with PID control operations. These were configured to control the raw gas flow via the raw
gas inlet valve, the lean solution flow via the main solution pump speed, and the reboiler steam flow
via the steam condensate flow valve.

PID control operations were added to control the temperatures of the regenerator overhead
condensers by manipulating the overhead condenser duties. This approach was equivalent to the
actual practice of manually switching the condenser fans on and off as required. These temperature
control loops replaced the temperature specifications on the material streams leaving the overhead
condensers. PID control operations were also included to control the liquid levels in the column
sumps and separator vessels. It should be noted that for CO2 train #7, a split-range controller was
implemented for the regenerator liquid level controller, as specified by the technical drawings in the

161
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

Santos TIMS database. The tuning of the above flow, level and temperature controllers will be
discussed in the next section.

The final modification to the process models was the removal of the material balance streams and the
tee operations for the lean solution streams. The replacement of the adjust operations with PID
controllers made these streams and tee operations redundant: they were no longer required to control
the lean solution flow.

162
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7
163

Figure 9.1.1: Process flow diagram of the simplified dynamic HYSYS® model for CO2 train #1.
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7
164

Figure 9.1.2: Process flow diagram of the simplified dynamic HYSYS® model for CO2 train #7.
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

9.1.2 Flow, Temperature and Level Control Loops


While some controller settings were available from Santos for the flow and liquid level control loops
illustrated in Figures 9.1.1 and 9.1.2, none were available for the temperature control loops since the
overhead condenser temperatures are currently under manual control. For the purposes of this work,
the flow, temperature and level control loops were tuned according to the PID tuning rules presented
in Tables 2.5.2 and 2.5.3.

For the flow control loops, only PI controller settings were used; derivative action was not applied due
to the extremely rapid dynamic nature of the flow control loops. Due to the insignificant dead time
associated with these flow loops, the Ziegler-Nichols frequency domain method was used to
determine their ultimate gains and ultimate periods. These parameters were then used to calculate
the controller settings via the Zeigler-Nichols and Tyreus-Luyben tuning rules in Table 2.5.2.

Table 9.1.1 presents the resulting flow loop characteristics and controller settings, while Figure 9.1.3
compares the loop responses to setpoint and load step changes for the two controller tuning rules. As
expected, more conservative controller settings were produced by the Tyreus-Luyben tuning rules,
thereby resulting in slower responses to the step changes in setpoint and load. Since the Ziegler-
Nichols settings were not observed to be overly aggressive for these flow loops (i.e. no oscillatory loop
responses), they were utilised for this work.

The Ziegler-Nichols frequency domain method was also used to determine the process characteristics
for the temperature control loops, which are given in Table 9.1.2. Since derivative action is often in
temperature control loops (Svrcek et al., 2006), Table 9.1.2 also compares the PI and PID controller
settings derived from the Ziegler-Nichols and Tyreus-Luyben tuning rules in Table 2.5.2. The resulting
loop responses to setpoint and load step changes are provided in Figure 9.1.4.

As for the flow control loops, the Ziegler-Nichols controller settings resulted in more aggressive loop
responses compared to the Tyreus-Luyben settings. It was also observed that the loop responses
corresponding to the PID settings for both sets of tuning rules were faster than those for the PI
settings, but were also more oscillatory for the setpoint step change. Consequently, the Tyreus-
Luyben PI controller settings were selected for this work as they still provided a reasonable response
time, as well as the least underdamped response.

For the liquid level control loops, the equipment sizing information and process specifications were
used to calculate the PID controller settings from the critically damped, minimum IAE and QDR
correlations provided in Table 2.5.3. The resulting settings are given in Table 9.1.3, while Figure 9.1.5
compares the corresponding loop responses to setpoint and load step changes. The QDR controller
settings produced the most oscillatory loop responses, which were far too aggressive in some cases,
whereas the critically damped settings gave the slowest and least underdamped responses. The

165
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

minimum IAE controller settings offered the best compromise between response speed and response
quality, and were therefore employed for this work.

It should be noted that the flow controllers were all reverse-acting, while the temperature and level
controllers were all direct-acting, with the exception of the reverse split-range action for manipulating
the makeup water flow to CO2 train #7.

Table 9.1.1: Flow control loop characteristics and controller settings for CO2 trains #1 and #7.

Flow Loop Raw Gas Lean Solution Reboiler Steam


CO2 Train #1 #7 #1 #7 #1 #7
Loop Characteristics
Ku 0.34 0.37 0.74 0.45 0.35 0.30
Pu (min/cycle) 0.02 0.02 0.03 0.03 0.02 0.02

Controller Settings
Ziegler-Nichols
Kc 0.15 0.17 0.34 0.20 0.16 0.14
WI (min) 0.01 0.01 0.03 0.03 0.01 0.01
Tyreus-Luyben
Kc 0.11 0.12 0.23 0.14 0.11 0.09
WI (min) 0.04 0.04 0.07 0.07 0.04 0.04

Ziegler-Nichols Tyreus-Luyben

2.0 2.0

1.5 1.5
Raw Gas

Raw Gas

1.0 1.0

0.5 0.5

0.0 0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0

2.0 2.0
Lean Solution

Lean Solution

1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0

1.5 1.5
Reboiler Steam

Reboiler Steam

1.0 1.0

0.5 0.5

0.0 0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0

Time (min) Time (min)

(a) (b)
Figure 9.1.3: Flow control loop responses. (a) CO2 train #1. (b) CO2 train #7. A 2% magnitude setpoint
step change occurred at 1 min, followed by a 2% magnitude load step change at 4 min.

166
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

Table 9.1.2: Temperature control loop characteristics and controller settings for CO2 trains #1 and #7.

CO2 Train #1 #7
Loop Characteristics
Ku 5.00 4.00
Pu (min/cycle) 0.17 0.17

Controller Settings PI PID PI PID


Ziegler-Nichols
Kc 2.27 3.00 1.82 2.40
WI (min) 0.14 0.08 0.14 0.08
WD (min) – 0.02 – 0.02
Tyreus-Luyben
Kc 1.56 2.30 1.25 1.84
WI (min) 0.37 0.37 0.37 0.37
WD (min) – 0.03 – 0.03

Ziegler-Nichols PI Tyreus-Luyben PI
Ziegler-Nichols PID Tyreus-Luyben PID

2.5 2.5

2.0 2.0
Overhead Condenser
Overhead Condenser

1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0
Time (min) Time (min)

(a) (b)
Figure 9.1.4: Temperature control loop responses. (a) CO2 train #1. (b) CO2 train #7. A 2% magnitude
setpoint step change occurred at 1 min, followed by a 2% magnitude load step change at 4 min.

Table 9.1.3: Liquid level controller settings for CO2 trains #1 and #7.
Sweet Gas Overhead
Liquid Level Loop Absorber Steam Receiver Regenerator
Separator Catchpot
CO2 Train #1 #7 #1 #7 #1 #7 #1 #7 #1 #7
Critically damped
Kc 2.96 2.96 2.96 2.96 2.96 2.96 2.96 2.96 2.96 2.96
WI (min) 4.01 6.32 34.20 68.70 9.38 4.78 2.25 2.85 19.47 46.74
Minimum IAE
Kc 1.28 1.28 1.28 1.28 1.28 1.28 1.28 1.28 1.28 1.28
WI (min) 0.44 0.69 3.76 7.55 1.03 0.53 0.25 0.31 2.14 5.13
QDR
Kc 2.00 2.00 2.00 2.00 2.00 2.00 2.00 2.00 2.00 2.00
WI (min) 1.10 1.73 9.36 18.81 2.57 1.31 0.62 0.78 5.33 12.80

167
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

Critically Damped Minimum IAE QDR

2.0 2.0

1.5 1.5
Absorber

Absorber
1.0 1.0

0.5 0.5

0.0 0.0
0 5 10 15 20 25 30 35 40 0 10 20 30 40 50 60 70

2.0 1.5
Sweet Gas Separator

Sweet Gas Separator


1.5
1.0

1.0

0.5
0.5

0.0 0.0
0 100 200 300 400 500 600 0 100 200 300 400 500 600

2.0 2.0
Overhead CatchPot

Overhead CatchPot
1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0
0 50 100 150 200 0 50 100 150 200 250

2.0 2.0
Steam Receiver

Steam Receiver

1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0
0 5 10 15 20 25 30 35 40 0 50 100 150 200 250

2.0 2.0

1.5 1.5
Regenerator

Regenerator

1.0 1.0

0.5 0.5

0.0 0.0
0 500 1000 1500 2000 0 500 1000 1500 2000 2500 3000 3500 4000

Time (min) Time (min)

(a) (b)
Figure 9.1.5: Liquid level control loop responses. (a) CO2 train #1. (b) CO2 train #7. 2% magnitude
setpoint and load step changes were implemented at the following times: For the absorber level loop, the
setpoint was changed at 5 min, followed by the load at 20 min for CO2 train #1 and at 30 min for CO2 train
#7. For the sweet gas separator level loop, the setpoint was changed at 10 mins, followed by the load at
400 min for CO2 train #1 and at 340 min for CO2 train #7. For the overhead catchpot level loop, the
setpoint was changed at 10 min, followed by the load at 110 min for CO2 train #1 and at 150 min for CO2
train #7. For the steam condensate receiver level loop, the setpoint was changed at 5 min for CO2 train
#1, followed by the load at 25 min; for CO2 train #7, the setpoint was changed at 10 min, followed by the
load at 130 min. Finally, for the regenerator level loop, the setpoint was changed at 100 min, followed by
the load at 1100 min for CO2 train #1 and at 2100 min for CO2 train #7.

168
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

9.2 Process Case Studies


After implementing the controller parameters derived in the previous section, the dynamic HYSYS®
process models of CO2 trains #1 and #7 were used to investigate the transient process behaviour of
the two CO2 trains. It should be noted that these dynamic simulations were strictly theoretical case
studies as the CO2 train dynamic process models were not validated against plant data due to
difficulties with obtaining sufficient online plant data. Online measurements were not available for the
solution composition 1 and the online CO2 analysers for the raw and sweet gas streams were not
always reliable (Barber, 2007).

Based on an inspection of the CO2 train plant data from 2002 to 2005, three different operating regions
were considered in the dynamic simulations: high, medium and low gas throughputs. These regions
corresponded respectively to 100, 80 and 60% of the raw gas flow rates from the first set of plant data
in Table 2.1.2, and 100, 90 and 70% of the equivalent raw gas to lean solution flow ratio data. All
other process conditions were taken to be the same as for the set of plant data; the exception was the
reboiler steam flow rates, which were adjusted to give the desired lean solution CO2 loading.

The control of the CO2 trains is primarily focussed on controlling three process variables: the sweet
gas CO2 content, the raw gas throughput and the sweet gas pressure to the LRP Plant (Santos Ltd,
1998). These were consequently identified as the three primary process outputs (controlled variables)
for each CO2 train. However, since the latter two variables are directly related to the flow rate of the
sweet gas, the sweet gas CO2 content and flow rate were instead taken to be the two key CO2 train
process outputs for this work.

An examination of the CO2 trains identified four main process inputs (manipulated variables) for each
CO2 train: the raw gas flow rate, the lean solution flow rate, the lean solution CO2 loading and the lean
solution strength. In the parametric studies in Chapter 5, the lean solution CO2 loading was shown to
be dependent on the reboiler steam flow rate, while the lean solution strength was dependent on the
flow rates of makeup water and water bypass, which are both used to control the regenerator liquid
level. Consequently, the key process inputs for the CO2 trains were taken to be the raw gas flow rate,
the lean solution flow rate, the reboiler steam flow rate and the regenerator liquid level. The
examination of the CO2 trains also identified the raw gas CO2 content as the primary process
disturbance.

1
A potential improvement to the Moomba CO2 trains is the implementation of online monitoring of the solution
composition. While direct online measurement is not feasible, the solution composition could be monitored
indirectly using a more easily measured property such as pH or conductivity. For example, pH was used to
monitor the CO2 loading in the amine solution for the Imperial College CO2-MEA pilot plant (Albrecht et al., 1980),
whereas van der Lee and co-workers (2003, 2004) used conductivity in their CO2-amine pilot plant. However, for
this approach to succeed, rigorous calibration studies are required to accurately correlate the solution
concentration with the alternate measurable property.

169
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

In order to determine the effect of the above process inputs and disturbance on the process outputs
for CO2 trains #1 and #7, process step response curves for the two trains were obtained for each
operating region. Based on the observed process variable fluctuations in the historical plant data,
±2% step changes were applied to the process inputs and disturbance, and the resulting averaged
step response curves for the process outputs are shown in Figures 9.2.1 to 9.2.5. The vertical axes
for these figures are dimensionless and are equivalent to those for the response curves depicted in
Table 2.5.1. To achieve this, the process outputs were expressed in perturbation form and were
scaled by dividing each output variable by its allowed range and then by the similarly scaled step
change.

Most of the step response curves depicted in Figures 9.2.1 to 9.2.5 were relatively non-oscillatory and
were readily described by the simple FOPDT model. The corresponding model parameters were
obtained by fitting the FOPDT model step response function from Table 2.5.1 to these response
curves via a least squares method. However, distinctly underdamped response curves were observed
for the step changes in the regenerator liquid level. This was most likely the result of interactions
arising from the various control loops in the process model, such as the regenerator liquid level control
loop. For these underdamped responses, USOPDT model parameters were fitted graphically using
the method shown in Table 2.5.1. These parameters were then used to determine the approximate
FOPDT model parameters from equation (2.5.1), as suggested by Panda and co-workers (2004),
since the quality of the validating data for the CO2 Removal Trains did not justify the use of more
complex models than the FOPDT model. The resulting set of FOPDT process transfer functions for
CO2 trains #1 and #7 are given in Tables 9.2.1 and 9.2.2.

Despite their dissimilar configurations, the two CO2 trains were generally seen to follow the same
behavioural trends over the three different operating regions, as indicated by the similarity between
their step response curves. The observed steady-state relationships between the process inputs and
outputs were also found to be consistent with the results of the parametric studies performed in
Chapter 5. While this was insufficient for validation purposes, it did indicate that the dynamic process
models were producing reasonable results.

It should be noted that one process output was not previously considered in Chapter 5: the sweet gas
flow rate. In the dynamic case studies, it was observed to follow the same trends as the sweet gas
CO2 content for the changes in the process inputs. This was not unexpected since the flow rate of the
sweet gas is directly related to its CO2 content for a constant CO2 content in the raw gas. In contrast,
the sweet gas flow rate was observed to be inversely related to the sweet gas CO2 content for
changes in the raw gas CO2 content, as it compensated for the changing volume of CO2 removed.

170
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

High Gas Throughput Medium Gas Throughput Low Gas Throughput

1.4 6.0

Sweet Gas CO2 Content 1.2

Sweet Gas CO2 Content


5.0

1.0
4.0
0.8
3.0
0.6
2.0
0.4

0.2 1.0

0.0 0.0
0 50 100 150 200 0 50 100 150 200

Time (min) Time (min)


1.2 1.4

1.0 1.2
Sweet Gas Flow Rate

Sweet Gas Flow Rate


1.0
0.8
0.8
0.6
0.6
0.4
0.4

0.2 0.2

0.0 0.0
0 50 100 150 200 0 50 100 150 200

Time (min) Time (min)

(a) (b)
Figure 9.2.1: Process response curves for a 2% magnitude step change in the raw gas flow rate at 0 min.
(a) CO2 train #1. (b) CO2 train #7. The points represent the process response, while the lines represent
the FOPDT model predictions.

High Gas Throughput Medium Gas Throughput Low Gas Throughput

0.00 0.00

-0.01 -0.05
Sweet Gas CO2 Content

Sweet Gas CO2 Content

-0.02 -0.10

-0.03 -0.15

-0.04 -0.20

-0.05 -0.25

-0.06 -0.30

-0.07 -0.35

-0.08 -0.40
0 100 200 300 400 500 0 50 100 150 200 250 300 350 400

Time (min) Time (min)


0.0000 0.000

-0.0005 -0.002
Sweet Gas Flow Rate

Sweet Gas Flow Rate

-0.004
-0.0010
-0.006
-0.0015
-0.008
-0.0020
-0.010
-0.0025
-0.012

-0.0030 -0.014

-0.0035 -0.016
0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400

Time (min) Time (min)

(a) (b)
Figure 9.2.2: Process response curves for a 2% magnitude step change in the lean solution flow rate at 0
min. (a) CO2 train #1. (b) CO2 train #7. The points represent the process response, while the lines
represent the FOPDT model predictions.

171
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

High Gas Throughput Medium Gas Throughput Low Gas Throughput

0.0 0.0

Sweet Gas CO2 Content

Sweet Gas CO2 Content


-0.2
-0.5

-0.4
-1.0
-0.6
-1.5
-0.8

-2.0
-1.0

-1.2 -2.5
0 50 100 150 200 250 300 350 400 0 50 100 150 200

Time (min) Time (min)


0.000 0.00

-0.005 -0.02
Sweet Gas Flow Rate

Sweet Gas Flow Rate


-0.010 -0.04

-0.015 -0.06

-0.020 -0.08

-0.025 -0.10

-0.030 -0.12
0 100 200 300 400 500 0 50 100 150 200

Time (min) Time (min)

(a) (b)
Figure 9.2.3: Process response curves for a 2% magnitude step change in the reboiler steam flow rate at 0
min. (a) CO2 train #1. (b) CO2 train #7. The points represent the process response, while the lines
represent the FOPDT model predictions.

High Gas Throughput Medium Gas Throughput Low Gas Throughput

0.10 1.4

1.2
Sweet Gas CO2 Content

Sweet Gas CO2 Content

0.08
1.0

0.06
0.8

0.6
0.04

0.4
0.02
0.2

0.00 0.0
0 100 200 300 400 500 0 100 200 300 400 500

Time (min) Time (min)


0.007 0.06

0.006 0.05
Sweet Gas Flow Rate

Sweet Gas Flow Rate

0.005
0.04
0.004
0.03
0.003
0.02
0.002

0.001 0.01

0.000 0.00
0 100 200 300 400 500 0 100 200 300 400 500

Time (min) Time (min)

(a) (b)
Figure 9.2.4: Process response curves for a 2% magnitude step change in the regenerator liquid level at 0
min. (a) CO2 train #1. (b) CO2 train #7. The points represent the process response, while the lines
represent the FOPDT model predictions.

172
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

High Gas Throughput Medium Gas Throughput Low Gas Throughput

0.7 2.0

0.6

Sweet Gas CO2 Content


Sweet Gas CO2 Content
1.6
0.5

1.2
0.4

0.3 0.8

0.2
0.4
0.1

0.0 0.0
0 50 100 150 200 0 50 100 150 200

Time (min) Time (min)


0.00 0.00
Sweet Gas Flow Rate

Sweet Gas Flow Rate


-0.02 -0.01

-0.04 -0.02

-0.06 -0.03

-0.08 -0.04

-0.10 -0.05
0 50 100 150 200 0 50 100 150 200

Time (min) Time (min)

(a) (b)
Figure 9.2.5: Process response curves for a 2% magnitude step change in the raw gas CO2 content at 0
min. (a) CO2 train #1. (b) CO2 train #7. The points represent the process response, while the lines
represent the FOPDT model predictions.

Table 9.2.1: Process transfer functions for CO2 train #1. The time constants and dead times are given in
minutes.

Lean Solution Reboiler Steam Regenerator Raw Gas CO2


Raw Gas Flow
Flow Flow Liquid Level Content
High Gas Throughput
Sweet Gas 1.17 ˜ e0.17˜s  0.0635 ˜ e 0.17˜s  0.989 ˜ e 0.67˜s 0.0412 ˜ e 51.94˜s 0.576 ˜ e 0.17˜s
CO2 Content 8.22 ˜ s  1 2.28 ˜ s  1 13.17 ˜ s  1 10.24 ˜ s  1 6.12 ˜ s  1
Sweet Gas 1.09 ˜ e0.17˜s  0.00263 ˜ e0.17˜s  0.0262 ˜ e 0.83˜s 0.00323 ˜ e 69.15˜s  0.0735 ˜ e 0.17˜s
Flow 1.62 ˜ s  1 3.37 ˜ s  1 16.02 ˜ s  1 9.50 ˜ s  1 3.44 ˜ s  1

Medium Gas Throughput


Sweet Gas 0.945 ˜ e 0.17˜s  0.0204 ˜ e 0.17˜s  0.781˜ e 0.67˜s 0.0368 ˜ e 48.87˜s 0.523 ˜ e 0.17˜s
CO2 Content 14.90 ˜ s  1 2.53 ˜ s  1 17.31˜ s  1 11.33 ˜ s  1 9.94 ˜ s  1
Sweet Gas 0.879 ˜ e 0.17˜s  0.000617 ˜ e0.17˜s  0.0207 ˜ e 0.83˜s 0.00288 ˜ e 77.51˜s  0.08108 ˜ e 0.17˜s
Flow 2.34 ˜ s  1 2.90 ˜ s  1 18.73 ˜ s  1 11.10 ˜ s  1 8.12 ˜ s  1

Low Gas Throughput


Sweet Gas 0.721 ˜ e 0.17˜s  0.00603 ˜ e 0.17˜s  0.659 ˜ e 0.67˜s 0.0307 ˜ e 46.01˜s 0.503 ˜ e 0.17˜s
CO2 Content 23.39 ˜ s  1 3.02 ˜ s  1 21.11˜ s  1 13.28 ˜ s  1 14.30 ˜ s  1
Sweet Gas 0.671˜ e0.17˜s  0.000398 ˜ e0.17˜s  0.0175 ˜ e 0.83˜s 0.00240 ˜ e 81.88˜s  0.0842 ˜ e 0.17˜s
Flow 2.80 ˜ s  1 2.22 ˜ s  1 20.85 ˜ s  1 8.65 ˜ s  1 13.31˜ s  1

173
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

Table 9.2.2: Process transfer functions for CO2 train #7. The time constants and dead times are given in
minutes.

Lean Solution Reboiler Steam Regenerator Raw Gas CO2


Raw Gas Flow
Flow Flow Liquid Level Content
High Gas Throughput
Sweet Gas 5.29 ˜ e 0.17˜s  0.284 ˜ e 0.17˜s  2.35 ˜ e 0.67 ˜s 0.577 ˜ e 41.10 ˜s 1.77 ˜ e 0.17˜s
CO2 Content 4.30 ˜ s  1 1.80 ˜ s  1 9.98 ˜ s  1 7.58 ˜ s  1 3.52 ˜ s  1
Sweet Gas 1.32 ˜ e 0.17˜s  0.0133 ˜ e 0.17˜s  0.0981 ˜ e 0.33 ˜s 0.0258 ˜ e 41.27˜s  0.0203 ˜ e 0.50˜s
Flow 2.10 ˜ s  1 1.84 ˜ s  1 10.52 ˜ s  1 7.58 ˜ s  1 11.66 ˜ s  1

Medium Gas Throughput


Sweet Gas 2.70 ˜ e 0.17˜s  0.0456 ˜ e 0.17 ˜s  1.61˜ e 0.67˜s 0.351 ˜ e 47.95˜s 0.978 ˜ e 0.17˜s
CO2 Content 9.93 ˜ s  1 2.60 ˜ s  1 14.97 ˜ s  1 6.82 ˜ s  1 6.70 ˜ s  1
Sweet Gas 1.18 ˜ e 0.17 ˜s  0.00237 ˜ e 0.17 ˜s  0.0535 ˜ e 0.33˜s 0.0132 ˜ e 44.99˜s  0.0433 ˜ e 0.50˜s
Flow 2.80 ˜ s  1 1.48 ˜ s  1 14.68 ˜ s  1 7.23 ˜ s  1 6.66 ˜ s  1

Low Gas Throughput


Sweet Gas 1.58 ˜ e 0.17˜s  0.0281 ˜ e 0.17˜s  1.20 ˜ e 0.67˜s 0.226 ˜ e 56.10˜s 0.588 ˜ e 0.17˜s
CO2 Content 20.66 ˜ s  1 2.99 ˜ s  1 19.23 ˜ s  1 6.96 ˜ s  1 10.54 ˜ s  1
Sweet Gas 1.12 ˜ e 0.17˜s  0.00195 ˜ e0.17˜s  0.0292 ˜ e 0.33˜s 0.00687 ˜ e 61.08˜s  0.0410 ˜ e 0.5˜s
Flow 2.91 ˜ s  1 1.38 ˜ s  1 17.88 ˜ s  1 6.63 ˜ s  1 4.71˜ s  1

To demonstrate the potential application of the HYSYS® CO2 train process models in the analysis and
design of control strategies for the CO2 trains, the simple FOPDT process transfer functions derived in
this section are used in the next chapter to analyse the multivariable controllability of the two CO2
trains, and to develop suitable diagonal control structures for them. As discussed in Section 2.5, the
application of diagonal control structures to the hot potassium carbonate process has not been
previously documented in literature, despite the diagonal control structure being the simplest and most
widely used form of multivariable control.

For this reason, only simple PID controllers were considered in the development of the diagonal
control structures for CO2 trains #1 and #7 for this work. However, both CO2 trains demonstrated
distinctly non-linear dynamic behaviour, as illustrated by the variation in the process transfer functions
and step response curves over the different operating regions. This observed process non-linearity
suggested that more complex adaptive controller algorithms, such as gain scheduling, may provide
more effective control of the CO2 trains than simple PID controllers. It is therefore recommended that
the application of more complex controller algorithms to the CO2 trains be investigated in any future
work.

174
Chapter 9: Dynamic HYSYS® Simulations of CO2 Removal Trains #1 and #7

9.3 Summary
In summary, this chapter presented simplified dynamic HYSYS® process models of CO2 trains #1 and
#7 and investigated the dynamic relationships between the key process outputs and inputs for the two
CO2 trains.

The dynamic HYSYS® models were developed from the detailed steady-state models presented in
Chapter 4. Due to the inherent differences between the dynamic and steady-state simulation modes
in HYSYS®, a number of process model modifications were required to convert the steady-state
models into dynamic ones. These changes included the amendment of various stream and unit
operation specifications, the replacement of the adjust operations with PID control operations, and the
addition of level control loops for all vessels.

The inclusion of the PID control operations necessitated the specification of suitable controller
parameters in order to complete the dynamic process models. Only some of the required parameters
were available from Santos due to various control loops being under manual control. Consequently,
for the purposes of this work, the model control loops were tuned according to the methods discussed
in Section 2.5.

Using the completed dynamic process models, the dynamic behaviour of CO2 trains #1 and #7 was
examined in a series of theoretical process case studies. Three different operating regions were
considered in these studies: high, medium and low gas throughputs. The resulting process response
curves were used to derive simple FOPDT process transfer functions which adequately described the
relationships between the key process outputs (the sweet gas CO2 content and flow rate) and the
primary process inputs (the raw gas flow rate, the lean solution flow rate, the lean solution CO2 loading
and the lean solution strength) and disturbance (the raw gas CO2 content). The two CO2 trains were
observed to behave quite similarly, despite the differences in their configurations.

The FOPDT process transfer functions are used in the next chapter to demonstrate the potential
application of the HYSYS® CO2 train process models in the analysis and development of suitable
diagonal control structures for the CO2 trains. As determined in the literature review in Chapter 2, the
application of this simple form of multivariable control has not been previously documented for hot
potassium carbonate process operations, thereby constituting a gap in the current literature.

175
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

CHAPTER 10

PROCESS CONTROL STUDIES


FOR CO2 TRAINS #1 AND #7

Simplified dynamic HYSYS® process models of Moomba CO2 trains #1 and #7 were developed in the
previous chapter, and were used to derive FODPT process transfer functions to describe the dynamic
behaviour of the two CO2 trains. In this chapter, these process transfer functions are used to analyse
the multivariable controllability of the two CO2 trains in order to develop potentially suitable diagonal
control structures for the overall automatic control of each CO2 train.

The analysis methods discussed in Chapter 2 are applied to the process transfer functions to evaluate
the possible diagonal control structures and structure configurations arising from the different
combinations and permutations of the key process inputs and outputs. The effect of the primary
process disturbance on the performance of the potential diagonal control structures is also examined.

Finally, controller parameters for the optimal diagonal control structures are determined using the BLT
tuning approach, and the selected diagonal control structures are tested via dynamic MATLAB®
simulations to determine their applicability to the two CO2 trains.

The simple case studies discussed in this chapter demonstrate a potentially useful application of the
HYSYS® process models of the Moomba CO2 trains. Further work is required in order to develop a
viable control strategy for the CO2 trains.

176
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

10.1 Selection of Diagonal Control Structure


In the previous chapter, four manipulated variables (process inputs) were identified for the CO2 trains,
in contrast to two controlled variables (process outputs). Consequently, six alternative diagonal
control structures were proposed for each train, one for each of the six possible combinations of
manipulated variables. However, three of these were immediately eliminated as they excluded the
raw gas flow rate, which is required for the effective control of the sweet gas flow rate.

The remaining sets of manipulated variable were denoted as RGF-LSF, RGF-RSF and RGF-RLL, with
RGF referring to the raw gas flow rate, LSF referring to the lean solution flow rate, RSF referring to the
reboiler steam flow rate, and RLL referring to the regenerator liquid level. The process transfer
function matrices for the resulting three potential control structures were derived from the simple
process transfer functions in Table 9.2.1 and 9.2.2, and are given in Table 10.1.1. These matrices
corresponded to a 2×2 system of the form:
ª y1 º ªG p,11 (s) G p,12 (s)º ª m1 º
« » « »˜« » (10.1.1)
¬y 2 ¼ ¬G p,21 (s) G p,22 (s)¼ ¬m 2 ¼

where y1 is the sweet gas CO2 content, y2 is the sweet gas flow rate, m1 is the raw gas flow rate, and
Gp,11(s) to Gp,21(s) are the process transfer functions relating the controlled variables y to the
manipulated variables m. The manipulated variable m2 varies depending on the control structure. For
the RGF-LSF structure, m2 is the lean solution flow rate; for the RGF-RSF structure, it refers to the
reboiler steam flow rate; and it is the regenerator liquid level for the RGF-RLL structure.

To identify the optimal set of manipulated variables, and therefore the optimal diagonal control
structure for CO2 trains #1 and #7, the MRI, CN, DCN and DC sensitivity indices were used to
investigate the sensitivity of the alternative diagonal control structures to process uncertainties and
disturbances. The necessary calculations were performed in the MATLAB® 5.2 numerical computing
environment, using equations (2.5.4) to (2.5.6). The dead times could not be directly implemented in
the MATLAB® calculations, and were instead specified as rational transfer functions using the third-
order Padé approximation. An example MATLAB® script for determining the sensitivity indices is
provided in Section I.1 of Appendix I.

The resulting values of the four sensitivity indices at steady-state (s=0 or Z=0 rad/min) are
summarised in Table 10.1.2, while the corresponding frequency plots are given in Figures 10.1.1 to
10.1.3. Of particular interest was the frequency range between 0.01 and 1 rad/min, due to the
importance of this frequency range to feedback control (Skogestad et al., 1990).

From Table 10.1.2 and Figure 10.1.1, it was observed that the MRI values for the three different
diagonal control structures were quite small (<1) over the examined frequency range. This indicated
that the three structures were relatively sensitive to process uncertainties. The RGF-RSF structure
was found to have the largest MRI values, making it the preferred option of the three. This preference

177
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

was confirmed by the CN values provided in Table 10.1.2 and Figure 10.1.1. The RGF-LSF and RGF-
RLL structures were quite ill-conditioned, having CN values that were generally substantially larger
than 10. In contrast, the RGF-RSF structure was reasonably well-conditioned, since most of its CN
values were below 10, with the exception of the CN values associated with CO2 train #7 at high gas
throughput. However, these were still considerably less than the corresponding CN values for the
other two control structures.

Unit magnitude changes in the raw gas CO2 content and in the setpoints for the sweet gas CO2
content and flow rate were considered for the DCN and DC calculations. From Table 10.1.2 and
Figures 10.1.2 and 10.1.3, it was evident that the DCN values for the RGF-LSF and RGF-RLL control
structures were considerably larger than the corresponding values for the RGF-RSF structure over the
examined frequency range. The same trend was observed for the DC values. This indicated that, of
the three potential control structures, the RGF-RSF structure was the least sensitive to process
disturbances and setpoint changes.

Consequently, from the above sensitivity analyses, the RGF-RSF structure was identified as the
optimal diagonal control structure for CO2 trains #1 and #7. It was therefore selected for the
interaction and stability analyses, which are discussed in the following section.

178
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

Table 10.1.1: Process transfer function matrices for the diagonal control structures. The time constants
and dead times are given in minutes.

RGF-LSF RGF-RSF RGF-RLL


CO2 Train #1
ª 1.17 ˜ e 0.17 ˜s  0.0635 ˜ e 0.17˜s º ª 1.17 ˜ e 0.17˜s  0.989 ˜ e 0.67˜s º ª 1.17 ˜ e 0.17˜s 0.041 ˜ e 51.94˜s º
« » « » « »
High Gas « 8.22 ˜ s  1 2.28 ˜ s  1 » « 8.22 ˜ s  1 13.17 ˜ s  1 » « 8.22 ˜ s  1 10.24 ˜ s  1 »
Throughput «  0.17 ˜ s  0.17 ˜s
» «  0.17˜s » « »
« 1.09 ˜ e  0.00263 ˜ e » «1.09 ˜ e  0.0262 ˜ e  0.83˜s » «1.09 ˜ e
 0.17 ˜s
0.00323 ˜ e  69.15˜s »
¬« 1.62 ˜ s  1 3.37 ˜ s  1 ¼» ¬« 1.62 ˜ s  1 16.02 ˜ s  1 ¼» ¬« 1.62 ˜ s  1 9.50 ˜ s  1 ¼»
ª 0.945 ˜ e 0.17˜s  0.0204 ˜ e 0.17˜s º ª 0.945 ˜ e 0.17 ˜s  0.781 ˜ e 0.67˜s º ª 0.945 ˜ e 0.17 ˜s 0.0368 ˜ e 48.87˜s º
« » « » « »
Medium Gas « 14.90 ˜ s  1 2.53 ˜ s  1 » « 14.90 ˜ s  1 17.31 ˜ s  1 » « 14.90 ˜ s  1 11.33 ˜ s  1 »
Throughput «  0.17 ˜s
» « » « »
« 0.879 ˜ e  0.000617 ˜ e  0.17˜s » « 0.879 ˜ e
 0.17 ˜s
 0.0207 ˜ e  0.83˜s » « 0.879 ˜ e  0.17˜s 0.00288 ˜ e  77.51˜s »
«¬ 2.34 ˜ s  1 2.90 ˜ s  1 »¼ «¬ 2.34 ˜ s  1 18.73 ˜ s  1 »¼ «¬ 2.34 ˜ s  1 11.10 ˜ s  1 »¼
ª 0.721 ˜ e 0.17˜s  0.00603 ˜ e 0.17˜s º ª 0.721 ˜ e 0.17˜s  0.659 ˜ e 0.67˜s º ª 0.721 ˜ e 0.17˜s 0.0307 ˜ e 46.01˜s º
« » « » « »
Low Gas « 23.39 ˜ s  1 3.02 ˜ s  1 » « 23.39 ˜ s  1 21.11 ˜ s  1 » « 23.39 ˜ s  1 13.28 ˜ s  1 »
Throughput «  0.17 ˜s
» « » « »
« 0.671 ˜ e  0.000398 ˜ e  0.17˜s » « 0.671 ˜ e
 0.17 ˜s
 0.0175 ˜ e  0.83˜s » « 0.671 ˜ e
 0.17 ˜s
0.00240 ˜ e  81.88˜s »
«¬ 2.80 ˜ s  1 2.22 ˜ s  1 »¼ «¬ 2.80 ˜ s  1 20.85 ˜ s  1 »¼ «¬ 2.80 ˜ s  1 8.65 ˜ s  1 »¼

CO2 Train #7
ª 5.29 ˜ e 0.17˜s  0.284 ˜ e 0.17˜s º ª 5.29 ˜ e 0.17˜s  2.35 ˜ e 0.67˜s º ª 5.29 ˜ e 0.17˜s 0.577 ˜ e 41.10˜s º
« » « » « »
High Gas « 4.30 ˜ s  1 1.80 ˜ s  1 » « 4.30 ˜ s  1 9.98 ˜ s  1 » « 4.30 ˜ s  1 7.58 ˜ s  1 »
Throughput «  0.17˜s » « » « »
« 1.32 ˜ e  0.0133 ˜ e  0.17˜s » « 1.32 ˜ e
 0.17˜ s
 0.0981 ˜ e  0.33˜s » « 1.32 ˜ e
 0.17 ˜s
0.0258 ˜ e  41.27˜s »
¬« 2.10 ˜ s  1 1.84 ˜ s  1 ¼» ¬« 2.10 ˜ s  1 10.52 ˜ s  1 ¼» ¬« 2.10 ˜ s  1 7.58 ˜ s  1 ¼»
ª 2.70 ˜ e 0.17 ˜s  0.0456 ˜ e 0.17˜s º ª 2.70 ˜ e 0.17˜s  1.61 ˜ e 0.67˜s º ª 2.70 ˜ e 0.17 ˜s 0.351 ˜ e 47.95˜s º
« » « » « »
Medium Gas « 9.93 ˜ s  1 2.60 ˜ s  1 » « 9.93 ˜ s  1 14.97 ˜ s  1 » « 9.93 ˜ s  1 6.82 ˜ s  1 »
Throughput «  0.17 ˜s
» « » « »
« 1.18 ˜ e  0.00237 ˜ e  0.17˜s » « 1.18 ˜ e
 0.17˜s
 0.0535 ˜ e  0.33˜s » « 1.18 ˜ e
 0.17˜s
0.0132 ˜ e  44.99˜s »
«¬ 2.80 ˜ s  1 1.48 ˜ s  1 »¼ «¬ 2.80 ˜ s  1 14.68 ˜ s  1 »¼ «¬ 2.80 ˜ s  1 7.23 ˜ s  1 »¼
ª 1.58 ˜ e 0.17˜s  0.0281 ˜ e 0.17˜s º ª 1.58 ˜ e 0.17˜s  1.20 ˜ e 0.67˜s º ª 1.58 ˜ e 0.17˜s 0.266 ˜ e 56.10˜s º
« » « » « »
Low Gas « 20.66 ˜ s  1 2.99 ˜ s  1 » « 20.66 ˜ s  1 19.23 ˜ s  1 » « 20.66 ˜ s  1 6.96 ˜ s  1 »
Throughput «  0.17 ˜s  0.17 ˜s
» «  0.17˜s » «  61.08˜s »
« 1.12 ˜ e  0.00195 ˜ e » «1.12 ˜ e  0.0292 ˜ e  0.33˜s » «1.12 ˜ e
 0.17˜s
0.00687 ˜ e »
¬« 2.91 ˜ s  1 1.38 ˜ s  1 ¼» ¬« 2.91 ˜ s  1 17.88 ˜ s  1 ¼» ¬« 2.91 ˜ s  1 6.63 ˜ s  1 ¼»

179
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

Table 10.1.2: Sensitivity analysis indices at steady-state.

CO2 Train #1 CO2 Train #7


RGF-LSF RGF-RSF RGF-RLL RGF-LSF RGF-RSF RGF-RLL
High Gas Throughput
MRI 0.041 0.585 0.026 0.056 0.437 0.115
CN 38.67 3.05 62.13 97.79 13.56 47.92
DCN - d 29.70 2.11 47.73 24.80 3.18 12.11
DCN - y1 26.33 1.86 42.31 23.72 3.04 11.58
DCN - y2 28.34 2.62 45.51 94.88 13.25 46.51
DC - d 10.79 0.69 17.34 8.05 0.95 3.92
DC - y1 16.47 1.04 26.48 4.34 0.51 2.11
DC - y2 17.73 1.47 28.48 17.37 2.24 8.48

Medium Gas Throughput


MRI 0.014 0.465 0.023 0.016 0.529 0.127
CN 95.88 3.09 56.26 184.19 6.28 23.32
DCN - d 75.29 2.20 44.17 80.95 2.46 10.19
DCN - y1 65.31 1.89 38.31 73.56 2.23 9.26
DCN - y2 70.21 2.64 41.21 168.87 5.96 21.43
DC - d 30.86 0.81 18.10 26.88 0.73 3.36
DC - y1 50.62 1.32 29.69 24.94 0.67 3.12
DC - y2 54.42 1.84 31.93 57.26 1.79 7.22

Low Gas Throughput


MRI 0.004 0.383 0.019 0.015 0.592 0.125
CN 246.40 2.93 51.46 131.16 3.72 15.56
DCN - d 195.34 2.067 40.79 83.27 2.08 9.84
DCN - y1 167.85 1.75 35.05 76.02 1.89 8.98
DCN - y2 180.38 2.55 37.70 106.89 3.35 12.75
DC - d 101.07 0.94 21.10 25.38 0.56 2.98
DC - y1 170.35 1.56 35.56 39.33 0.86 4.62
DC - y2 183.08 2.27 38.25 55.30 1.52 6.57
Notation: d is the raw gas CO2 content, y1 is the sweet gas CO2 content, and y2 is the sweet gas flow rate.

180
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

RGF-LSF RGF-RSF RGF-RLL

1.0.E+00 1.0.E+00

1.0.E-01 1.0.E-01

1.0.E-02 1.0.E-02

MRI
MRI

1.0.E-03 1.0.E-03

1.0.E-04 1.0.E-04

1.0.E-05 1.0.E-05
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02 1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02
Frequency (rad/min) Frequency (rad/min)

1.0.E+03 1.0.E+03

1.0.E+02 1.0.E+02
CN

CN

1.0.E+01 1.0.E+01

1.0.E+00 1.0.E+00
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02 1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02
Frequency (rad/min) Frequency (rad/min)
(a) (b)
Figure 10.1.1: Frequency plots of the MRI and CN. (a) CO2 train #1. (b) CO2 train #7. The solid lines (––)
represent the high gas throughput conditions, the dashed lines (---) represent the medium gas throughput
conditions, and the crosses (×××) represent the low gas throughput conditions.

181
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

RGF-LSF RGF-RSF RGF-RLL

1.0.E+03 1.0.E+03

1.0.E+02
1.0.E+02
DCN

DC
1.0.E+01

1.0.E+01
1.0.E+00

1.0.E+00 1.0.E-01
1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02 1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02

Frequency (rad/min) Frequency (rad/min)


(a)

1.0.E+03 1.0.E+05

1.0.E+04

1.0.E+02
1.0.E+03
DCN

DC

1.0.E+02
1.0.E+01

1.0.E+01

1.0.E+00 1.0.E+00
1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02 1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02

Frequency (rad/min) Frequency (rad/min)


(b)

1.0.E+03 1.0.E+05

1.0.E+04

1.0.E+02
1.0.E+03
DCN

DC

1.0.E+02
1.0.E+01

1.0.E+01

1.0.E+00 1.0.E+00
1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02 1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02

Frequency (rad/min) Frequency (rad/min)


(c)
Figure 10.1.2: Frequency plots of DCN and DC for CO2 train #1. Unit magnitude change in (a) the raw gas
CO2 content, (b) the sweet gas CO2 content setpoint and (c) the sweet gas flow rate setpoint. The solid
lines (––) represent the high gas throughput conditions, the dashed lines (---) represent the medium gas
throughput conditions, and the crosses (×××) represent the low gas throughput conditions.

182
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

RGF-LSF RGF-RSF RGF-RLL

1.0.E+03 1.0.E+02

1.0.E+02 1.0.E+01
DCN

DC
1.0.E+01 1.0.E+00

1.0.E+00 1.0.E-01
1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02 1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02

Frequency (rad/min) Frequency (rad/min)


(a)

1.0.E+03 1.0.E+05

1.0.E+02 1.0.E+03
DCN

DC

1.0.E+01 1.0.E+01

1.0.E+00 1.0.E-01
1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02 1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02

Frequency (rad/min) Frequency (rad/min)


(b)

1.0.E+03 1.0.E+04

1.0.E+03
1.0.E+02
DCN

DC

1.0.E+02

1.0.E+01
1.0.E+01

1.0.E+00 1.0.E+00
1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02 1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02

Frequency (rad/min) Frequency (rad/min)


(c)
Figure 10.1.3: Frequency plots of DCN and DC for CO2 train #7. Unit magnitude change in (a) the raw gas
CO2 content, (b) the sweet gas CO2 content setpoint and (c) the sweet gas flow rate setpoint. The solid
lines (––) represent the high gas throughput conditions, the dashed lines (---) represent the medium gas
throughput conditions, and the crosses (×××) represent the low gas throughput conditions.

183
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

10.2 Selection of Diagonal Control Structure Configuration


Based on the sensitivity analyses performed in the previous section, the RGF-RSF diagonal control
structure was identified as the optimal control structure for CO2 trains #1 and #7. In this section, the
loop interactions and system stability are evaluated in order to determine the best configuration for this
control structure. As in the previous section, all necessary calculations were performed in the
MATLAB® 5.2 numerical computing environment.

As an initial analysis, the process stability and controllability were checked by determining the system
poles and zeros. An example of the MATLAB® scripts for performing these calculations is provided in
Section I.2.1 in Appendix I. For the purpose of brevity, the calculated poles and zeros for the RGF-
RSF control structure have not been included in the main body of this thesis, but are instead provided
in Table I.2.1 in Appendix I.

No right-hand-plane (RHP) poles were identified, which indicated that the control structure was stable
(Skogestad and Postlethwaite, 1996) for both CO2 trains over the different operating regions. Due to
the third-order Padé approximations for the dead times, several RHP zeros were identified for each
system, which indicated that the systems would have inverse response behaviour. However, these
RHP zeros were not located near the origin, which would have otherwise implied poor control
performance (Skogestad and Postlethwaite, 1996).

To determine the loop interactions, the RGA was derived as a function of frequency using equation
(2.5.7). An example of the MATLAB® scripts used to perform these calculations is included in Section
I.2.2 in Appendix I. The steady-state values for the diagonal and off-diagonal elements O11 and O12,
respectively, are given in Table 10.2.1, while the corresponding frequency plots are depicted in Figure
10.2.1. It should be noted that for 2×2 systems, O11=O22 and O12=O21 since the elements of each row
and column of the RGA sum to 1 (Skogestad and Postlethwaite, 1996).

It was observed that the diagonal elements of the RGA were small and negative, whereas the off-
diagonal elements were positive and close to 1 over the examined frequency range. This strong off-
diagonal dominance favoured the off-diagonal variable pairings for minimising the loop interactions.
However, some degree of interaction was still expected, especially for CO2 train #7, since the off-
diagonal elements were greater than 1 (Svrcek et al., 2006). The off-diagonal pairings were also
preferred due to their positive values, which ensured the systems corresponding to the off-diagonal
pairings were decentralised integral controllable (Skogestad and Postlethwaite, 1996). This indicated
that a diagonal control structure with integral action in each loop could be implemented such that the
feedback system was stable and each loop could be detuned independently without introducing
instability.

To confirm the suitability of the off-diagonal variable pairings, the process transfer function matrices
were reordered such that the paired variables were along the diagonal, as shown in Table 10.2.2.

184
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

Using these reordered transfer function matrices, the NI and MIC were calculated from equations
(2.5.8) and (2.5.9). An example MATLAB® script for these calculations is included in Section I.2.3 in
Appendix I, while the resulting values are given in Table 10.2.1. For both CO2 trains, the NI and MIC
values were positive. This signified that the RGF-RSF systems described by the process transfer
function matrices in Table 10.2.2 were definitely closed-loop stable.

As a result of the above analyses, the optimal configuration for the RGF-RSF diagonal control
structure for CO2 trains #1 and #7 was identified as the one in which the raw gas and reboiler steam
flow rates were used to control the sweet gas flow rate and CO2 content, respectively, corresponding
to the y1-m2 and y2-m1 variable pairings from equation (10.1.1). This was not unexpected since the
sweet gas flow rate is directly dependent on the raw gas flow rate.

The performance of the optimal diagonal control structure configuration is analysed in the next section.

Table 10.2.1: Steady-state results for the interaction and stability analyses for the RGF-RSF control
structure. The NI and MIC values correspond to the off-diagonal RGA pairings.

O 11, O12 NI MIC


CO2 Train #1
High Gas Throughput -0.03, 1.03 0.97 0.86, 1.22
Medium Gas Throughput -0.03, 1.03 0.97 0.68, 0.98
Low Gas Throughput -0.03, 1.03 0.97 0.55, 0.78

CO2 Train #7
High Gas Throughput -0.20, 1.20 0.83 2.72, 0.95
Medium Gas Throughput -0.08, 1.08 0.92 1.83, 0.96
Low Gas Throughput -0.04, 1.04 0.97 1.38, 0.94

High gas throughput Medium gas throughput Low gas throughput

1.5 1.5

1.0 1.0
O11 and O12

O11 and O12

0.5 0.5

0.0 0.0

-0.5 -0.5
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02 1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02

Frequency (rad/min) Frequency (rad/min)


(a) (b)
Figure 10.2.1: Frequency plots for the RGA elements. (a) CO2 train #1. (b) CO2 train #7. The solid lines
(––) represent the diagonal elements O, while the dashed lines (---) represent the off-diagonal elements
O.

185
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

Table 10.2.2: Reordered process transfer function matrices for the RGF-RSF diagonal control structure.
The time constants and dead times are given in minutes.

CO2 Train #1 CO2 Train #7


ª  0.989 ˜ e 0.67˜s 1.17 ˜ e º
0.17 ˜ s ª  2.35 ˜ e 0.67 ˜s 5.29 ˜ e 0.17 ˜s º
« » « »
High Gas « 13.17 ˜ s  1 8.22 ˜ s  1 » « 9.98 ˜ s  1 4.30 ˜ s  1 »
Throughput « » «  0.33 ˜s
»
«  0.0262 ˜ e
 0.83 ˜ s
1.09 ˜ e  0.17˜s » «  0.0981 ˜ e 1.32 ˜ e  0.17˜s »
«¬ 16.02 ˜ s  1 1.62 ˜ s  1 »¼ «¬ 10.52 ˜ s  1 2.10 ˜ s  1 »¼
ª  0.781 ˜ e 0.67 ˜s 0.945 ˜ e 0.17 ˜s º ª  1.61 ˜ e 0.67 ˜s 2.70 ˜ e 0.17 ˜s º
« » « »
Medium Gas « 17.31 ˜ s  1 14.90 ˜ s  1 » « 14.97 ˜ s  1 9.93 ˜ s  1 »
Throughput «  0.83 ˜ s » « »
«  0.0207 ˜ e 0.879 ˜ e  0.17 ˜s » «  0.0535 ˜ e
 0.33 ˜s
1.18 ˜ e  0.17 ˜s »
«¬ 18.73 ˜ s  1 2.34 ˜ s  1 »¼ «¬ 14.68 ˜ s  1 2.80 ˜ s  1 »¼
ª  0.659 ˜ e 0.67 ˜s 0.721 ˜ e 0.17 ˜s º ª  1.20 ˜ e 0.67˜s 1.58 ˜ e 0.17˜s º
« » « »
Low Gas « 21.11 ˜ s  1 23.39 ˜ s  1 » « 19.23 ˜ s  1 20.66 ˜ s  1 »
Throughput «  0.83 ˜ s » « »
«  0.0175 ˜ e 0.671 ˜ e  0.17 ˜s » «  0.0292 ˜ e
 0.33 ˜ s
1.12 ˜ e  0.17˜s »
«¬ 20.85 ˜ s  1 2.80 ˜ s  1 »¼ «¬ 17.88 ˜ s  1 2.91 ˜ s  1 »¼

186
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

10.3 Analysis of Diagonal Control Structure Performance


Having identified a suitable configuration for the RGF-RSF diagonal control structure for CO2 trains #1
and #7, the performance of the resulting control systems are analysed in this section using the
frequency-dependent PRGA and CLDG described by equations (2.5.10) and (2.5.11). The steady-
state values for the PRGA and CLDG are provided in Table 10.3.1, while the corresponding frequency
plots are given in Figure 10.3.1. An example of the MATLAB® scripts used to perform the necessary
calculations is included in Section I.3 in Appendix I.

It was observed that the PRGA and CLDG elements were larger for CO2 train #7 than for CO2 train
#1. This implied that the former CO2 train was more difficult to control than the latter, which was
consistent with the greater degree of interaction indicated by the larger off-diagonal RGA elements for
CO2 train #7 in Table 10.2.1. It was also noted that the PRGA and CLDG elements decreased with
decreasing gas throughput, which indicated that the trains, more notably CO2 train #7, became more
difficult to control with increasing gas throughput. This corresponded well with the observed trend for
the off-diagonal RGA elements in Figure 10.2.1.

To achieve acceptable diagonal control structure performance for the two trains, condition (2.5.12) for
the open loop transfer functions GOL,i(s) had to be satisfied for each control loop. For the sweet gas
CO2 content loop (denoted as Loop 1), |PRGA11| and |PRGA12| were considered for setpoint tracking,
while |CLDG11| was considered for disturbance rejection. In the case of the sweet gas flow rate loop
(denoted as Loop 2), |PRGA22| and |PRGA21| were considered for setpoint tracking, while |CLDG21|
was considered for disturbance rejection.

For CO2 train # 1, both |PRGA11| and |PRGA12| were larger than |CLDG11| over the examined
frequency range. This implied that changes in the setpoints for the sweet gas CO2 content and flow
rate had a greater effect on the sweet gas CO2 content than any changes in the raw gas CO2 content.
At low frequencies (up to 0.01, 0.03 and 0.04 rad/min respectively for the low, medium and high gas
throughput conditions), |PRGA12| was marginally larger than |PRGA11|, which remained relatively
constant over the frequency range. Consequently, to ensure satisfactory performance for Loop 1 for
CO2 train #1 over the three operating regions, |1+GOL,1(s)| had to be greater than the high gas
throughput |PRGA12| values at frequencies below 0.04 rad/min. For higher frequencies at which
setpoint tracking was still required, |1+ GOL,1(s)| had to be greater than the corresponding |PRGA11|
values.

For CO2 train #7, it was observed that |CLDG11| was greater than 1 for frequencies up to 0.05 and
0.25 rad/min respectively for the medium and high gas throughput conditions. This implied that control
was needed in Loop 1 for disturbance rejection for frequencies up to 0.25 rad/min. Even so, tracking
the sweet gas flow rate setpoint was still more difficult due to the larger values of |PRGA12| over the
entire frequency range. However, the sweet gas CO2 content setpoint became the more difficult
setpoint to track (|PRGA12|<|PRGA11|) for the low and medium gas throughput conditions at

187
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

frequencies above 0.04 and 0.19 rad/min respectively. Nevertheless, to ensure acceptable control
performance for Loop 1 for CO2 train #7 over the three operating regions, |1+GOL,1(s)| had to be
greater than the high gas throughput |PRGA12| values over the frequency range required for setpoint
tracking.

For both CO2 trains, |CLDG21| was observed to be less than 1 over the examined frequency range,
indicating that changes in the raw gas CO2 content had little effect on the sweet gas flow rate. The
|PRGA22| values were considerably larger than those for |PRGA21| and |CLDG21|, which unsurprisingly
implied that changes in the sweet gas flow rate setpoint had the most significant effect on the sweet
gas flow rate. Consequently, to ensure satisfactory performance for Loop 2 for CO2 trains #1 and #7
over the three operating regions, |1+GOL,2(s)| had to be greater than the high gas throughput |PRGA22|
values over the frequency range required for setpoint tracking.

To test the validity of the above findings, BLT tuning was applied to tune the SISO controllers for the
RGF-RSF diagonal control structures for the two CO2 trains. This is discussed in the following section.

Table 10.3.1: Steady-state values for the PRGA and CLDG for the selected configuration for the RGF-RSF
diagonal control structure.

CO2 Train #1 CO2 Train #7


PRGA CLDG PRGA CLDG
High Gas ª 1.03  1.11º ª 0.67 º ª 1.20  4.80º ª 2.23 º
« » « » « » « »
Throughput ¬ 0.03 1.03 ¼ ¬ 0.09 ¼ ¬ 0.05 1.20 ¼ ¬ 0.11¼
Medium Gas ª 1.03  1.11º ª 0.63 º ª 1.08  2.48º ª 1.17 º
« » « » « » « »
Throughput ¬ 0.03 1.03 ¼ ¬ 0.10 ¼ ¬ 0.04 1.08 ¼ ¬ 0.08 ¼
Low Gas ª 1.03  1.11º ª 0.61 º ª 1.04  1.46 º ª 0.67 º
« » « » « » « »
Throughput ¬ 0.03 1.03 ¼ ¬ 0.10 ¼ ¬ 0.03 1.04 ¼ ¬ 0.06 ¼

188
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

PRGA11,
PRGA11, PRGA
PRGA22
22 PRGA(1,2)
PRGA12 PRGA(2,1)
PRGA21 CLDG(1,1)
CLDG11 CLDG(2,1)
CLDG21

1.0.E+01 1.0.E+01

1.0.E+00 1.0.E+00

1.0.E-01 1.0.E-01
|PRGAij|

|CLDGij|
1.0.E-02 1.0.E-02

1.0.E-03 1.0.E-03

1.0.E-04 1.0.E-04

1.0.E-05 1.0.E-05
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02 1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02
Frequency (rad/min) Frequency (rad/min)
(a)
1.0.E+01 1.0.E+01

1.0.E+00 1.0.E+00

1.0.E-01 1.0.E-01
|PRGAij|

|CLDGij|
1.0.E-02 1.0.E-02

1.0.E-03 1.0.E-03

1.0.E-04 1.0.E-04

1.0.E-05 1.0.E-05
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02 1.0.E-03 1.0.E-02 1.0.E-01 1.0.E+00 1.0.E+01 1.0.E+02

Frequency (rad/min) Frequency (rad/min)


(b)
Figure 10.3.1: Frequency plots for |PRGAij| and |CLDGij|. (a) CO2 train #1. (b) CO2 train #7. The solid lines
(––) represent the high gas throughput conditions, the dashed lines (---) represent the medium gas
throughput conditions, and the crosses (×××) represent the low gas throughput conditions.

189
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

10.4 BLT Tuning


As described in Section 2.5.6, BLT tuning is a well-known and simple method for tuning diagonal
control structures. This section discusses the application of this method to determine the controller
settings for the RGF-RSF diagonal control structure defined in Table 10.2.2 for CO2 trains #1 and #7.
These controller settings were required in order to perform dynamic simulations of the diagonal control
structure so that its performance could be assessed, as described in the next section.

Using equation (2.5.14), the cross-over frequency Zco was obtained with the Microsoft Excel® Solver
tool. This parameter was used to calculate the ultimate gain Ku and period Pu via equations (2.5.15)
and (2.5.16). The resulting values for Zco, Ku and Pu are given in Table 10.4.1, along with the
corresponding Ziegler-Nichols controller gain KcZN and integral time constant WIZN. The Ziegler-Nichols
settings were obtained using the PI tuning rules provided in Table 2.5.2. Derivative action was not
considered for the RGF-RSF diagonal control structure since the SISO loops were not slow enough,
as indicated by their moderate time constants, to require derivative action to tighten the dynamic
response. Additionally, signal noise is expected in practice for the CO2 trains, which precludes the use
of derivative action (Svrcek et al., 2006).

Following the method outlined in Section 2.5.6, detuning factors F were determined for each CO2 train
and operating region. An example MATLAB® script for BLT tuning is provided in Section I.4 of
Appendix I. The resulting F factors and SISO controller settings for the diagonal control structures are
included in Table 10.4.1, while plots of the corresponding W functions, as defined by equation
(2.5.19), are provided as Figure 10.4.1. According to these plots, the BLT tuned control systems were
closed-loop stable since the W curves did not encircle the critical point (-1, 0). The controller settings
obtained from the BLT tuning procedure were therefore considered valid.

It was also observed that the CO2 train #7 loops were significantly more detuned by the BLT tuning
method, compared to their equivalent loops in CO2 train #1. Consequently, more conservative
controller settings (i.e. smaller Kc and larger WI values) were required for CO2 train #7 in order to
achieve a similar level of compromise between robustness and performance as for CO2 train #1. This
reflected the more difficult control problem posed by the former CO2 train, which was implied in the
previous sections by the PRGA, CLDG and RGA.

190
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

Table 10.4.1: BLT tuning parameters for the RGF-RSF diagonal control structures for CO2 trains #1 and
#7.

Process Parameters Controller Parameters


Zco Ku Pu KcZN WI
ZN
F Kc WI
(rad/min) (min) (min)  (min)
CO2 Train #1
High Gas Loop 1 2.40 -31.67 2.61 -14.39 1.19 -6.76 2.53
2.14
Throughput Loop 2 9.80 15.88 0.64 7.22 0.29 3.39 0.62
Medium Gas Loop 1 2.40 -41.43 2.63 -18.83 1.19 -7.50 3.00
2.52
Throughput Loop 2 9.69 22.67 0.65 10.31 0.29 4.11 0.74
Low Gas Loop 1 2.39 -50.37 2.63 -22.90 1.20 -8.15 3.36
2.82
Throughput Loop 2 9.65 27.06 0.65 12.30 0.30 4.38 0.83

CO2 Train #7
High Gas Loop 1 2.42 -24.16 2.60 -10.98 1.18 -2.73 4.75
4.02
Throughput Loop 2 9.72 20.43 0.65 9.29 0.29 2.31 1.18
Medium Gas Loop 1 2.40 -35.92 2.62 -16.33 1.19 -4.25 4.57
3.85
Throughput Loop 2 9.65 26.99 0.65 12.27 0.30 3.20 1.14
Low Gas Loop 1 2.39 -45.94 2.63 -20.88 1.20 -6.33 3.95
3.29
Throughput Loop 2 9.64 28.09 0.65 12.77 0.30 3.87 0.98

High gas throughput Medium gas throughput Low gas throughput

1.0 1.0

0.5 0.5
Imaginary(W)

Imaginary(W)

0.0 0.0

-0.5 -0.5

-1.0 -1.0

-1.5 -1.5

-2.0 -2.0
-2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0

Real(W) Real(W)
(a) (b)
Figure 10.4.1: Plots of the scalar function W. (a) CO2 train #1. (b) CO2 train #7. The critical point (-1, 0) is
highlighted in red.

191
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

10.5 Diagonal Control Structure Dynamic Simulations


Having obtained suitable controller settings from the BLT tuning method, a series of dynamic
simulations were performed to analyse the dynamic performance of the RGF-RSF diagonal control
structure for CO2 trains #1 and #7. The necessary calculations were performed in MATLAB® 5.2, and
an example MATLAB® script is provided in Section I.5 in Appendix I.

As a matter of interest, each set of controller settings was applied to all three operating regions for
their respective CO2 train. The purpose of this comparison was to determine whether it was possible
to achieve acceptable control performance for the two CO2 trains using the same set of controller
settings across the different operating regions.

For each simulation case, the open-loop transfer functions GOL,i(s) were first determined to verify if
condition (2.5.12) for achieving acceptable disturbance rejection and setpoint performance was
satisfied over the relevant frequency ranges identified in Section 10.3. Figure 10.5.1 compares the
resulting frequency-dependent |1+GOL,i(s)| curves against the relevant PRGA elements identified in
Section 10.3. It was observed that the |1+GOL,i(s)| values were greater than the corresponding PRGA
elements at frequencies below approximately 0.1 to 1 rad/min. These frequencies corresponded to
the frequency range usually needed for setpoint tracking, as required to satisfy condition (2.5.12) for
the PRGA elements. This indicated that the controller settings obtained from the BLT tuning method
could produce a reasonable control performance for both CO2 trains, regardless of operating region.

However, a comparison of Figures 10.3.1 and 10.5.1 showed that for CO2 train #7, the |1+GOL,1(s)|
values were not always larger than the corresponding CLDG elements for which control was required
for disturbance rejection (|CLDGik|>1). This implied that the Loop 1 controller settings for CO2 train #7
may not provide adequate disturbance rejection at higher frequencies.

In order to test the control performance of the BLT tuned controller settings for the RGF-RSF diagonal
control structure, the servo and regulator closed-loop transfer function matrices GCLS(s) and GCLR(s)
for setpoint tracking and disturbance rejection, respectively, were defined as follows:

G CLS (s) >I  Gp (s) ˜ G c (s)@1 ˜ Gp (s) ˜ G c (s) (10.5.1)

G CLR (s) >I  Gp (s) ˜ G c (s)@1 ˜ G d (s) (10.5.2)

where I is the identity matrix, Gp(s) is the process transfer function matrix, Gc(s) is the controller
transfer function matrix, and Gd(s) is the disturbance transfer function matrix.

Unit magnitude disturbance and setpoint changes were applied separately to GCLS(s) and GCLR(s) for
each case to obtain the closed-loop step response curves for each controlled variable. The resulting
response curves were observed to be quite similar for the three operating conditions. Consequently
for brevity, the closed-loop step response curves for the high gas throughput conditions are given in

192
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

Figures 10.5.2 and 10.5.3, while the corresponding curves for the medium and low gas throughput
conditions are included as Figures I.5.1 to I.5.4 in Appendix I. The vertical axes for these figures are
dimensionless and are equivalent to those for the response curves depicted in Table 2.5.1. To
achieve this, the process outputs were expressed in perturbation form and were scaled by dividing
each output variable by its allowed range and then by the similarly scaled step change.

It was observed that the closed-loop step responses for all the simulation cases were reasonably rapid
with regard to the process time constants for the two CO2 trains, as steady-state was achieved within
one hour. The controlled variables responded most quickly at the high gas throughput conditions,
while the low gas throughput conditions produced the slowest closed-loop responses. It was also
noted that the closed-loop step responses for Loop 2 tended to be much faster than the corresponding
Loop 1 responses. These findings were consistent with the observed trends for the process time
constants given in Table 10.2.2. The process time constants for Loop 1 were much larger than those
for Loop 2, and the process time constants for both loops increased with increasing gas throughput.
No offset was observed at steady-state due to the integral action of the controllers.

For CO2 train #1, the effect of a unit magnitude change in the raw gas CO2 content on the closed-loop
response for the different controller settings are illustrated in the leftmost plots in Figures 10.5.2, I.5.1
and I.5.3. The corresponding step responses for unit magnitude changes in the sweet gas CO2
content and flow rate setpoints are depicted in the middle and rightmost plots in these three figures.
For both disturbance rejection and setpoint tracking, it was observed that the most rapid responses
with the largest overshoots were produced by the controller settings derived for the high gas
throughput conditions. In contrast, the low gas throughput controller settings generated the most
damped step responses.

This observed trend was consistent with the high gas throughput controller settings having the
smallest integral time constants, while the largest integral time constants were associated with the low
gas throughput controller settings. It should be noted that despite its slightly more aggressive nature,
the high gas throughput controller settings still produced relatively damped closed-loop step
responses for the three operating regions. Each of the three sets of controller settings for CO2 train #1
in Table 10.4.1 therefore generated acceptable control performances for the RGF-RSF diagonal
control structure, regardless of the operating region, which was consistent with the earlier analysis of
the PRGA elements and |1+GOL,i(s)| values.

For CO2 train #7, the leftmost plots in Figures 10.5.3, I.5.2 and I.5.4 show the closed-loop step
responses for a unit magnitude change in the raw gas CO2 content, while the centre and rightmost
plots illustrate the corresponding effects of unit magnitude changes in the two setpoints. In contrast to
CO2 train #1, the high gas throughput controller settings were found to generate the most damped
closed-loop step responses for the disturbance rejection and setpoint tracking cases. The
corresponding responses produced by the low gas throughput controller settings were the least

193
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

damped, and in the case of the high gas throughput conditions, the responses were distinctly under-
damped.

This observed control behaviour was consistent with the high gas throughput controller settings having
the smallest controller gains and largest integral time constants, while the low gas throughput settings
were the most aggressive with the largest controller gains and the integral time constants. Since none
of the controller settings for CO2 train #7 produced overly oscillatory or unstable closed-loop
responses, the three sets of controller settings were each applicable to the RGF-RSF diagonal control
structure, regardless of the operating region. While this agreed with the earlier analysis of the PRGA
elements and |1+GOL,i(s)| values, it also contrasted with the analysis of the CLDG elements and
|1+GOL,i(s)| values since reasonable disturbance rejection was achieved with the different controller
settings over the three operating regions.

The results of the above-discussed process control studies therefore indicated that the implementation
of the RGF-RSF diagonal control structure identified in this work could potentially enable the automatic
control of the two key process outputs, the sweet gas CO2 content and flow rate, for the two most
dissimilar CO2 trains, CO2 trains #1 and #7. These studies demonstrated a potential application for
the HYSYS® process models of the Moomba CO2 trains, in that they could be used to facilitate the
development of a fully automated control strategy for the CO2 trains.

As the above described control studies were purely for demonstrative purposes, and not a key focus
of this work, the diagonal control structures were not rigorously tested for performance in the presence
of process noise. Neither were they implemented in HYSYS® to investigate the model mismatch
between the HYSYS® and MATLAB® process models. It should be noted that these measures are
highly recommended for any future work as they are essential to the development of a viable control
scheme for the CO2 trains. However, before such work can proceed, the dynamic process models
need to be validated against detailed high quality step-test data from the CO2 trains, which was not
accessible at the time.

A final point to note, the reasonable control performances observed in the dynamic process control
studies indicated that the simple PID controllers in the diagonal control structure provided sufficient
control action, despite the non-linear behaviour of the CO2 trains. This justified the exclusion of more
advanced controllers from this work, and suggested that more complex multivariable controllers may
not be necessary in order to achieve reasonable control of the Moomba CO2 trains.

194
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

1+GOLH(s)
1+GOL,H(s) 1+GOLH(s)
1+GOL,M(s) 1+GOLH(s)
1+GOL,L(s) PRGA11,PRGA22
PRGA11, PRGA22 PRGA12
PRGA12

1.0E+02 1.0E+02

|1+GLOL,1(s)|, |PRGA1j|
1.0E+01
|1+GOL,1(s)|, |PRGA1j|

1.0E+01

1.0E+00
1.0E+00
1.0E-01

1.0E-01
1.0E-02

1.0E-02 1.0E-03
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02 1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02
Frequency (rad/min) Frequency (rad/min)
1.0E+01 1.0E+02

1.0E+01

|1+GOL,2(s)|, |PRGA2j|
|1+GOL,2(s)|, |PRGA2j|

1.0E+00

1.0E+00
1.0E-01
1.0E-01

1.0E-02
1.0E-02

1.0E-03 1.0E-03
1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02 1.0E-03 1.0E-02 1.0E-01 1.0E+00 1.0E+01 1.0E+02
Frequency (rad/min) Frequency (rad/min)
(a) (b)
Figure 10.5.1: Frequency plots of |1+GOL,i(s)| and |PRGAij|. (a) CO2 trains #1. (b) CO2 train #7. The solid
lines (––) represent the high gas throughput conditions, the dashed lines (---) represent the medium gas
throughput conditions, and the crosses (×××) represent the low gas throughput conditions. The
subscripts H, M and L respectively denote the controller settings corresponding to the high, medium and
low gas throughput conditions.

195
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7
High gas throughput controller settings Medium gas throughput controller settings Low gas throughput controller settings

0.160 1.500 0.400

1.200

Sweet Gas CO2 Content


Sweet Gas CO2 Content

Sweet Gas CO2 Content


0.120 0.300

0.900
0.080 0.200

0.600

0.040 0.100
0.300

0.000 0.000
0.000

-0.040 -0.300 -0.100


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60
196

Time (min) Time (min) Time (min)


0.000 0.006 1.400

0.005 1.200
Sweet Gas Flow Rate

Sweet Gas Flow Rate

Sweet Gas Flow Rate


-0.002
0.004 1.000

0.003 0.800
-0.004
0.002 0.600

0.001 0.400
-0.006

0.000 0.200

-0.008 -0.001 0.000


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60

Time (min) Time (min) Time (min)

(a) (b) (c)


Figure 10.5.2: CO2 train #1 closed-loop step response curves at the high gas throughput conditions. (a) Unit magnitude change in the raw gas CO2 content at
0 min. (b) Unit magnitude change in the sweet gas CO2 content setpoint at 0 min. (c) Unit magnitude change in the sweet gas flow rate setpoint at 0 min.
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7
High gas throughput controller settings Medium gas throughput controller settings Low gas throughput controller settings

0.600 1.800 3.000

0.500 1.500
Sweet Gas CO2 Content

Sweet Gas CO2 Content

Sweet Gas CO2 Content


2.000
0.400 1.200

1.000
0.300 0.900

0.200 0.600
0.000

0.100 0.300
-1.000
0.000 0.000

-0.100 -0.300 -2.000


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60
197

Time (min) Time (min) Time (min)


0.002 0.030 1.200

0.000 1.000
0.020
Sweet Gas Flow Rate

Sweet Gas Flow Rate

Sweet Gas Flow Rate


-0.002 0.800
0.010

-0.004 0.600

0.000
-0.006 0.400

-0.010
-0.008 0.200

-0.010 -0.020 0.000


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60

Time (min) Time (min) Time (min)

(a) (b) (c)


Figure 10.5.3: CO2 train #7 closed-loop step response curves at the high gas throughput conditions. (a) Unit magnitude change in the raw gas CO2 content at
0 min. (b) Unit magnitude change in the sweet gas CO2 content setpoint at 0 min. (c) Unit magnitude change in the sweet gas flow rate setpoint at 0 min.
Chapter 10: Process Control Studies for CO2 Removal Trains #1 and #7

10.6 Summary
In summary, a simple and fully automatic control strategy was developed in this chapter for each of
the two most dissimilar CO2 trains, CO2 trains #1 and #7, to control the two key process outputs: the
sweet gas flow rate and CO2 content. This demonstrated the potential application of the HYSYS®
process models of the Moomba CO2 trains in facilitating the development of a fully automated control
system for the CO2 trains.

As previously mentioned, the diagonal control structure is the simplest and most widely used form of
multivariable control. However, its application to hot potassium carbonate process operations, like the
CO2 trains, has not been documented in the literature. Consequently, this chapter focussed on the
development of diagonal control structures for the two CO2 trains, based on the simple FOPDT
process transfer functions derived in the previous chapter. All the necessary calculations for this
exercise were performed in the MATLAB® 5.2 numerical computing environment.

Due to the different possible combinations of the process inputs and outputs, a number of alternative
diagonal control structures had to be considered for each CO2 train. Their sensitivity to process
uncertainties and disturbances was evaluated, and the optimal set of process inputs was found to
comprise of the raw gas flow rate and the reboiler steam flow rate, which corresponded to the RGF-
RSF diagonal control structure. The optimal control structure configuration was then identified from
the possible variable pairing permutations by analysing the system stability and loop interactions. It
was determined that for the RGF-RSF structure, the raw gas flow rate should be paired with the sweet
gas flow rate, while the sweet gas CO2 content should be paired with the reboiler steam flow rate.

Dynamic MATLAB® simulations were performed to test the performance of the RGF-RSF diagonal
control structure for the two CO2 trains. The relevant controller parameters were obtained using the
BLT tuning method, and the frequency ranges of interest were determined from the analysis of the
control performance of the RGF-RSF structure. It was observed that acceptable disturbance rejection
and setpoint tracking was achieved with the selected diagonal control structure for both CO2 trains.
This indicated that the RGF-RSF diagonal control structure could potentially be used to facilitate the
fully automatic control of the Moomba CO2 trains. However, further work is required to confirm this.

198
Chapter 11: Conclusions and Recommendations

CHAPTER 11

CONCLUSIONS AND
RECOMMENDATIONS

199
Chapter 11: Conclusions and Recommendations

11.1 Conclusions
In this thesis, a new rigorous non-equilibrium rate-based model for chemical absorption and
desorption was developed, and was successfully applied to the hot potassium carbonate process CO2
Removal Trains at the Santos Moomba Processing Facility. Unlike previous rate-based models for the
hot potassium carbonate process, the model presented in this work included rigorous electrolyte
thermodynamics for both the absorber and regenerator columns. This model also differed in that it
considered the use of reaction rate expressions for representing the effect of chemical reactions.
However, enhancement factors were used instead in the final form of the model as preliminary
simulations showed that minimal improvement was obtained with the more rigorous rate expressions,
despite the greater associated computing time and model accuracy.

The non-equilibrium rate-based process models of the CO2 trains were implemented in the equation-
based Aspen Custom Modeler® simulation platform. This simulation environment enabled rigorous
thermodynamic and physical property calculations via the Aspen Properties® software package.
Literature data were used to determine the parameters for the relevant Aspen Properties® property
models. Where these were found to be inadequate, the data were instead used to regress the
parameters for alternative property models from literature or to develop empirical correlations.

Preliminary simulations identified the need for correction factors in order to achieve closer agreement
between the model predictions for the absorber columns and the plant data. In contrast, the results for
the regenerator column model agreed well with the plant data. A liquid phase enthalpy correction was
developed for the absorber column model to generate a more reasonable temperature profile, while
adjustment factors for the absorber effective interfacial area were derived to provide closer agreement
between the predicted and data values for the sweet gas CO2 content. Following the application of
these model adjustments, the CO2 train process models were successfully validated against an
independent set of steady-state plant data.

The validation of these Aspen Custom Modeler® process models demonstrated that the non-
equilibrium rate-based model presented in this work was suitable for modelling the hot potassium
carbonate process. However, this approach could not be easily implemented in HYSYS®, Santos’s
preferred simulation environment, due to the lack of electrolyte components and property models and
the limitations of the HYSYS® column operations in accommodating chemical reactions and non-
equilibrium column behaviour. Importation of the Aspen Custom Modeler® process models into
HYSYS® was considered, but was disregarded due to the significant associated computation time.
Instead, a novel approach involving hypothetical HYSYS® components and the HYSYS® column
stage efficiencies was developed to compensate for the restricted capabilities of this simulation
platform.

Hypothetical K2CO3* and H2O* components were created to facilitate the simulation of the hot
potassium carbonate system in HYSYS®. In addition, the standard Peng-Robinson property package

200
Chapter 11: Conclusions and Recommendations

model was modified to include tabular physical property models to accommodate the properties of this
electrolyte system. Literature data were used to derive parameters for the relevant thermodynamic
and physical property models. To compensate for the limitations of the standard Peng-Robinson
property package, stage efficiency correlations were derived from parametric studies performed on the
Aspen Custom Modeler® CO2 train process models. These correlations related various operating
parameters to the column performance, and were used to represent the effects of the non-equilibrium
column behaviour and the liquid phase chemistry for the hot potassium carbonate process.
Preliminary simulations indicated that the stage efficiency correlations were only necessary for the
absorber columns as the regenerator columns were satisfactorily represented using the default
equilibrium stages.

The combination of the stage efficiency correlations and the hypothetical components was found to
enable the reasonable representation of the potassium carbonate system in HYSYS®. Consequently,
detailed steady-state process models of the CO2 trains were created in this simulation environment
using technical data from the Santos TIMS controlled documentation database. As for the Aspen
Custom Modeler® process models, the HYSYS® CO2 train process models were successfully
validated against an independent set of steady-state plant data. It was noted that HYSYS® process
models predicted a considerably higher H2S content for the sweet gas, compared to the Aspen
Custom Modeler® process models. However, this discrepancy was deemed acceptable for this work
since the primary focus of the Moomba CO2 trains is the removal of CO2, and in that regard, the
HYSYS® CO2 train process models were found to produce excellent results.

To demonstrate a potential application of the HYSYS® process models, dynamic process models of
CO2 trains #1 and #7, the two most dissimilarly configured trains, were created by modifying the
steady-state HYSYS® process models. These changes to the model configuration and specifications
were necessary due to the fundamental differences between the two simulation modes. The dynamic
process models were used to investigate the transient behaviour of the two CO2 trains over a range of
operating conditions. Particular attention was given to the key process outputs (the sweet gas CO2
content and flow rate), inputs (the raw gas flow rate, the lean solution flow rate, the reboiler steam flow
rate and the regenerator liquid level) and disturbance (the raw gas CO2 content). Simple first-order
plus dead time process transfer functions relating these variables were derived from the results of the
dynamic process case studies.

The process transfer functions were used to develop individual diagonal control structures to fully
automate the control of CO2 trains #1 and #7. This simple form of multivariable control was
considered in this work because despite its wide usage, its application to hot potassium carbonate
process operations has not been documented in literature. A number of analysis techniques were
applied to evaluate the sensitivity, stability and controllability of the potential diagonal control
structures. The two CO2 trains were found to share the same optimal diagonal control structure, in

201
Chapter 11: Conclusions and Recommendations

which the sweet gas CO2 content was controlled by the reboiler steam flow rate while the sweet gas
flow rate was controlled by the raw gas flow rate.

The results of the controllability analyses were confirmed by dynamic simulations in the MATLAB®
numerical computing environment, which showed that effective disturbance rejection and setpoint
tracking were achieved for both CO2 trains with this control structure. This suggested that the
identified control system was independent of the CO2 train configuration, and could potentially be
applied to the remaining five CO2 trains to enable the fully automatic control of all seven CO2 trains.
The process control studies performed in this work therefore demonstrated a possible useful
application for the HYSYS® process models of the Moomba CO2 trains, in that they could be used to
facilitate the development of a fully automated control strategy for the Moomba CO2 Removal Trains.

In conclusion, a new rigorous non-equilibrium rate-based model was developed in this work for the hot
potassium carbonate process. To demonstrate this model, it was implemented in the Aspen Custom
Modeler® simulation platform and was successfully used to simulate the CO2 Removal Trains at the
Santos Moomba Processing Facility. However, this model could not be readily applied in HYSYS®,
the preferred simulation environment for Santos, so a novel approach was followed to develop
detailed CO2 train process models in HYSYS® using the Aspen Custom Modeler® process models. A
potential use for the resulting HYSYS® process models was then demonstrated through the
development of diagonal control structures to facilitate the fully automatic control of the CO2 trains.

Further work is recommended to extend the application and improve the modelling accuracy of the
Aspen Custom Modeler® and HYSYS® CO2 train process models. It is also suggested that work be
performed on the process control of the CO2 trains to investigate the effectiveness of more complex
control algorithms compared to the simple diagonal control structure. These recommendations are
listed in the next section.

202
Chapter 11: Conclusions and Recommendations

11.2 Recommendations
The recommendations for future work are divided into the three categories:

1. Recommendations for further Aspen Custom Modeler® simulation work:


(a) Validate the Aspen Custom Modeler® column models against a larger set of plant data,
including data for the current conditions at the Moomba Processing Facility and data for start-
up and shut-down conditions;
(b) Investigate the effects of different types of random packing on the hydrodynamic behaviour
and performance of the column models; and
(c) Extend the model process chemistry to include the effects of amine activators, such as MEA,
DEA and ACT-1, and other solution additives like anti-foaming agents and corrosion inhibitors.

2. Recommendations for further HYSYS® simulation work:


(a) Extend the operating range over which the HYSYS® column stage efficiency correlations are
valid to include a wider range of conditions, such as start-up and shut-down;
(b) Incorporate the effect of different types of random packing into the column stage efficiency
correlations;
(c) Extend the model component system to include amine activators and other solution additives;
(d) Adjust the physical and thermodynamic properties of the hypothetical K2CO3* component to
enable the entrainment of the potassium carbonate solution, as observed in practice;
(e) Convert the steady-state process models for CO2 trains #2 to #6 into dynamic models; and
(f) Validate the dynamic CO2 train process models against online plant data.

3. Recommendations for further process control studies:


(a) Derive process transfer functions for CO2 trains #2 to #6 and, following the procedure
described in this work for CO2 trains #1 and #7, develop a diagonal control structure for each
of these CO2 trains which could facilitate the fully automatic control of the Moomba CO2 trains;
and
(b) Develop alternative control systems for the CO2 trains using more complex control algorithms,
such as multivariable controllers and model predictive control, and compare their effectiveness
against the simple diagonal control structure.

203
References

References
Abrams, D. S. and Prausnitz, J. M. (1975). "Statistical thermodynamics of liquid mixtures: a new expression for the
Gibbs energy of partly or completely miscible systems." The American Institute of Chemical Engineers
Journal 21: 116 - 128.

Alatiqi, I. M., Sabri, M. F. and Alper, E. (1993). "Multivariable control system design of CO2/amine
absorber/desorber units by using a rigorous steady-state model." Gas Separation and Purification 7(2):
119 - 121.

Albrecht, J. J., Kershenbaum, L. S. and Pyle, D. L. (1980). "Identification and linear multivariable control in an
absorption-desorption pilot plant." AIChE Journal 26(3): 496 - 504.

Al-Ramdhan, H. A. (2001). A Rate-Based Model for the Design and Simulation of a Carbon Dioxide Absorber
Using the Hot Potassium Carbonate Process. PhD Thesis. Golden, Colorado School of Mines

Anderko, A., Wang, P. and Rafal, M. (2002). "Electrolyte solutions: from thermodynamic and transport property
models to the simulation of industrial processes." Fluid Phase Equilibria 194 -197: 123 - 142.

Armand Products Company (1998). Potassium Carbonate Handbook. Princeton NJ, Armand Products Company.

Aseyev, G. G. and Zaytsev, I. D. (1996). Volumetric Properties of Electrolyte Solutions: Estimation Methods and
Experimental Data. New York, Begell House.

Aspen Technology Inc. (2003). Aspen Physical Property System: Physical Property Methods and Models. v12.1.
Cambridge MA, Aspen Technology Inc.

Aspen Technology Inc. (2006). Solution ID: 118492. Available from http://support.aspentech.com. Accessed: 10
August 2006.

Astarita, G., Savage, D. W. and Bisio, A. (1983). Gas Treating with Chemical Solvents. New York, John Wiley and
Sons.

Astarita, G., Savage, D. W. and Longo, J. M. (1981). "Promotion of CO2 mass transfer in carbonate solutions."
Chemical Engineering Science 36(3): 581 - 588.

Åström, K. J. and Hagglünd, T. (1984). "Automatic tuning of simple regulators with specifications on phase and
amplified margins." Automatica 20(5): 645 - 651.

Åström, K. J. and Hagglünd, T. (1995). PID Controllers: Theory, Design and Tuning. 2nd ed. Research Triangle
Park, Instrument Society of America.

Austgen, D. M., Rochelle, G. T., Peng, X. and Chen, C.-C. (1989). "Model of vapour-liquid equilibria for aqueous
acid gas-alkanolamine systems using the Electrolyte-NRTL equation." Industrial and Engineering
Chemistry Research 28(7): 1060 - 1073.

Barber, C. (2007). Personal communication. 6 June.

Baroncini, C., Filippo, P. D., Latini, G. and Pacetti, M. (1980). "An improved correlation for the calculation of liquid
thermal conductivity." International Journal of Thermophysics 1(2): 159 - 175.

Bartoo, R. K. (1984). "Removing acid gas by the Benfield Process." Chemical Engineering Progress 80(10): 35 -
39.

Bedekar, S. G. (1955). "Properties of sodium carbonate-sodium bicarbonate solutions." Journal of Applied


Chemistry 5(2): 72 - 75.

Benson, H. E. and Field, J. H. (1959). Method for Separating CO2 and H2S from Gas Mixtures. Patent No.
2,886,405. United States.

Benson, H. E., Field, J. H. and Jimeson, R. M. (1954). "CO2 absorption employing hot potassium carbonate
solutions." Chemical Engineering Progress 50(7): 356 - 364.

204
References

Benson, H. E. and McCrea, D. H. (1979). Removal of acid gases from hot gas mixtures. Patent No. 4,160,810.
United States.

Benson, H. E. and Parrish, R. W. (1974). "HiPure process removes CO2/H2S." Hydrocarbon Processing 53(4): 81 -
82.

Bergman, D. F. (1976). Predicting the Phase Behaviour of Natural Gas in Pipelines. PhD Thesis. Ann Arbor,
University of Michigan

Bergman, D. F., Tek, M. R. and Katz, D. L. (1975). Retrograde Condensation in Natural Gas Pipelines. New York,
American Gas Association.

Billet, R. and Schultes, M. (1999). "Prediction of mass transfer columns with dumped and arranged packings:
Updated summary of the calculation method of Billet and Schultes." Chemical Engineering Research and
Design : Transactions of the Institution of Chemical Engineers A 77: 498 - 504.

Bingham, E. C. (1922). Fluidity and Plasticity. New York, McGraw-Hill.

Bird, R. B., Stewart, W. E. and Lightfoot, E. N. (1960). Transport Phenomena. New York, John Wiley and Sons.

Bocard, J. P. and Mayland, B. J. (1962). "New charts for hot carbonate process." Hydrocarbon Processing and
Petroleum Refiner 41(4): 128 - 132.

Bravo, J. L. and Fair, J. R. (1982). "Generalised correlation for mass transfer in packed distillation columns."
Industrial and Engineering Chemistry Process Design and Development 21: 162 - 170.

Brelvi, S. W. and O'Connell, J. P. (1972). "Corresponding states correlations for liquid compressibility and partial
molal volumes of gases at infinite dilution in liquids." AIChE Journal 18(6): 1239 - 1243.

Breslau, B. R. and Miller, I. F. (1972). "On the viscosity of concentrated aqueous electrolyte solutions." Journal of
Physical Chemistry 74(5): 1056 - 1061.

Breslau, B. R., Welsh, P. B. and Miller, I. F. (1974). "Estimating the viscosities of aqueous electrolytic solutions."
Chemical Engineering 81(8): 112 - 113.

Bristol, E. H. (1966). "On a new measure of interactions for multivariable process control." IEEE Transactions on
Automatic Control AC-11: 133 -134.

Britt, H. I. and Luecke, R. H. (1973). "The estimation of parameters in nonlinear, implicit models." Technometrics
15(2): 223 - 247.

Brock, J. R. and Bird, R. B. (1955). "Surface tension and the principle of corresponding states." AIChE Journal 1:
174 - 177.

Cavett, R. H. (1964). Physical data for distillation calculations: vapour-liquid equilibria. 27th Midyear Meeting, API
Division of Refining, San Francisco, USA.

Chakravarty, T., Phukan, U. K. and Weiland, R. H. (1985). "Reaction of acid gases with mixtures of amines."
Chemical Engineering Progress 81(4): 32 - 36.

Charpentier, J.-C. (1981). "Mass-transfer rates in gas-liquid absorbers and reactors." Advances in Chemical
Engineering 11: 1 - 133.

Chase, M. W., Jr. (1998). NIST-JANAF Thermochemical Tables. 4th ed. Woodbury, American Institute of Physics.

Chen, C.-C. (1980). Computer Simulation of Chemical Processes with Electrolytes. Sc. D. Thesis. Cambridge,
Massachusetts Institute of Technology

Chen, C. C., Britt, H. I., Boston, J. F. and Clarke, W. M. (1983). "Thermodynamic property evaluation in computer-
based flowsheet simulation for aqueous electrolyte systems." AIChE Symposium Series 79(229): 126 -
134.

205
References

Chen, C.-C., Britt, H. I., Boston, J. F. and Evans, L. B. (1982). "Local composition model for excess Gibbs energy
of electrolyte systems." The American Institute of Chemical Engineers Journal 28(4): 588 - 596.

Chen, C.-C. and Evans, L. B. (1986). "A local composition model for the excess Gibbs energy of aqueous
electrolyte systems." The American Institute of Chemical Engineers Journal 32(3): 444 - 454.

Chernen'kaya, E. I. and Revenko, S. S. (1975). "Viscosity and density of K2CO3 and KHCO3 solutions." Zhurnal
Prikladnoi Khimii 48(1): 126 - 130.

Chernen'kaya, E. I. and Vernigora, G. A. (1973). "Thermal conductivities of solutions of Na2CO3, K2CO3, KCl and
K2SO4 at various temperatures." Zhurnal Prikladnoi Khimii 46(7): 1603 - 1605.

Chung, T. H., Ajlan, M., Lee, L. L. and Starling, K. E. (1988). "Generalised multiparameter correlation for nonpolar
and polar fluid transport properties." Industrial and Engineering Chemistry Research 27(4): 671 - 679.

Correia, R. J. and Kestin, J. (1980). "Viscosity and density of aqueous Na2CO3 and K2CO3 solutions in the
temperature range 20-90°C and the pressure range 0-30 MPa." Journal of Chemical and Engineering
Data 25: 201 - 206.

Criss, C. M. and Cobble, J. W. (1964a). "The thermodynamic properties of high temperature aqueous solutions.
IV. Entropies of the ions up to 200° and the correspondence principle." Journal of the American Chemical
Society 86(12): 5385 - 5390.

Criss, C. M. and Cobble, J. W. (1964b). "The thermodynamic properties of high temperature aqueous solutions. V.
The calculation of ionic heat capacities up to 200°, entropies and heat capacities above 200°." Journal of
the American Chemical Society 86(12): 5390 - 5393.

Cullinane, J. T. and Rochelle, G. T. (2004). "Carbon dioxide absorption with aqueous potassium carbonate
promoted by piperazine." Chemical Engineering Science 59: 3619 - 3630.

Danckwerts, P. V. (1951). "Significance of liquid film coefficients in gas absorption." Industrial and Engineering
Chemistry 43(6): 1460 - 1467.

Danckwerts, P. V. (1970). Gas-Liquid Reactions. New York, McGraw-Hill Book Company.

Danckwerts, P. V. and Kennedy, A. M. (1954). "Kinetics of liquid-film process in gas absorption. Part I: Models of
the absorption process." Transactions of the Institution of Chemical Engineers 32(Supplement No 1): S49
- S53.

Danckwerts, P. V. and Sharma, M. M. (1966). "The absorption of carbon dioxide into solutions of alkalis and
amines (with some notes on hydrogen sulfide and carbonyl sulfide)." The Chemical Engineer (202): 244 -
280.

de Hemptinne, J.-C., Dhima, A. and Shakir, S. (2000). The Henry constant for 20 hydrocarbons, CO2, H2S in water
as a function of pressure and temperature. 14th Symposium on Thermophysical Properties, Boulder,
Colorado, USA.

de Leye, L. and Froment, G. F. (1986). "Rigorous simulation and design of columns for gas absorption and
chemical reaction - I." Computers and Chemical Engineering 10(5): 493 - 504.

Dean, D. E. and Stiel, L. I. (1965). "The viscosity of nonpolar gas mixtures at moderate and high pressure." AIChE
Journal 11(3): 526 - 532.

Drummond, S. E. (1981). Boiling and mixing of hydrothermal fluids: chemical effects on mineral precipitation. PhD
Thesis. University Park, Pennsylvania State University

Ebrahimi, S., Picioreanu, C., Kleerebezem, R., Heijnen, J. J. and van Loosdrecht, M. C. M. (2003). "Rate-based
modelling of SO2 absorption into aqueous NaHCO3/Na2CO3 solutions accompanied by the desorption of
CO2." Chemical Engineering Science 58: 3589 - 3600.

Ely, J. F. and Hanley, H. J. M. (1981). "Prediction of transport properties. 1. Viscosity of fluids and mixtures."
Industrial and Engineering Chemistry Fundamentals 20(4): 323 - 332.

206
References

Ely, J. F. and Hanley, H. J. M. (1983). "Prediction of transport properties. 2. Thermal conductivity of pure fluids
and mixtures." Industrial and Engineering Chemistry Fundamentals 22(1): 90 - 97.

Fernández-Prini, R., Alvarez, J. L. and Harvey, A. H. (2003). "Henry's constants and vapour-liquid distribution
constants for gaseous solutes in H2O and D2O at high temperatures." Journal of Physical and Chemical
Reference Data 32(2): 903 - 916.

Ferrell, R. T. and Himmelblau, D. M. (1967). "Diffusion coefficients of nitrogen and oxygen in water." Journal of
Chemical and Engineering Data 12(1): 111 - 115.

Field, J. H., Benson, H. E., Johnson, G. E., Tosh, J. S. and Forney, A. J. (1962). Pilot-Plant Studies of the Hot-
Carbonate Process for Removing Carbon Dioxide and Hydrogen Sulfide. Bulletin 597. Washington,
United States Department of the Interior, Bureau of Mines

Fortescue, T. R., Kershenbaum, L. S. and Ydstie, B. E. (1981). "Implementation of self-tuning regulators with
variable forgetting factors." Automatica 17(6): 831 - 835.

Fürst, W. and Renon, H. (1993). "Representation of excess properties of electrolyte solutions using a new
equation of state." The American Institute of Chemical Engineers Journal 39(2): 335 - 343.

Garcia, C. E., Prett, D. M. and Morari, M. (1989). "Model predictive control: Theory and practice - a survey."
Automatica 25(3): 335 - 348.

Golub, G. H. and Kahan, W. (1965). "Calculating the singular values and pseudo-inverse of a matrix." Journal of
the Society for Industrial and Applied Mathematics: Series B, Numerical Analysis 2(2): 205 - 224.

Gonçalves, F. A. and Kestin, J. (1981). "The viscosity of Na2CO3 and K2CO3 aqueous solutions in the range 20 -
60°C." International Journal of Thermophysics 2(4): 315 - 322.

Graboski, M. S. and Daubert, T. E. (1978). "A modified Soave equation of state for phase equilibrium calculations.
1. Hydrocarbon systems." Industrial and Engineering Chemistry Process Design and Development 17(4):
443 - 448.

Grosdidier, P., Morari, M. and Holt, B. R. (1985). "Closed-loop properties from steady-state gain information."
Industrial and Engineering Chemistry Fundamentals 24: 221 - 235.

Grover, B. S. (1987). Process for the removal of acid gases from gas mixtures. Patent No. 4,702,898. United
States.

Haar, L., Gallagher, J. S. and Kell, G. S. (1984). NBS/NRC Steam Tables: Thermodynamic and Transport
Properties and Computer Programs for Vapour and Liquid States of Water in SI Units. Washington,
Hemisphere Publishing Corporation.

Hankinson, R. W. and Thomson, G. H. (1979). "A new correlation for saturated densities of liquids and their
mixtures." AIChE Journal 25(4): 653 - 663.

Harned, H. S. and Owen, B. B. (1958). The Physical Chemistry of Electrolytic Solutions. 3rd ed. New York,
Reinhold Publishing Corporation.

Hayduk, W. and Laudie, H. (1974). "Prediction of diffusion coefficients for nonelectrolytes in dilute aqueous
solutions." AIChE Journal 20(3): 611 - 615.

Higbie, R. (1935). "The rate of absorption of a pure gas into a still liquid during short periods of exposure."
Transactions of the American Institute of Chemical Engineers 31: 365 - 383.

Hilliard, M. D. (2005). Thermodynamics of Aqueous Piperazine/Potassium Carbonate/Carbon Dioxide


Characterised by the Electrolyte NRTL Model within Aspen Plus. M.S. Thesis. Austin, The University of
Texas

Horvath, A. L. (1985). Handbook of Aqueous Electrolyte Solutions: Physical Properties, Estimation and Correlation
Methods. Chichester, Ellis Horwood Limited.

207
References

Huang, H.-P., Jeng, J.-C., Chiang, C.-H. and Pan, W. (2003). "A direct method for multi-loop PI/PID controller
design." Journal of Process Control 13: 769 - 786.

Hudson Products Corporation (2002). Hudson Products Corporation website. http://www.hudsonproducts.com.


Accessed: 10 October 2005.

Hyprotech Ltd (2001a). HYSYS Simulation Basis. Calgary, Hyprotech Ltd.

Hyprotech Ltd (2001b). HYSYS Dynamic Modelling. Calgary, Hyprotech Ltd.

Hyprotech Ltd (2001c). HYSYS Steady State Modelling. Calgary, Hyprotech Ltd.

Joback, K. G. (1984). A Unified Approach to Physical Property Estimation Using Multivariate Statistical
Techniques. M.S. Thesis. Cambridge, Massachusetts Institute of Technology

Jones, G. and Dole, M. (1929). "The viscosity of aqueous solutions of strong electrolytes with special reference to
barium chloride." Journal of the American Chemical Society 51(10): 2950 - 2964.

Joseph, B. (1999). A tutorial on inferential control and its applications. 1999 American Control Conference, San
Diego, USA.

Joshi, S. V., Astarita, G. and Savage, D. W. (1981). "Prediction of pilot plant performance for a chemical gas
absorption process." The American Institute of Chemical Engineers Symposium Series 77(202): 63 - 77.

Kenig, E. Y., Schneider, R. and Górak, A. (1999). "Rigorous dynamic modelling of complex reactive absorption
processes." Chemical Engineering Science 54: 5195 - 5203.

Kenig, E. Y., Schneider, R. and Górak, A. (2001). "Reactive absorption: Optimal process design via optimal
modelling." Chemical Engineering Science 56: 343 - 350.

Kent, R. L. and Eisenberg, B. (1976). "Better Data for Amine Treating." Hydrocarbon Processing 55(2): 87 - 90.

Kershenbaum, L. S. (2000). "Experimental testing of advanced algorithms for process control." Chemical
Engineering Research and Design : Transactions of the Institution of Chemical Engineers 78(A4): 509 -
521.

Kershenbaum, L. S. and Pérez-Correa, J. R. (1989). "An extended horizon feedback/feedforward self-tuning


controller." AIChE Journal 35(11): 1835 - 1844.

Kesler, M. G. and Lee, B. I. (1976). "Improve prediction of enthalpy of fractions." Hydrocarbon Processing 55(3):
153 - 158.

Kister, H. Z. (1992). Distillation Design. Boston, McGraw-Hill.

Koch-Glitsch (2003a). Intalox Packed Tower Systems: Metal Random Packing. Bulletin KGMRP-1 2M0303B.

Koch-Glitsch (2003b). Intalox Packed Tower Systems: IMTP® High Performance Packing. Bulletin KGIMTP-2
2MI303E.

Kohl, A. L. and Nielson, R. B. (1997). Alkaline Salt Solutions for Acid Gas Removal. Gas Purification. Houston,
Gulf Publishing Company: 330 - 414.

Kolev, N. (1976). "Operation of randomly packed columns." Chemie Ingenieur Technik 48(12): 1105 - 1112.

Krishnamurthy, R. and Taylor, R. (1985). "Simulation of packed distillation and absorption columns." Industrial and
Engineering Chemistry Process Design and Development 24(3): 513 - 524.

Kucka, L., Kenig, E. Y. and Górak, A. (2002). "Kinetics of the gas-liquid reaction between carbon dioxide and
hydroxide ions." Industrial and Engineering Chemistry Research 41: 5952 - 5957.

Kucka, L., Müller, I., Kenig, E. Y. and Górak, A. (2003). "On the modelling and simulation of sour gas absorption
by aqueous amine solutions." Chemical Engineering Science 58: 3571 - 3578.

208
References

Lee, J. H. and Cooley, B. (1997). Recent advances in model predictive control and other related areas. 5th
International Conference on Chemical Process Control, AICHE and CACHE, Tahoe City, USA.

Letsou, A. and Stiel, L. I. (1973). "Viscosity of saturated nonpolar liquids at elevated pressures." AIChE Journal
19(2): 153 - 158.

Lewin, D. R. (1996). "A simple tool for disturbance resiliency diagnosis and feedforward control design."
Computers and Chemical Engineering 20(1): 13 - 25.

Lewis, W. K. and Whitman, W. G. (1924). "Principles of gas absorption." Industrial and Engineering Chemistry
16(1215 - 1220).

Loehe, J. R. and Donohue, M. D. (1997). "Recent advances in modelling thermodynamic properties of aqueous
strong electrolyte systems." The American Institute of Chemical Engineers Journal 43(1): 180 - 195.

Luyben, W. L. (1986). "Simple method for tuning SISO controllers in multivariable systems." Industrial and
Engineering Chemistry: Process Design and Development 25: 654 - 660.

Luyben, W. L. (1996). "Tuning proportional-integral-derivative controllers for integrator/dead time processes."


Industrial and Engineering Chemistry Research 35(3480 - 3483).

Luyben, W. L. and Luyben, M. L. (1997). Essentials of Process Control. New York, McGraw-Hill.

Marini, L., Clement, K., Georgakis, C. and Suenson, M. M. (1985). "Experimental and theoretical investigation of
an absorber-stripper pilot plant under nonequilibrium conditions." Industrial and Engineering Chemistry
Fundamentals 24(3): 296 - 301.

Mason, E. A. and Saxena, S. C. (1958). "Approximate formula for the thermal conductivity of gas mixtures."
Physics of Fluids 1: 361 - 369.

Mayer, J. (2002). Experimentelle und theoretische Untersuchungen zur chemischen Absorption am Biespiel der
Koksofengasreinigung. Berlin, Logos Verlag.

Misic, D. and Thodos, G. (1961). "The thermal conductivity of hydrocarbon gases at normal pressures." AIChE
Journal 7(2): 264 - 267.

Misic, D. and Thodos, G. (1963). "Atmospheric thermal conductivities for gases of simple molecular structure."
Journal of Chemical and Engineering Data 8(4): 540 - 544.

Mock, B., Evans, L. B. and Chen, C. C. (1986). "Thermodynamic representation of phase equilibria of mixed-
solvent electrolyte systems." AIChE Journal 32(10): 1655 - 1664.

Monica, T. J., Yu, C. C. and Luyben, W. L. (1988). "Improved multiloop single-input/single-output (SISO)
controllers for multivariable processes." Industrial and Engineering Chemistry Research 27: 969 - 973.

Morari, M. (1983). "Design of resilient processing plants. III. A general framework for the assessment of dynamic
resilience." Chemical Engineering Science 38: 1881 - 1891.

Murphree, E. V. (1925). "Rectifying column calculations with particular reference to N component mixtures."
Industrial and Engineering Chemistry 17(7): 747 - 750.

Niederlinski, A. (1971). "A heuristic approach to the design of linear multivariable interacting control systems."
Automatica 7(6): 691 - 701.

Noeres, C., Kenig, E. Y. and Górak, A. (2003). "Modelling of reactive separation processes: reactive absorption
and reactive distillation." Chemical Engineering and Processing 42: 157 - 178.

O'Dwyer, A. (2006). Handbook of PI and PID Controller Tuning Rules. 2nd ed. London, Imperial College Press.

Onda, K., Kobayashi, T., Fujine, M. and Takahashi, M. (1971). "Concentration profile of sodium ion in gas
absorption." Chemical Engineering Science 26: 2027 - 2035.

209
References

Onda, K., Sada, E. and Takeuchi, H. (1968b). "Gas absorption with chemical reaction in packed columns." Journal
of Chemical Engineering of Japan 1(1): 62 - 66.

Onda, K., Takeuchi, H. and Okumoto, Y. (1968a). "Mass transfer coefficients between gas and liquid phases in
packed columns." Journal of Chemical Engineering of Japan 1(1): 56 - 62.

Onsager, L. and Samaras, N. N. T. (1934). "The surface tension of Debye-Hückel electrolytes." The Journal of
Chemical Physics 2(8): 528 - 536.

Ooi, S. M., Colby, C. B., Humphris, K., King, K. D. and O'Neill, B. K. (2005). The evaluation of different electrolyte
thermodynamic models for simulating the hot potassium carbonate process. 7th World Congress of
Chemical Engineering, Glasgow, Scotland.

Panda, R. C., Yu, C.-C. and Huang, H.-P. (2004). "PID tuning rules for SOPDT systems: Review and some new
results." ISA Transactions 43: 283 - 295.

Park, K. Y. and Edgar, T. F. (1984). "Simulation of the hot carbonate process for removal of CO2 and H2S from
medium Btu gas." Energy Progress 4(3): 174 - 180.

Peng, D. Y. and Robinson, D. B. (1976). "A two constant equation of state." Industrial and Engineering Chemistry
Fundamentals 15: 59 - 64.

Perry, R. H. and Green, D. W. (1997). Perry's Chemical Engineers' Handbook. 7th ed. New York, McGraw-Hill.

Pitzer, K. S. (1973). "Thermodynamics of electrolytes. I. Theoretical basis and general equations." The Journal of
Physical Chemistry 77(2): 268 - 277.

Pitzer, K. S. (1980). "Electrolytes. From dilute solutions to fused salts." Journal of American Chemical Society
102(9): 2902 - 2906.

Pitzer, K. S., Lippmann, D. Z., Curl, R. F., Jr., Huggins, C. M. and Petersen, D. E. (1955). "Volumetric and
thermodynamic properties of fluids. II. Compressibility factor, vapour pressure and enthalpy of
vaporization." Journal of American Chemical Society 77(13): 3433 - 3440.

Pohorecki, R. and Kucharski, E. (1991). "Desorption with chemical reaction in the system CO2-aqueous solution of
potassium carbonate." The Chemical Engineering Journal 46: 1 - 7.

Pohorecki, R. and Moniuk, W. (1988). "Kinetics of reactions between carbon dioxide and hydroxyl ions in aqueous
electrolyte solutions." Chemical Engineering Science 43(7): 1677 - 1684.

Poling, B. E., Prausnitz, J. M. and O'Connell, J. P. (2001). The Properties of Gases and Liquids. 5th ed. New York,
McGraw-Hill.

Popolizio, B. (2005). Personal communication. 27 October.

Puranik, S. S. and Vogelpohl, A. (1974). "Effective interfacial area in irrigated packed columns." Chemical
Engineering Science 29(2): 501 - 507.

Qin, S. J. and Badgwell, T. A. (2003). "A survey of industrial model predictive control technology." Control
Engineering Practice 11: 733 - 764.

Rahimpour, M. R. and Kashkooli, A. Z. (2004). "Enhanced carbon dioxide removal by promoted hot potassium
carbonate in a split-flow absorber." Chemical Engineering and Processing 43: 857 - 865.

Rao, D. P. (1990). "Design of a packed column for absorption of carbon dioxide in DEA promoted hot K2CO3
solution." Gas Separation and Purification 4: 58 - 61.

Rao, D. P. (1991). "Design of packed towers for absorption and desorption of carbon dioxide using hot promoted
K2CO3 solution." Gas Separation and Purification 5(9): 177 - 180.

Ratcliff, G. A. and Holdcroft, J. G. (1963). "Diffusivities of gases in aqueous electrolyte solutions." Transactions of
the Institution of Chemical Engineers 41: 315 - 319.

210
References

Redlich, O. and Kwong, J. N. S. (1949). "On the thermodynamics of solutions. V. An Equation of State. Fugacities
of gaseous solutions." Chemical Reviews 44: 233 - 245.

Reid, R. C., Prausnitz, J. M. and Sherwood, T. K. (1977). The Properties of Gases and Liquids. 3rd ed. New York,
McGraw-Hill.

Renon, H. (1996). "Models for excess properties of electrolyte solutions: molecular bases and classification, needs
and trends for new developments." Fluid Phase Equilibria 116: 217 - 224.

Renon, H. and Prausnitz, J. M. (1968). "Local compositions in thermodynamic excess functions for liquid
mixtures." The American Institute of Chemical Engineers Journal 14: 135 - 144.

Riedel, L. (1951). "The thermal conductivity of aqueous solutions of strong electrolytes." Chemie Ingenieur
Technik 23: 59 - 64.

Riedel, L. (1954). "A new universal vapour-pressure equation. I. The extension of the theorem of corresponding
states." Chemie Ingenieur Technik 26: 83 - 89.

Robinson, R. A. and Stokes, R. H. (1965). Electrolyte Solutions. 2nd (Revised) ed. London, Butterworths.

Rosenbrock, H. H. (1974). Computer Aided Control System Design. London, Academic Press.

Rowley, R. L., Wilding, W. V., Oscarson, J. L. and Yang, Y. (1998). DIPPR® Project 801 Database of Evaluated
Process Design Data. http://dippr.byu.edu. Brigham Young University. Accessed: 2005.

Sander, B., Fredenslund, A. and Rasmussen, P. (1986a). "Calculation of vapour-liquid equilibria in mixed
solvent/salt systems using an extended UNIQUAC equation." Chemical Engineering Science 41(5): 1171
- 1183.

Sander, B., Rasmussen, P. and Fredenslund, A. (1986b). "Calculation of vapour-liquid equilibria in nitric acid-
water-nitrate salt systems using an extended UNIQUAC equation." Chemical Engineering Science 41(5):
1185 - 1195.

Santos Ltd (1997). South Australian Operations: Moomba Processing Facilities, Santos Ltd.

Santos Ltd (1998). Section 4.0: CO2 Removal (Benfield) Plants. Moomba Gas Plant Technical Manual, Santos
Ltd: 4.10.

Sanyal, D., Vasishtha, N. and Saraf, D. N. (1988). "Modelling of carbon dioxide absorber using hot carbonate
process." Industrial and Engineering Chemistry Research 27(11): 2149 - 2156.

Schneider, R. and Górak, A. (2001). "Model optimization for the dynamic simulation of reactive absorption
processes." Chemical Engineering and Technology 24(10): 979 - 989.

Seader, J. D. and Henley, E. J. (1998). Separation Process Principles. New York, John Wiley and Sons.

Seborg, D. E., Edgar, T. F. and Mellichamp, D. A. (1989). Process Dynamics and Control. New York, John Wiley
and Sons.

Seborg, D. E., Edgar, T. F. and Shah, S. L. (1986). "Adaptive control strategies for process control: A survey."
AIChE Journal 32(6): 881 - 913.

Sherwood, T. K. and Holloway, F. A. L. (1940a). "Performance of packed towers - Experimental studies of


absorption and desorption." Transactions of the American Institute of Chemical Engineers 36(1): 21 -37.

Sherwood, T. K. and Holloway, F. A. L. (1940b). "Performance of packed towers - Liquid-film data for several
packings." Transactions of the American Institute of Chemical Engineers 36(1-2): 39 - 90.

Shinskey, F. G. (1979). Process Control Systems : Application, Design, Adjustment. 2nd ed. New York, McGraw-
Hill.

Shulman, H. L., Ullrich, C. F., Proulx, A. Z. and Zimmerman, J. O. (1955). "Performance of packed columns. 1.
Total, static, and operating holdups." AIChE Journal 1(2): 247 - 253.

211
References

Skogestad, S. (2003). "Simple analytic rules for model reduction and PID controller tuning." Journal of Process
Control 13: 291 - 309.

Skogestad, S., Jacobsen, E. W. and Morari, M. (1990). "Inadequacy of steady-state analysis for feedback control:
distillate-bottom control of distillation columns." Industrial and Engineering Chemistry Research 29(12):
2339 - 2346.

Skogestad, S. and Morari, M. (1987). "Effect of disturbance direction on closed-loop performance." Industrial and
Engineering Chemistry Research 26(10): 2029 - 2035.

Skogestad, S. and Postlethwaite, I. (1996). Multivariable Feedback Control: Analysis and Design. Chichester,
John Wiley and Sons.

Soave, G. (1972). "Equilibrium constants from a modified Redlich-Kwong equation of state." Chemical Engineering
Science 27: 1197 - 1203.

Spencer, C. F. and Danner, R. P. (1972). "Improved equation for prediction of saturated liquid density." Journal of
Chemical and Engineering Data 17(2): 236 - 241.

Stanley, G., Marino-Galarraga, M. and McAvoy, T. J. (1985). "Shortcut operability analysis. 1. The relative
disturbance gain." Industrial and Engineering Chemistry Process Design and Development 24: 1181 -
1188.

Staton, J. S. (1985). Performance and Modelling of a Hot Potassium Carbonate Acid Gas Removal System in
Treating Coal Gas. PhD Thesis. Raleigh, North Carolina State University

Stephanopoulos, G. (1984). Chemical Process Control: An Introduction to Theory and Practice. Englewood Cliffs,
Prentice-Hall.

Stephanopoulos, G. and Han, C. (1996). "Intelligent systems in process engineering: A review." Computers and
Chemical Engineering 20(6/7): 743 - 791.

Stichlmair, J., Bravo, J. L. and Fair, J. R. (1989). "General model for prediction of pressure drop and capacity of
countercurrent gas/liquid packed columns." Gas Separation and Purification 3: 19 - 28.

Stichlmair, J. and Fair, J. R. (1998). Distillation: Principles and Practice. New York, Wiley-VCH.

Stiel, L. I. and Thodos, G. (1964). "The thermal conductivity of nonpolar substances in the dense gaseous and
liquid regions." AIChE Journal 10(1): 26 - 30.

Stone, D. (2005). Personal communication. 23 January.

Stryjek, R. and Vera, J. H. (1986a). "PRSV: An improved Peng-Robinson equation of state for pure compounds
and mixtures." Canadian Journal of Chemical Engineering 64(2): 323 - 333.

Stryjek, R. and Vera, J. H. (1986b). "PRSV: An improved Peng-Robinson equation of state with new mixing rules
for strongly non-ideal mixtures." Canadian Journal of Chemical Engineering 64(2): 334 - 340.

Stumm, W. and Morgan, J. J. (1996). Aquatic Chemistry: Chemical Equilibria and Rates in Natural Waters. 3rd ed.
New York, John Wiley and Sons.

Suenson, M. M., Georgakis, C. and Evans, L. B. (1985). "Steady-state and dynamic modelling of a gas absorber-
stripped system." Industrial and Engineering Chemistry Fundamentals 24(3): 288 - 295.

Sulzer Chemtech (2006). Metal Random Packing: Superior random packings combined with innovative mass
transfer technology. Brochure 22.64.06.40 - VI.06 - 50.

Svrcek, W. Y., Mahoney, D. P. and Young, B. R. (2006). A Real-Time Approach to Process Control. 2nd ed.
Chichester, John Wiley and Sons.

Tamimi, A., Rinker, E. B. and Sandall, O. C. (1994). "Diffusion coefficients for hydrogen sulfide, carbon dioxide,
and nitrous oxide in water over the temperature range 293 - 368 K." Journal of Chemical and Engineering
Data 39(2): 330 - 332.

212
References

Taylor, R. and Krishna, R. (1993). Multicomponent Mass Transfer. New York, John Wiley and Sons.

Thiele, R. (2007). Selektive Absorption innerhalb der Kokereigasreinigung. PhD Thesis. Berlin, Technischen
Universität Berlin

Thiele, R., Repke, J.-U., Diekjakobs, B., Thielert, H. and Wozny, G. (2005a). A general rate-based model for
industrial reactive absorption and desorption processes in sour gas treatment. 2005 AIChE Spring
National Meeting, Atlanta, USA.

Thiele, R., Repke, J.-U., Diekjakobs, B., Thielert, H. and Wozny, G. (2005b). Rate-based modelling of reactive
absorption and desorption processes in sour gas treatment. 7th World Congress of Chemical
Engineering, Glasgow, Scotland.

Thomsen, K. and Rasmussen, P. (1999). "Modelling of vapour-liquid-solid equilibrium in gas-aqueous electrolyte


systems." Chemical Engineering Science 54: 1787 - 1802.

Todinca, T. (1994). "Mathematical model for the gas desorber used for carbon dioxide desorption from DEA-
promoted hot potassium carbonate." Chemical Bulletin of the Technical University of Timisoara 39(53):
127 - 134.

Todinca, T. (1995). "Mathematical modelling of absorption-desorption processes: Carbon dioxide removing plants
using diethanolamine promoted hot potassium carbonate." Chemical Bulletin of the Technical University
of Timisoara 40(54): 11 - 20.

Todinca, T., Moszkowicz, P., Târnoveanu, M. and Benadda, B. (1997). "Absorption of carbon dioxide in
diethanolamine promoted hot potassium carbonate processes." International Chemical Engineering
Symposium Series 142: 429 - 437.

Tosh, J. S., Field, J. H., Benson, H. E. and Anderson, R. B. (1960). Equilibrium Pressures of Hydrogen Sulfide and
Carbon Dioxide over Solutions of Potassium Carbonate. Report of Investigations 5622. Washington,
United States Department of the Interior, Bureau of Mines

Tosh, J. S., Field, J. H., Benson, H. E. and Haynes, B. S. (1959). Equilibrium Study of the System Potassium
Carbonate, Potasstium Bicarbonate, Carbon Dioxide, and Water. Report of Investigations 5484.
Washington, United States Department of the Interior, Bureau of Mines

Twu, C. H. (1985). "Internally consistent correlation for predicting liquid viscosities of petroleum fractions."
Industrial and Engineering Chemistry Process Design and Development 24(4): 1287 -1293.

Tyreus, B. D. and Luyben, W. L. (1992). "Tuning PI controllers for integrator/dead time processes." Industrial and
Engineering Chemistry Research 31: 2625 - 2628.

UOP Gas Processing (1998). UOP Benfield ACT-1 Activator Operating Manual. Des Plaines, UOP

UOP Gas Processing (2000). Benfield Process Technical Sheet. Des Plaines, UOP

van der Lee, J. H., Swallow, J. A. and Young, B. R. (2003). "CO2 loading measurement and control in an amine
absorption/stripping pilot plant." Technical Papers of ISA 444: 59 - 65.

van der Lee, J. H., Young, B. R. and Svrcek, W. Y. (2001). Model predictive control of a pilot amine
absorption/stripping plant. 51st Canadian Chemical Engineering Conference, Halifax, Canada.

van der Lee, J. H., Young, B. R., Swallow, J. A. and Svrcek, W. Y. (2004). Monitoring CO2 loading of rich and lean
streams in an amine absorption/stripping pilot plant. 49th Annual ISA Analysis Division Symposium,
Cincinnati, USA.

van Krevelen, D. W. and Hoftijzer, P. J. (1948). "Kinetics of simultaneous absorption and chemical reaction."
Chemical Engineering Progress 44: 529 - 536.

van Swaaij, W. P. M. and Versteeg, G. F. (1992). "Mass transfer accompanied with complex reversible chemical
reactions in gas-liquid systems: An overview." Chemical Engineering Science 47(13 - 14): 3181 - 3195.

213
References

Wade, H. L. (2004). Basic and Advanced Regulatory Control: System Design and Application. 2nd ed. Research
Triangle Park, The Instrumentation, Systems and Automation Society.

Wang, G. Q., Yuan, X. G. and Yu, K. T. (2005). "Review of mass-transfer correlations for packed columns."
Industrial and Engineering Chemistry Research 44: 8715 - 8729.

Wang, J. H. (1965). "Self-diffusion coefficients of water." Journal of Physical Chemistry 69(12): 4412.

Wilke, C. R. (1950). "A viscosity equation for gas mixtures." The Journal of Chemical Physics 18(4): 517 - 519.

Willcocks, L. (2008). Personal communication. 26 May.

Wilson, G. M. (1980). A New Correlation of Ammonia, Carbon Dioxide, and Hydrogen Sulfide Volatility Data from
Aqueous Sour Water Systems. Washington DC, American Petroleum Institute

Witcher, M. and McAvoy, T. J. (1977). "Interacting control systems: Steady state and dynamic measurement of
interaction." ISA Transactions 16: 35 - 41.

Witherspoon, P. A. and Saraf, D. N. (1965). "Diffusion of methane, ethane, propane and n-butane in water from 25
to 43°." Journal of Physical Chemistry 69(11): 3752 - 3755.

Wittenmark, B. (1997). A survey of adaptive control applications. Dynamic Modelling Control Applications for
Industry Workshop, IEEE Industry Applications Society, Vancouver, Canada.

Ydstie, B. E., Kershenbaum, L. S. and Sargent, R. W. H. (1985). "Theory and application of an extended horizon
self-tuning controller." AIChE Journal 31(11): 1771 - 1780.

Yen, L. C. and Woods, S. S. (1966). "A generalised pressure correlation for gas-solvent-reservoir oil systems."
AIChE Journal 12(1): 95 - 99.

Yih, S.-M. and Lai, H.-C. (1987). "Simultaneous absorption of carbon dioxide and hydrogen sulfide in hot
carbonate solutions in a packed absorber-stripper unit." Chemical Engineering Communications 51(1-6):
277 - 290.

Yu, C. C. and Luyben, W. L. (1986). "Design of multiloop SISO controllers in multivariable processes." Industrial
and Engineering Chemistry: Process Design and Development 25: 498 - 503.

Zemaitis, J. F., Jr, Clark, D. M., Rafal, M. and Scrivner, N. C. (1986). Handbook of Aqueous Electrolyte
Thermodynamics: Theory and Application. New York, Design Institute for Physical Property Data.

Ziegler, J. G. and Nichols, N. B. (1942). "Optimum settings for automatic controllers." Transactions of the ASME
64: 759 - 768.

Ziegler, J. G. and Nichols, N. B. (1943). "Process lags in automatic control circuits." Transactions of the ASME 65:
433 - 444.

214
Volume of Appendices

for

Development and Demonstration of a New Non-


Equilibrium Rate-Based Process Model for the
Hot Potassium Carbonate Process

by

Su Ming Pamela Ooi

Thesis submitted for the degree of


Doctor of Philosophy

in

The University of Adelaide


School of Chemical Engineering
July 2008
Table of Contents

Table of Contents
Appendix A: Thermodynamic Model Equations ….......... A-1
A.1 Reference States ….......... A-2
A.2 The Electrolyte NRTL Model ….......... A-3
A.3 Cubic Equations of State ….......... A-8
Appendix B: Property Models for Aspen Custom Modeler® ….......... A-10
B.1 Thermodynamic Property Models ….......... A-11
B.2 Physical and Transport Property Models ….......... A-17
Appendix C: Electrolyte NRTL Adjustable Parameters ….......... A-38
C.1 Parameter Values ….......... A-39
C.2 Data Regression Procedure ….......... A-41
C.3 Data Regression Results ….......... A-44
Appendix D: Aspen Custom Modeler® Simulation Results ….......... A-49
D.1 The Different Modelling Approaches ….......... A-50
D.2 Model Adjustments ….......... A-52
D.3 CO2 Train Model Validation ….......... A-56
Appendix E: Hypothetical K2CO3* HYSYS® Component Properties ….......... A-62
E.1 Base Properties ….......... A-63
E.2 Additional Point Properties ….......... A-66
E.3 Temperature Dependent Properties ….......... A-67
Appendix F: Property Models for HYSYS® ….......... A-70
F.1 Thermodynamic Property Models ….......... A-71
F.2 Physical and Transport Property Models ….......... A-72
Appendix G: Enhanced PR Binary Interaction Parameters ….......... A-84
G.1 Data Regression Procedure ….......... A-85
G.2 Data Regression Results ….......... A-87
Appendix H: HYSYS® Simulation Results ….......... A-88
H.1 Preliminary Column Model Simulations ….......... A-89
H.2 Steady-State CO2 Train Models ….......... A-95
H.3 CO2 Train Model Validation ….......... A-101
Appendix I: Process Control Studies of the CO2 Trains ….......... A-107
I.1 Selection of Diagonal Control Structure ….......... A-108
I.2 Selection of Diagonal Control Structure Configuration ….......... A-111
I.3 Analysis of Diagonal Control Structure Performance ….......... A-114
I.4 BLT Tuning ….......... A-115
I.5 Diagonal Control Structure Dynamic Simulations ….......... A-117

A-i
List of Figures

List of Figures
Figure B.1.1: Comparison between the predicted and experimental solution
heat capacities. ….......... A-16
Figure B.2.1: Comparison between the predicted and experimental solution
mass densities. ….......... A-20
Figure B.2.2: Comparison between the predicted and experimental solution
viscosities. ….......... A-25
Figure B.2.3: Comparison between the predicted and experimental solution
surface tensions. ….......... A-28
Figure B.2.4: Comparison between the predicted and experimental solution
thermal conductivities. ….......... A-33
Figure D.1.1: Results of the different modelling approaches for CO2 trains #2
to #4. ….......... A-50
Figure D.1.2: Results of the different modelling approaches for CO2 trains #5
and #6. ….......... A-51
Figure D.2.1: Temperature profiles for CO2 trains #2 to #4. ….......... A-52
Figure D.2.2: Temperature profiles for CO2 trains #5 and #6. ….......... A-53
Figure D.2.3: Effect of the effective interfacial area adjustment factor on the
absorber CO2 and H2S vapour phase profiles. ….......... A-55
Figure D.3.1: CO2 and H2S vapour and liquid phase column profiles for the
first set of plant data. ….......... A-56
Figure D.3.2: CO2 and H2S vapour and liquid phase column profiles for the
first set of plant data. ….......... A-57
Figure D.3.3: CO2 and H2S vapour and liquid phase column profiles for the
second set of plant data. ….......... A-58
Figure D.3.4: CO2 and H2S vapour and liquid phase column profiles for the
second set of plant data. ….......... A-59
Figure D.3.5: Column temperature profiles for the first set of plant data. ….......... A-60
Figure D.3.6: Column temperature profiles for the second set of plant data. ….......... A-61
Figure F.2.1: Comparison between the predicted and experimental solution
mass densities. ….......... A-74
Figure F.2.2: Comparison between the predicted and experimental solution
viscosities. ….......... A-77
Figure F.2.3: Comparison between the solution surface tensions predicted by
the tabular model and the empirical correlation. ….......... A-79
Figure F.2.4: Comparison between the solution thermal conductivities
predicted by the tabular model and the empirical correlation. ….......... A-83
Figure H.1.1: Equilibrium stage simulation results for the absorber and
regenerator columns. ….......... A-89
Figure H.1.2: Equilibrium stage simulation results for the absorber and
regenerator columns. ….......... A-90
Figure H.1.3: Effect of the correlated overall stage efficiencies on the steady-
state absorber columns. ….......... A-91
Figure H.1.4: Effect of the correlated overall stage efficiencies on the steady-
state absorber columns. ….......... A-92

A-ii
List of Figures

Figure H.1.5: Effect of the correlated overall stage efficiencies on the steady-
state behaviour of the dynamic absorber columns. ….......... A-93
Figure H.1.6: Effect of the correlated overall stage efficiencies on the steady-
state behaviour of the dynamic absorber columns. ….......... A-94
Figure H.2.1: Process flow diagram for the steady-state model of CO2 train #2. ….......... A-96
Figure H.2.2: Process flow diagram for the steady-state model of CO2 train #3. ….......... A-97
Figure H.2.3: Process flow diagram for the steady-state model of CO2 train #4. ….......... A-98
Figure H.2.4: Process flow diagram for the steady-state model of CO2 train #5. ….......... A-99
Figure H.2.5: Process flow diagram for the steady-state model of CO2 train #6. ….......... A-100
Figure H.3.1: CO2 and H2S vapour and liquid phase column profiles for the
first set of plant data. ….......... A-101
Figure H.3.2: CO2 and H2S vapour and liquid phase column profiles for the
first set of plant data. ….......... A-102
Figure H.3.3: CO2 and H2S vapour and liquid phase column profiles for the
second set of plant data. ….......... A-103
Figure H.3.4: CO2 and H2S vapour and liquid phase column profiles for the
second set of plant data. ….......... A-104
Figure H.3.5: Column temperature profiles for the first set of plant data. ….......... A-105
Figure H.3.6: Column temperature profiles for the second set of plant data. ….......... A-106
Figure I.5.1: CO2 train #1 closed-loop step response curves at the medium
gas throughput conditions. ….......... A-121
Figure I.5.2: CO2 train #7 closed-loop step response curves at the medium
gas throughput conditions. ….......... A-122
Figure I.5.3: CO2 train #1 closed-loop step response curves at the low gas
throughput conditions. ….......... A-123
Figure I.5.4: CO2 train #7 closed-loop step response curves at the low gas
throughput conditions. ….......... A-124

A-iii
List of Tables

List of Tables
Table B.1.1: Component critical properties (Poling et al., 2001). ….......... A-11
Table B.1.2: Ideal gas heat capacity coefficients and enthalpies of formation
(Poling et al., 2001). ….......... A-12
Table B.1.3: Ionic species thermodynamic properties (Zemaitis et al., 1986). ….......... A-15
Table B.1.4: Temperature dependence of Henry’s Law constants. ….......... A-15
Table B.1.5: Criss-Cobble entropy parameters (Criss and Cobble, 1964ab). ….......... A-16
Table B.1.6: Atmospheric solution heat capacity data. ….......... A-16
Table B.1.7: Parameter values for the Aspen Properties® heat capacity
polynomial. ….......... A-16
Table B.2.1: Antoine equation coefficients (Rowley et al., 1998). ….......... A-17
Table B.2.2: Parameter values for the modified Rackett equation (Spencer
and Danner, 1972). ….......... A-19
Table B.2.3: Atmospheric solution mass density data. ….......... A-19
Table B.2.4: Pair parameter values for the Clarke Aqueous Electrolyte
Volume model. ….......... A-19
Table B.2.5: Component characteristic volumes. ….......... A-21
Table B.2.6: Coefficients for the DIPPR vapour viscosity model (Rowley et al.,
1998). ….......... A-23
Table B.2.7: Coefficients for the Andrade liquid viscosity equation (Reid et al.,
1977). ….......... A-24
Table B.2.8: Atmospheric solution viscosity data. ….......... A-25
Table B.2.9: Parameter values for the Jones-Dole viscosity equation. ….......... A-25
Table B.2.10: Coefficients for the DIPPR surface tension equation (Rowley et
al., 1998). ….......... A-27
Table B.2.11: Atmospheric solution surface tension data. ….......... A-28
Table B.2.12: Surface tension correlation coefficients. ….......... A-28
Table B.2.13: Coefficients for the DIPPR vapour thermal conductivity model
(Rowley et al., 1998). ….......... A-30
Table B.2.14: Coefficients for the DIPPR liquid thermal conductivity equation
(Rowley et al., 1998). ….......... A-32
Table B.2.15: Atmospheric solution thermal conductivity data. ….......... A-32
Table B.2.16: Liquid phase thermal conductivity correlation coefficients. ….......... A-32
Table B.2.17: Normal boiling points and the corresponding liquid molar
volumes (Poling et al., 2001). ….......... A-35
Table B.2.18: Ionic conductivities at infinite dilution. ….......... A-37
Table B.2.19: Diffusivities in water at 25°C. ….......... A-37
Table C.1.1: The Electrolyte NRTL adjustable parameters used in this work. ….......... A-39
Table C.1.2: The Electrolyte NRTL adjustable parameters used in this work. ….......... A-40
Table C.3.1: Statistical results for the CO2-K2CO3-KHCO3-H2O system data
regression runs. ….......... A-45
Table C.3.2: Suitable parameter value sets for the CO2-K2CO3-KHCO3-H2O
system. ….......... A-46

A-iv
List of Tables

Table C.3.3: Statistical results for the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O


system data regression runs. ….......... A-47
Table C.3.4: Suitable parameter value sets for the CO2-H2S-K2CO3-KHCO3-
KHS-K2S-H2O system. ….......... A-48
Table D.2.1: Alternative mass transfer coefficient and effective interfacial area
correlations. ….......... A-54
Table E.3.1: Temperature dependent property correlation coefficients. ….......... A-67
Table F.2.1: Coefficient values for the HYSYS® liquid density tabular model. ….......... A-73
Table F.2.2: Coefficient values for the HYSYS® liquid viscosity tabular
model. ….......... A-77
Table F.2.3: Coefficient values for the HYSYS® liquid surface tension tabular
model. ….......... A-79
Table F.2.4: Coefficient values for the HYSYS® liquid thermal conductivity
tabular model. ….......... A-83
Table G.2.1: Statistical results for the CO2-K2CO3-H2O system data
regression runs. ….......... A-87
Table G.2.2: Statistical results for the CO2-H2S-K2CO3-H2O system data
regression runs. ….......... A-87
Table I.2.1: System poles and zeros for the SGC-RSF control structure. ….......... A-111

A-v
Appendix A: Thermodynamic Model Equations

APPENDIX A

THERMODYNAMIC MODEL
EQUATIONS

This appendix presents the model equations for the thermodynamic models utilised in this thesis. A
brief discussion of reference states is provided, followed by summaries of the Electrolyte NRTL model,
the Soave-Redlich-Kwong equation of state and the Peng-Robinson equation of state.

A-1
Appendix A: Thermodynamic Model Equations

A.1 Reference States


The difference between J and J* arises from the different reference states for solvent species (i.e.
water) and for solute species (e.g. CO2). In this work, the reference state for solvents is taken to be
the pure solvent species at the system temperature and pressure. This is known as the symmetric
convention and it defines the activity coefficient Js of a solvent species s such that:
J s o 1 as x s o 1 (A.1.1)

where xs is the liquid phase mole fraction of solvent species s.

For undissociated electrolytes, molecular solutes and ions, the reference state is the species at infinite
dilution in pure water at the system temperature and pressure. According to this un-symmetric
convention, which is denoted by an asterisk, the activity coefficient J *j of a solute species j is defined

so that:
J *j o 1 as x j o 0 (A.1.2)

where xj is the liquid phase mole fraction of solute species j.

The two conventions are related by:


ln J *j ln J j  ln J fj (A.1.3)

where J fj is the symmetrically normalised activity coefficient of species j at infinite dilution. J fj is

achieved in the limit where the composition of species j approaches zero and all the remaining solute
species k in the solution are set to zero:
J fj lim J j (A.1.4)
x j o0
xk z j 0

A-2
Appendix A: Thermodynamic Model Equations

A.2 The Electrolyte NRTL Model


A summary of the theory behind the Electrolyte Non-Random Two-Liquid (NRTL) model is presented
below. A more thorough discussion of the model theory and formulation can be found in the
publications by Chen and co-workers (1982), Chen and Evans (1986) and Mock and co-workers
(1986).

The Electrolyte NRTL model is an extension of the traditional NRTL model presented by Renon and
Prausnitz (1968) for calculating liquid activity coefficients. This model is based on two fundamental
assumptions (Chen et al., 1982):
1. The like-ion repulsion assumption: Due to the extremely large repulsive forces between ions
of like charge, the local composition of cations around cations is zero and the local
composition of anions around anions is zero.
2. The local electroneutrality assumption: The distribution of cations and anions around a central
molecular species is such that the net local ionic charge is zero.

The excess Gibbs energy expression proposed by Chen and co-workers (1982) involves two types of
contribution terms: one for the short-range local interactions around any central species and the other
for the long-range ion-ion interactions. The short-range local contribution term is represented by the
traditional NRTL model, while the long-range contribution term is represented by the un-symmetric
Pitzer-Debye-Hückel (PDH) model. An un-symmetric Born correction term is applied to account for
the effect of mixed solvents on the long-range contribution term. The expression for the un-symmetric
E*
excess Gibbs energy G is therefore:

G E* G E *,PDH G E *,Born G E*,NRTL


  (A.2.1)
R˜T R˜T R˜T R˜T

where R is the gas constant and T is the absolute temperature. The un-symmetric activity coefficient
is obtained from the partial differentiation of GE* with respect to the mole number nj of species j:

R ˜ T ˜ ln J *j «

ª w n T ˜ G E* º
»
(A.2.2)
¬« wn j ¼» T,P,n kz j

where nT is the total number of moles. Similarly, the symmetric activity coefficient Jj for a species j is
obtained from the partial differentiation of the symmetric excess Gibbs energy GE:

R ˜ T ˜ ln J j «

ª w n T ˜ GE º» (A.2.3)
«¬ wn j »¼ T,P,n
kzj

Substituting equation (A.2.2) into equation (A.2.1) gives the following expression for the un-symmetric
activity coefficient:
ln J *j ln J *,jPDH  ln J *,jBorn  ln J *,jNRTL (A.2.4)

A-3
Appendix A: Thermodynamic Model Equations

For the Pitzer-Debye-Hückel model, the un-symmetric excess Gibbs energy is expressed as (Pitzer,
1980):


0.5
G E*,PDH § 1000 · 4 ˜ A II x
R˜T
 ¦j
x j ˜ ¨¨
© MW solvent
¸
¸
¹
˜
p
˜ ln 1  p ˜ I x
0 .5
(A.2.5)

and the corresponding expression for the un-symmetric activity coefficient is:
ª2 ˜ z j2 z j ˜ Ix  2 ˜ Ix º

0. 5 2 0. 5 1 .5
§ 1000 ·
¨¨ ¸ ˜ AI ˜ « »
0. 5
ln J *,jPDH ¸ ˜ ln 1  p ˜ I x  (A.2.6)
© MW solvent ¹ « p 1  p ˜ Ix
0 .5 »
¬ ¼

where MW is the molecular weight, x is the liquid phase mole fraction, p is the closest approach
parameter, and z is the ionic charge. The Debye-Hückel parameter AI is defined as:
1 .5
1 § 2 ˜ S ˜ No ˜ Usolvent ·
0. 5
§ e2 ·
AI ˜¨ ¸ ˜¨ ¸ (A.2.7)
¨H ¸
3 © 1000 ¹ © solvent ˜ k ˜ T ¹

where No is Avogadro’s number,  is the density, e is the charge of an electron,  is the dielectric
constant, and k is the Boltzmann constant. The mole fraction ionic strength Ix is defined as:

Ix 0 .5 ˜ ¦xj
j ˜ zj
2
(A.2.8)

In mixed solvents, there is a change of reference state due to the difference in dielectric constants.
The un-symmetric Born correction term was introduced to maintain a reference state of infinite dilution
in water for the ionic species (Mock et al., 1986). The Born excess Gibbs energy is defined as
(Robinson and Stokes, 1965):

G E *,Born
e 2§ 1 1 ·
¦x j ˜ zj
2

˜¨ ¸˜ ˜ 10 2
j
 (A.2.9)
R˜T ¨
2 ˜ k ˜ T © H solvent H H2O ¸ rj
¹

and the corresponding un-symmetric activity coefficient is:

e2 § 1 1 · z j2
ln J *,jBorn ˜¨  ¸˜ ˜ 10 2 (A.2.10)
2˜k ˜T ¨ H solvent H H O ¸ rj
© 2 ¹

where r is the Born radius. This correction accounts for the difference in Gibbs energies between ionic
species in a mixed solvent and in water. Since the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system
only involves a single solvent species, i.e. water, the Born terms have not been included in this work.

Unlike the above two long-range contribution terms, the short-range NRTL term is developed as a
symmetric model and has to be normalised to obtain the un-symmetric model:

A-4
Appendix A: Thermodynamic Model Equations

G E*,NRTL G E,NRTL
R˜T R˜T
 ¦x
j
j ˜ ln J fj ,NRTL (A.2.11)

where J f,NRTL is the symmetrically normalised NRTL activity coefficient at infinite dilution. The
symmetric excess Gibbs energy is given as:

G E,lc
¦X j ˜ G jm ˜ W jm

¦X
j
˜
¦X
m
R˜T m k ˜ G km
k

X a'
¦X j ˜ G jc,a'c ˜ W jc,a'c

¦X ˜¦X
j
 ˜ ˜ (A.2.12)
¦X ¦X
c a'
c a' a '' k ˜ G kc,a'c
a '' k

X c'
¦X j ˜ G ja,c 'a ˜ W ja,c 'a

¦X ˜¦X
j
 ˜ ˜
¦X ¦X
a c'
a c' c '' k ˜ G ka,c 'a
c '' k

where
G jm
exp  D jm ˜ W jm G ca,m exp  D ca,m ˜ W ca,m (A.2.13)

G jc,ac
exp  D jc,ac ˜ W jc,ac G ja,ca
exp  D ja,ca ˜ W ja,ca (A.2.14)

¦X ˜G
a
a ca.m ¦X ˜G
c
c ca,m
G cm G am (A.2.15)
¦X a'
a' ¦X c'
c'

W mc,ac W cm  W ca,m  W m,ca W ma,ca W am  W ca,m  W m,ca (A.2.16)

¦X ˜D a ca.m ¦X ˜D c ca,m
D cm a
D am c
(A.2.17)
¦X a'
a' ¦X c'
c'

W is the binary interaction energy parameter, and D is the non-randomness factor. The subscripts a, a’,
c, c’, m and m’ refer to anions, cations and molecular species, respectively, while the subscripts j and
k refer to any species. The effective local mole fraction Xj of a species j is defined as:
Xj xj ˜ Cj (A.2.18)

where Cj is equal to zj for ions and to 1 for molecular species.

The expressions for the symmetric NRTL activity coefficients are:

A-5
Appendix A: Thermodynamic Model Equations

¦X j ˜ G jm ˜ W jm
X m' ˜ G mm'
§
¨ ¦X ˜G ˜Wk km' km'
·
¸
¦ ˜ ¨ W mm'  ¸
j
ln J NRTL  k

¦ ¦ ¦X ˜G
m
X k ˜ G km X k ˜ G km' ¨ ¸
m' ¨ k km' ¸
k k © k ¹

X a' X c ˜ G mc,a'c
§
¨ ¦X ˜G k kc,a'c ˜ W kc,a'c ·¸
 ¦¦ ˜ ¨ W mc ,a'c  k ¸ (A.2.19)
c a' ¦ a ''
X a'' ¦k
X k ˜ G kc,a'c ¨
¨
©
¦X ˜G k
k kc,a'c
¸
¸
¹

X c' X a ˜ G ma,c 'a ¨


§ ¦X ˜G k ka,c 'a˜ W ka,c 'a ·¸
 ¦¦ ˜ ¨ W ma,c 'a  k ¸
a c' ¦X ¦c ''
c ''
k
X k ˜ G ka,c 'a ¨
¨
©
¦X ˜G k
k ka,c 'a
¸
¸
¹

X m ˜ G cm
§
¨ ¦X ˜G ˜W k km km
·
¸
¦
1
ln J NRTL ˜ ¨ W cm  k ¸
¦X ¦X ˜G
c
zc ˜ G km ¨ ¸
m k ¨ k km ¸
k © k ¹

X a'
¦X ˜G k kc,a'c ˜ W kc,a'c
 ¦ ˜ k
(A.2.20)
a' ¦
a ''
X a'' ¦X ˜G k
k kc,a'c

X c' X a ˜ G ca,c 'a


§
¨ ¦X ˜G k ka,c 'a ˜ W ka,c 'a ·¸
 ¦¦ ˜ ¨ W ca,c 'a  k ¸
a c' ¦ c ''
X c '' ¦k
X k ˜ G ka,c 'a ¨
¨
©
¦X ˜G k
k ka,c 'a
¸
¸
¹

X m ˜ G am
§
¨ ¦X ˜G ˜W k km km
·
¸
¦
1
ln J NRTL ˜ ¨ W am  k ¸
¦ ¦X ˜G
a
za X k ˜ G km ¨ ¸
m ¨ k km ¸
k © k ¹

X c'
¦X ˜G k ka,c 'a ˜ W ka,c 'a
c ¦ ˜ k
(A.2.21)
c' ¦X c ''
c '' ¦X ˜G k
k ka,c 'a

X a' X c ˜ G ac,a'c
§
¨ ¦X ˜G k kc,a'c ˜ W kc,a'c ·¸
 ¦¦ ˜ ¨ W ac,a'c  k ¸
c a' ¦ a ''
X a'' ¦X
k
k ˜ G kc,a'c ¨
¨
©
¦X ˜G k
k kc,a'c
¸
¸
¹

When the solvent is pure water, the corresponding symmetrically normalised NRTL activity coefficients
at infinite dilution are given by:
f,NRTL
ln J m W wm  G wm ˜ W wm (A.2.22)

1
¦X ˜W a' wa,c 'a
ln J cf,NRTL G cw ˜ W cw  a'
(A.2.23)
zc ¦X a ''
a''

1
¦X ˜W c' wc,a'c
ln J af,NRTL G aw ˜ W aw  c'
(A.2.24)
za ¦X c ''
c ''

A-6
Appendix A: Thermodynamic Model Equations

where the subscript w refers to water. Following equation (A.1.3), the un-symmetric NRTL activity
coefficients are calculated from:
f,NRTL
*,NRTL
ln J m ln J NRTL
m  ln J m (A.2.25)

ln J c*,NRTL ln J NRTL
c  ln J cf,NRTL (A.2.26)

ln J a*,NRTL ln J NRTL
a  ln J af,NRTL (A.2.27)

In the absence of electrolytes, the Electrolyte NRTL model reduces to the NRTL model.

A-7
Appendix A: Thermodynamic Model Equations

A.3 Cubic Equations of State


A.3.1 The Soave-Redlich-Kwong Equation of State
The pressure explicit form of the Soave-Redlich-Kwong (SRK) (Soave, 1972) cubic equation of state
is:
R˜T a
P  (A.3.1)
V  b V ˜ V  b

where P is the system pressure, V is the molar volume, T is the absolute system temperature, and R
is the gas constant. In this work, the model parameters a and b are given by the classical mixing rules
proposed by Redlich and Kwong (1949):
a ¦¦ x j
˜ x k ˜ a jk ˜ 1  k jk with k jk k kj and k jj 0 (A.3.2)
j k

a jk a j ˜ a k 0.5 (A.3.3)
2
R 2 ˜ Tc, j
aj 0.42748 ˜ D j ˜ (A.3.4)
Pc, j

b ¦xj
j ˜bj (A.3.5)

R ˜ Tc, j
bj 0.08664 ˜ (A.3.6)
Pc, j

where x is the mole fraction, kjk is the binary interaction parameter for the component pair j-k, Pc is the
critical pressure, and Tc is the critical temperature, while the function Dj is given by the expression
proposed by Graboski and Daubert (1978):
2
ª § § 0 .5 ·º
Dj
«

«1  0.48508  1.55171 ˜ Z  0.15613 ˜ Z 2
j j ¨
˜ ¨1  ¨
¨
¨ © c, j
T
T ·
¸
¸
¸»
¸»
¸
(A.3.7)
¬« © ¹ ¹¼»

where  is the acentricity.

The expression for the fugacity coefficient Mj for the SRK equation of state is:
bj § P˜b ·
ln M j ˜ Z  1  ln¨ Z  ¸
b © R˜T¹
(A.3.8)
§2 bj · §

a
˜¨ ˜
b ˜ R ˜ T ¨© a k ¦ b·
x k ˜ a jk ˜ 1  k jk  ¸ ˜ ln¨1  ¸
b ¸¹ © V¹

where Z is the compressibility, which is represented by:


P˜V
Z (A.3.9)
R˜T

A-8
Appendix A: Thermodynamic Model Equations

A.3.2 The Peng-Robinson Equation of State


The pressure explicit form of the Peng-Robinson (PR) (Peng and Robinson, 1976) cubic equation of
state is:
R˜T a
P  (A.3.10)
V  b V ˜ V  b  b ˜ V  b

where P is the system pressure, V is the molar volume, T is the absolute system temperature, and R
is the gas constant. The parameters a and b are given by equations (A.3.2), (A.3.3) and (A.3.5) with
the following component parameters aj and bj:
2
R 2 ˜ Tc, j
aj 0.457235 ˜ D j ˜ (A.3.11)
Pc, j

R ˜ Tc, j
bj 0.077796 ˜ (A.3.12)
Pc, j

where Pc is the critical pressure and Tc is the critical temperature, while Dj is given by:
2
ª § § 0 .5 ·º
Dj
«

«1  0.37464  1.54226 ˜ Z  0.26992 ˜ Z 2
j j ¨
˜ ¨1  ¨
¨
T
¨ © Tc, j
·
¸
¸
¸»
¸»
¸
(A.3.13)
«¬ © ¹ ¹»¼

where  is the acentricity. In HYSYS®, when  > 0.49, Dj is instead given by (Hyprotech Ltd, 2001a):
2
ª § 0.379642  1.48503 ˜ Z · §¨ § T ·
0 .5 ·º
«1  ¨ j
¸ ˜ 1 ¨ ¸ ¸»
Dj (A.3.14)
« ¨  0.164423 ˜ Z 2  1.016666 ˜ Z 3 ¸ ¨¨ ¨ Tc, j ¸ ¸»
¸
¬« ©
j j ¹
© © ¹ ¹¼»

The expression for the fugacity coefficient Mj for the PR equation of state is:
bj § P˜b ·
ln M j ˜ Z  1  ln¨ Z  ¸
b © R˜T¹
§2

bj · § V  2  1 ˜ b ·¸ (A.3.15)
¦
a
 1 .5 ˜¨ ˜ x k ˜ a jk ˜ 1  k jk  ¸ ˜ ln¨
2 ˜ b ˜ R ˜ T ¨© a k b ¸¹ ¨© V  2  1 ˜ b ¸¹
where Z is the compressibility given by equation (A.3.9).

An enhanced version of the PR equation of state is the basis of the HYSYS® PR and Sour PR
property packages. In the Sour PR property package, the enhanced PR equation of state is combined
with the Wilson API-Sour Model (Wilson, 1980) for sour water systems. The vapour and liquid phase
fugacities and the enthalpies for the vapour, liquid and aqueous phases are determined by the
enhanced PR equation of state while the Wilson model accounts for the ionisation of CO2 and H2S in
water via the aqueous phase K-values (Hyprotech Ltd, 2001a). In the absence of an aqueous phase,
the Sour PR property package produces identical results to the PR property package.

A-9
Appendix B: Property Models for Aspen Custom Modeler®

APPENDIX B

PROPERTY MODELS FOR ASPEN


CUSTOM MODELER®

This appendix presents the various thermodynamic and physical property models utilised in the Aspen
Custom Modeler® process models for the CO2 trains. Where necessary, model parameters are
regressed from literature data.

A-10
Appendix B: Property Models for Aspen Custom Modeler®

B.1 Thermodynamic Property Models


B.1.1 Fugacity Coefficients
The fugacity coefficients used in this work were calculated from the Soave-Redlich-Kwong (SRK)
equation of state, which is described in Section A.3.1 in Appendix A. This equation of state requires
values for the component critical properties and the binary interaction parameters kjk to provide an
accurate representation of the vapour phase thermodynamics. The kjk values were taken from the
Aspen Properties® databanks, and the component critical properties used in this work are listed in
Table B.1.1. Also given are values for the molecular weight MW, critical molar volume Vc, critical
compressibility Zc, acentricity Z, and dipole moment G, of the components present in the raw natural
gas to the Moomba CO2 trains.

Table B.1.1: Component critical properties (Poling et al., 2001).

NOTE:
This table is included on page A-11 of the print copy of
the thesis held in the University of Adelaide Library.

B.1.2 Enthalpy and Heat Capacity Calculations


B.1.2.1 Vapour Phase Enthalpy and Heat Capacity
In this work, the vapour phase enthalpy hG was calculated from the sum of the ideal gas enthalpies
hIG,j of each vapour phase component and the residual enthalpy h res
G :

NC NC § T ·
¨ f ¸
hG ¦ y j ˜ h IG, j  h res
G ¦ y j ˜ ¨ 'hIG
¨
,j  ³ C pIG, j ˜ dT
¸
res
¸  hG (B.1.1)
j 1 j 1
© Tref ¹

f
where 'hIG is the ideal gas standard enthalpy of formation at the reference temperature Tref (298.15
K), CpIG is the ideal gas heat capacity, T is the absolute temperature, and NC is the number of vapour
phase components. Coefficient values for the following ideal gas heat capacity polynomial equation
(Poling et al., 2001):

A-11
Appendix B: Property Models for Aspen Custom Modeler®

C pIG
A  B ˜ T  C ˜ T2  D ˜ T3  E ˜ T4 (B.1.2)
R

where the units of CpIG depend on the value used for the gas constant R, were adapted to fit the
Aspen Properties® Ideal Gas Heat Capacity polynomial (Aspen Technology Inc, 2003) which was
used in this work:
C pIG C1  C 2 ˜ T  C 3 ˜ T 2  C 4 ˜ T 3  C 5 ˜ T 4  C 6 ˜ T 5 for C 7  T  C 8 (B.1.3)

Here, CpIG is in kJ/kmol·K and T is in K. Table B.1.2 lists the coefficient values for equation (B.1.2),
which are valid between 200 and 1000 K, along with values of 'h IG
f
,j .

The residual enthalpy h res


G was defined according to the equation of state approach:

VG
ª § wP · º
h res
G h G  h IG R ˜ T ˜ Z  1  ³«T ˜ ¨ ¸  P» ˜ dVG (B.1.4)
« © wT ¹ VG

»¼

which in terms of the SRK equation of state is:

1 § da · § b ·
h res R ˜ T ˜ Z  1  ˜ ¨a  T ˜ ¸ ˜ ln¨¨1  ¸ (B.1.5)
VG ¸¹
G
b © dT ¹ ©

where Z is compressibility, VG is the molar volume, and P is the pressure.

The vapour phase heat capacity CpG was determined from the vapour phase enthalpy hG:
§ dh G ·
C pG ¨¨ ¸¸ (B.1.6)
© dT ¹P

Table B.1.2: Ideal gas heat capacity coefficients and enthalpies of formation (Poling et al., 2001).

NOTE:
This table is included on page A-12 of the print copy of
the thesis held in the University of Adelaide Library.

A-12
Appendix B: Property Models for Aspen Custom Modeler®

B.1.2.2 Liquid Phase Enthalpy and Heat Capacity


In this work, the liquid phase enthalpy hL was obtained from the mole-fraction-weighted sum of the
liquid phase partial molar enthalpies hL,j:

§ § w ln J *j · ·¸
¦ | ¦x
NC NC NC
¨
hL ¦ x j ˜ h L, j x j ˜ h Lideal
,j  h Lex, j j ˜ ¨ h ideal
¨
aq, j  R ˜ T
2
˜¨
¨ wT
¸
¸ ¸¸
(B.1.7)
j 1 j 1 j 1
© © ¹ P,x ¹

where x is the liquid phase mole fraction, NC is the number of liquid phase species, R is the gas
constant, T is the system absolute temperature, J* is the species un-symmetric activity coefficient, and

P is the system pressure. h Lideal


,j is the ideal partial molar enthalpy, which is approximated by the

aqueous ideal partial molar enthalpy h ideal ideal


aq, j since h aq, j is evaluated at the saturation pressure of water

and not at the system pressure like h Lideal


, j . A correction factor should be included to account for the

enthalpy difference between the system pressure and the saturation pressure; however, at pressures
below 100 bar, the pressure effect is minimal and can be ignored (Zemaitis et al., 1986). According to
the un-symmetric reference state, h ideal
aq, j refers to the partial molar enthalpy of pure solvent for the

solvent H2O and to the infinite dilution partial molar enthalpy for the solute species. The temperature
derivative term represents the excess partial molar enthalpy h Lex, j , and was calculated from the

Electrolyte NRTL model. For a detailed derivation of equation (B.1.7), please refer to the work by
Zemaitis and co-workers (1986).

The aqueous ideal partial molar enthalpy h ideal


aq,H O
for H2O was determined from the expression:
2

T
h ideal ³C  'h HvapO
f
aq,H O
'hIG,H2O  pIG,H2O ˜ dT  h res
G,H O
(B.1.8)
2 2 2
Tref

where 'hIG
f
,H2O is the ideal gas standard enthalpy of formation at the reference temperature Tref

(298.15 K), CpIG,H2O is the ideal gas heat capacity, hres


G,H2O is the residual vapour enthalpy at the

saturation pressure of water, and 'hHvapO is the enthalpy of vaporisation at the saturation pressure of
2

water. 'hHvapO was obtained from the Watson equation (Reid et al., 1977):
2

§ ·
a b˜¨¨ 1 T ¸
§ 1 T · © Tc ,H2 O ¸¹
¨ Tc,H2O ¸
'h HvapO 'h H O T1 ˜ ¨
vap ¸ (B.1.9)
2 2 ¨ T1 ¸
¨ 1  Tc,H O ¸
© 2 ¹

A-13
Appendix B: Property Models for Aspen Custom Modeler®

where 'hHvapO is in kJ/kmol, Tc,H2O is the critical temperature of water in K, and T is the system
2

temperature in K. The parameters T1, 'h HvapO T1 , a and b have the values 373.200 K, 40683.136
2

kJ/kmol, 0.311 and 0.000, respectively.

The infinite dilution aqueous ideal partial molar enthalpy h ideal


aq, j for a solute species j was expressed as:

³C
ff f
h ideal
aq, j 'h aq ,j  paq, j ˜ dT for ionic solutes (B.1.10)
Tref

T
d ln H j
h ideal 'hIG ³
f
aq, j ,j  C pIG, j ˜ dT  R ˜ T 2 ˜ for molecular solutes (B.1.11)
dT
Tref

ff f
where 'h aq , j is the aqueous infinite dilution standard enthalpy of formation at Tref, C paq, j is the

aqueous infinite dilution heat capacity, 'hIG


f
, j is the ideal gas standard enthalpy of formation at Tref,

CpIG,j is the ideal gas heat capacity, and Hj is the Henry’s Law constant at infinite dilution in water at
ff
the system temperature T. Table B.1.3 presents literature values for 'h aq , j , and Table B.1.4 lists

coefficients for determining Hj at temperatures between 25 and 150°C.

Criss and Cobble (1964ab) developed the following correlation for calculating the aqueous infinite
f
dilution heat capacities C paq, j for any ionic species j:

f
T Sf f
T, j  S 25, j
C paq,j for j = H3O+ (B.1.12)
25 § T  273.15 ·
ln¨ ¸
© 298.15 ¹

f
T
a T, j  b T, j  1 ˜ S f
25, j  20.934 ˜ z j
C paq,j for j  H3O+ (B.1.13)
25 § T  273.15 ·
ln¨ ¸
© 298.15 ¹

T
f f
where C paq,j is the average value of C paq, j in kJ/kmol·K between 25°C and temperature T in °C,
25

S fT, j is the infinite dilution entropy in kJ/kmol·K at T, and zj is the ionic charge. Values for S f25, j are

given in Table B.1.3, while Table B.1.5 lists the temperature dependent values for S Tf,H O  and the
3

parameters aT,j and bT,j.

f
Aspen Properties® also provided a polynomial for calculating C paq, j (Aspen Technology Inc, 2003):

f
C 4, j C 5, j C 6, j
C paq,j C1, j  C 2, j ˜ T  C 3, j ˜ T 2   2
 (B.1.14)
T T T

A-14
Appendix B: Property Models for Aspen Custom Modeler®

f 2- - -
where C paq, j is in kJ/kmol·K and T is in K. However, parameter values for the CO3 , HCO3 and HS

ions were missing from the Aspen Properties® databanks. These values had to be regressed from
the literature data in Table B.1.6 using the Aspen Properties® DRS, and are listed in Table B.1.7.
There was no available heat capacity data for potassium bisulfide solutions, so the parameter values
for the HS- ion were set such that the polynomial predicted the same value for the infinite dilution heat
capacity at 25°C as the Criss-Cobble correlation. Figure B.1.1 compares the solution heat capacities
predicted by the Aspen Properties® polynomial against the literature values. The average absolute
deviation was 2.3%, compared to 10.7% for the heat capacities predicted by the Criss-Cobble
correlation. Consequently, the Aspen Properties® polynomial was used in this work.

The liquid phase heat capacity CpL was determined from the liquid phase enthalpy hL:
§ dh L ·
C pL ¨¨ ¸¸ (B.1.15)
© dT ¹P

Table B.1.3: Ionic species thermodynamic properties (Zemaitis et al., 1986).

NOTE:
This table is included on page A-15 of the print copy of
the thesis held in the University of Adelaide Library.

Table B.1.4: Temperature dependence of Henry’s Law constants.

Component A B C D
B
ln H A   C ˜ ln T  D ˜ T H in bar, T in K
T
CO2 a 159.1997 -8477.711 -21.9574 0.00578075
H2S b 346.6251 -13236.800 -55.0551 0.05956500
N2 c 143.4487 -6832.880 -19.1622 0
CH4 d 239.4526 -11999.000 -33.1160 0
C2H6 d 204.0936 -10500.000 -27.8321 0
C3H8 d 246.6016 -12650.000 -33.9873 0
n-C4H10 d 212.5476 -11000.700 -28.9762 0
i-C4H10 d 247.0126 -12650.000 -33.9873 0
n-C5H12 d 221.5876 -11000.600 -30.4489 0
i-C5H12 d 244.5856 -11999.100 -33.8838 0
n-C6H14 d 427.2776 -21850.000 -60.1168 0
n-C7H16 d 339.8506 -17500.000 -47.2333 0
a b c
Chen (1980) Austgen and co-workers (1989) modified from Fernández-Prini and co-workers
d
(2003) de Hemptinne and co-workers (2000)

A-15
Appendix B: Property Models for Aspen Custom Modeler®

Table B.1.5: Criss-Cobble entropy parameters (Criss and Cobble, 1964ab).

K+ OH-, HS-, S2- HCO3- CO32- H3O+


T (°C)
aT,j bT,j aT,j bT, aT,j bT, aT,j bT, SfT, j
25 0.00 1.000 0.00 1.000 0.00 1.000 0.00 1.000 -20.92 a
60 16.32 0.955 -21.34 0.969 -56.49 1.380 -58.58 1.217 -10.46
100 43.09 0.876 -54.39 1.000 -126.79 1.894 -129.72 1.476 8.37
150 67.78 0.792 -89.12 0.989 -836.92 2.381 -194.17 1.687 27.20
a
This value is used in equation (B.1.12) instead of the value in Table B.1.3.

Table B.1.6: Atmospheric solution heat capacity data.

Solution Concentration CO2 Loading Temperature Range


K2CO3 a 2 – 50 wt% K2CO3 - 30 – 130°C
KHCO3 a 4 – 20 wt% KHCO3 - 30 – 130°C
K2CO3–KHCO3 b 30 wt% equivalent K2CO3 0.2 – 0.9 70 – 120°C
a b
Aseyev and Zaytsev (1996) UOP Gas Processing (1998)

Table B.1.7: Parameter values for the Aspen Properties® heat capacity polynomial.

CO32- HCO3- HS-


Parameter a a a
Value Std. Dev. Value Std. Dev. Value Std. Dev.
CPAQO-1 0 0b 0 0b 0 0b
CPAQO-2 -0.1047 0.0103 0.0528 0.0101 -0.8426 0b
CPAQO-3 0 0b 0 0b 0 0b
CPAQO-4 0 0b 0 0b 0 0b
CPAQO-5 0 0b 0 0b 0 0b
CPAQO-6 0 0b 0 0b 0 0b
a b
Heat capacity in kJ/kmol·K. Parameter was fixed so no standard deviation was determined.

4.5

Potassium carbonate
Predicted Solution Heat Capacity (kJ/kg·K)

Potassium bicarbonate
4.0 Potassium carbonate-bicarbonate

3.5

3.0

2.5

2.0
2.0 2.5 3.0 3.5 4.0 4.5
Experimental Solution Heat Capacity (kJ/kg·K)

Figure B.1.1: Comparison between the predicted and experimental solution heat capacities. The dashed
lines (---) represent the ± 5% lines.

A-16
Appendix B: Property Models for Aspen Custom Modeler®

B.2 Physical and Transport Property Models


B.2.1 Molecular Weight
In this work, the molecular weight MW of a phase was determined from the mole-fraction-weighted
sums of the individual apparent component molecular weights:
NC
MW ¦x
j 1
j ˜ MW j (B.2.1)

where x is the apparent mole fraction and NC is the number of apparent components.

No ionic species were taken to exist in the vapour phase, so the vapour phase apparent composition
was considered to be equivalent to its true composition. In the liquid phase, the speciation and
chemical reactions occurring in the liquid phase were ignored and the apparent components were
taken to be the vapour phase components and K2CO3. The vapour phase component molecular
weights are given in Table B.1.1 and the molecular weight of K2CO3 is 138.206 kg/kmol.

B.2.2 Vapour Pressure


The pure component vapour pressures Ps were calculated in this work using the extended Antoine
equation (Rowley et al., 1998):
B
ln P s A  C ˜ T  D ˜ ln T  E ˜ T F (B.2.2)
T

where Ps is in bar and T is the temperature in K. The relevant coefficients are listed in Table B.2.1.

Table B.2.1: Antoine equation coefficients (Rowley et al., 1998).

NOTE:
This table is included on page A-17 of the print copy of
the thesis held in the University of Adelaide Library.

A-17
Appendix B: Property Models for Aspen Custom Modeler®

B.2.3 Density
B.2.3.1 Vapour Phase Density
The vapour phase molar density CG is the inverse of the vapour phase molar volume VG:
1
CG (B.2.3)
VG

which, in this work, was determined from the following form of the SRK equation of state:

3 R˜T 2 a  b ˜ R ˜ T  P ˜ b2 a˜b
VG  ˜ VG  ˜ VG  0 (B.2.4)
P P P

where VG is the largest positive real root. The vapour phase mass density UG was calculated from CG
and vapour phase molecular weight MWG:
UG C G ˜ MW G (B.2.5)

B.2.3.2 Liquid Phase Density


In this work, the liquid phase mass density UL was determined from the relation:
UL C L ˜ MWL (B.2.6)

where MWL is the liquid phase molecular weight and CL is the liquid phase molar density, which was
calculated from Amagat’s Law (Aspen Technology Inc, 2003):
NC xj
¦C
1
(B.2.7)
CL j 1 L, j

where CL,j is the liquid phase molar density of component j, x is the liquid phase mole fraction, and NC
is the number of liquid phase components.

The liquid phase molar densities of the solvent H2O and the molecular solutes were obtained from the
modified Rackett equation (Spencer and Danner, 1972):

1 R ˜ Tc, j
˜ Z RA, j
>1 1T T c, j 2 / 7 @ (B.2.8)
C L, j Pc, j

where R is the gas constant, T is the absolute temperature, Tc is the critical temperature, Pc is the
critical pressure, and ZRA is the Rackett parameter, for which values are presented in Table B.2.2.
The liquid phase electrolyte molar densities CL,ca were calculated using the Clarke Aqueous Electrolyte
Volume model (Chen et al., 1983):

1 f x ca
Vca  A ca ˜ (B.2.9)
C L,ca 1  x ca

A-18
Appendix B: Property Models for Aspen Custom Modeler®

f
where xca is the apparent electrolyte mole fraction for the electrolyte ca. Databank values for Vca and
Aca were available in Aspen Properties® for a number of cation-anion pairs, but not for the four key
pairs of interest: (K+,CO32-), (K+,HCO3-), (K+,HS-) and (K+,S2-). Given the relatively negligible quantities
of HS- and S2- ions present in the liquid phase for the process of interest, it was considered reasonable
to neglect the effect of the (K+,HS-) and (K+,S2-) pairs on the liquid phase density. The parameter
values for the (K+,CO32-) and (K+,HCO3-) pairs were regressed from the literature data in Table B.2.3
using the Aspen Properties® DRS, and are presented in Table B.2.4. Figure B.2.1 compares the
solution mass densities predicted by equations (B.2.6) to (B.2.9) against the literature values. The
average absolute deviation between the predicted and literature values is 0.4%.

Table B.2.2: Parameter values for the modified Rackett equation (Spencer and Danner, 1972).

Component ZRA
CO2 0.2736
H2S 0.2851
H2O 0.2432
N2 0.2905
CH4 0.2876
C2H6 0.2789
C3H8 0.2763
n-C4H10 0.2728
i-C4H10 0.2750
n-C5H12 0.2685
i-C5H12 0.2716
n-C6H14 0.2635
n-C7H16 0.2611

Table B.2.3: Atmospheric solution mass density data.

Solution Concentration CO2 Loading Temperature Range


K2CO3 a 20 – 40 wt% K2CO3 - 30 – 100°C
K2CO3 b 20 – 40 wt% K2CO3 - 30 – 90°C
KHCO3 b 1 – 30 wt% KHCO3 - 30 – 80°C
K2CO3–KHCO3 c 22 – 30 wt% K2CO3 e 0 – 0.6 70 – 115°C
K2CO3–KHCO3 d 30 wt% K2CO3 e 0–1 70 – 130°C
a b c
Armand Products Company (1998) Chernen’kaya and Revenko (1975) Bocard and Mayland (1962)
d e
UOP Gas Processing (1998) These concentrations are in terms of equivalent K2CO3 weight percent.

Table B.2.4: Pair parameter values for the Clarke Aqueous Electrolyte Volume model.

Parameter Component j Component k Parameter Value a Standard Deviation


f + 2-
Vca K CO3 0.007465 0.001818
f + -
Vca K HCO3 0.018397 0.002006
+ 2-
Aca K CO3 0.136841 0.009026
Aca K+ HCO3- 0.077836 0.009483
a 3
Molar volume in m /kmol.

A-19
Appendix B: Property Models for Aspen Custom Modeler®

1600
Potassium carbonate
(Armand Product Company, 1998)
1500 Potassium carbonate
(Chernen'kaya & Revenko, 1975)

Predicted Solution Mass Density (kg/m 3)


Potassium bicarbonate
1400 (Chernen'kaya & Revenko, 1975)
Potassium carbonate-bicarbonate
(Bocard & Mayland, 1962)
Potassium carbonate-bicarbonate
1300
(UOP, 1998)

1200

1100

1000

900

800
800 900 1000 1100 1200 1300 1400 1500 1600
3
Experimental Solution Mass Density (kg/m )

Figure B.2.1: Comparison between the predicted and experimental solution mass densities. The dashed
lines (---) represent the ± 1% lines.

B.2.3.3 Partial Molar Volume


The partial molar volume v fj of a component j at infinite dilution in water is required to determine the

effect of pressure on its Henry’s Law constant. The method of Brelvi and O’Connell (1972) was used
in this work to calculate v fj :

v fj
1  C $jw (B.2.10)
Z $j ˜R ˜ T

where v fj is in cm3/mol, Z $j is the isothermal compressibility at infinite dilution in water in 1/atm:

§ VLs,H O ·¸
¨
ln¨1  $ 2 ¸ 0.42704 ˜ ~
v j  1  2.089 ˜ ~
v j  1  0.42367 ˜ ~
2
vj 1
3
(B.2.11)
¨ Zj ˜R ˜ T ¸
© ¹

C$jw is the reduced volume integral of component j at infinite dilution in water:

§ § v* ·
0.62 ·
¨ ¸
ln¨  C $jw ˜ ¨ w* ¸
¸ 2.4467  2.12074 ˜ ~
vj for 2.0 d ~
v j d 2.785 (B.2.12)
¨v ¸
¨ © j ¹ ¸
© ¹

§ § v* ·
0.62 ·
¨ ¸
ln¨  C $jw ˜ ¨ w* ¸ 3.02214  1.87085 ˜ ~
v j  0.71955 ˜ ~
v j for 2.785 d ~
2
¸ v j d 3 .2 (B.2.13)
¨v ¸
¨ © j ¹ ¸
© ¹

A-20
Appendix B: Property Models for Aspen Custom Modeler®

T is the system temperature in K, R is the gas constant (82.06 atm·cm3/mol·K), VLs,H2O is the liquid

phase molar volume of water in at its saturation pressure in cm3/mol, v *j is the characteristic volume

for which values are listed in Table B.2.5, and ~


v j is the dimensionless reduced molar volume:

~ v *j
vj (B.2.14)
VLs,H O
2

Table B.2.5: Component characteristic volumes.

Component v* (cm3/mol)
CO2 a 93.94
H2S a 88.74
H2O a 46.40
N2 b 89.64
CH4 b 99.21
C2H6 b 148.17
C3H8 b 203.49
n-C4H10 b 254.92
i-C4H10 b 263.24
n-C5H12 b 303.16
i-C5H12 b 306.51
n-C6H14 b 369.44
n-C7H16 b 431.76
a
Austgen and co-workers (1989)
b
Zemaitiis and co-workers (1986)

B.2.4 Viscosity
B.2.4.1 Vapour Phase Viscosity
In this work, the vapour phase viscosity PG was determined in two steps. First, the low pressure

vapour phase viscosity P LP


G was determined from the Wilke method (Wilke, 1950):

NC y j ˜ P LP
¦
G, j
P LP
G NC
(B.2.15)
j 1
¦y
k 1
k ˜ I jk

2
ª § LP ·
0.5 0.25 º
«1  ¨ P ¸ § MWk · »
G, j
˜ ¨ ¸
« ¨© P LP ¸
G,k ¹ © MW j ¹ »
¬ ¼
I jk 0 .5
(B.2.16)
ª § MW j ·º
«8 ˜ ¨1  MW
¸»
¬ © k ¹¼

where yj and yk are the vapour phase mole fractions of components j and k, P LP
G, j and P G,k are the
LP

component low pressure vapour phase viscosities, MWj and MWk are the component molecular

A-21
Appendix B: Property Models for Aspen Custom Modeler®

weights, and NC is the number of vapour phase components. The Dean-Stiel pressure correction
(Dean and Stiel, 1965) was then applied to P LP
G to give PG:

§ V §V ·
1.858
·
1.08 u 10  4 ¨ 1.439˜ c ,G VG 1.111˜¨ c ,G
© V ¸

¸
PG  P LP
G ˜ ¨e e ¸ (B.2.17)
[ ¨ ¸
© ¹
1/ 6
Tc,G
[ 1/ 2 2/3
(B.2.18)
MWG ˜ Pc,G

where PG and P LP
G are in cP, MWG is the vapour phase molecular weight in kg/kmol, VG is the vapour

phase molar volume, Vc,G is the vapour phase critical molar volume, Tc,G is the vapour phase critical
temperature in K, and Pc,G is the vapour phase critical pressure in atm. The vapour phase critical
properties were obtained as follows:
NC
Vc,G ¦y
j 1
j ˜ Vc, j (B.2.19)

NC
Tc,G ¦y
j 1
j ˜ Tc, j (B.2.20)

NC
Z c,G ¦y
j 1
j ˜ Z c, j (B.2.21)

Z c,G ˜ R ˜ Tc,G
Pc,G (B.2.22)
Vc,G

where R is the gas constant, Zc,G is the vapour phase critical compressibility, and Vc,j, Tc,j and Zc,j are
the component critical molar volume, temperature and compressibility, respectively.

The component low pressure vapour phase viscosities P LP


G, j were calculated using the DIPPR vapour

viscosity model (Rowley et al., 1998):

A ˜ TB
P LP
G, j (B.2.23)
1 C  D 2
T T

where P LP
G, j is in cP and T is the system temperature in K. The relevant coefficients are given in Table

B.2.6.

A-22
Appendix B: Property Models for Aspen Custom Modeler®

Table B.2.6: Coefficients for the DIPPR vapour viscosity model (Rowley et al., 1998).

NOTE:
This table is included on page A-23 of the print copy of
the thesis held in the University of Adelaide Library.

B.2.4.2 Liquid Phase Viscosity


In this work, the liquid phase viscosity PL was determined from the Jones-Dole equation (Jones and
Dole, 1929):

§ ·
PL P L,solvent ˜ ¨1 
¨
©
¦ 'P
ca
ca
¸
¸
¹
(B.2.24)

where PL,solvent is the viscosity of the solvent, which was determined from the mole-fraction-weighted
sum of the component liquid viscosities PL,j for H2O and any molecular species present in the liquid
phase:
NC
P L,solvent ¦x
j 1
j ˜ P L, j (B.2.25)

x is the liquid phase mole fraction and NC is the number of molecular species present in the liquid
phase. The values of PL,j were calculated from the Andrade equation (Reid et al., 1977) with the
coefficients given in Table B.2.7:
Bj
ln P L, j Aj  (B.2.26)
T

where PL,j is in cP and T is in K.

The contribution term 'Pca was determined using the Breslau-Miller equation (Breslau and Miller,
1972; Breslau et al., 1974):

2.5 ˜ Ve ˜ C L,ca  10.05 ˜ Ve ˜ C L,ca


2
'P ca (B.2.27)

where CL,ca is the molar concentration of the apparent electrolyte ca:

A-23
Appendix B: Property Models for Aspen Custom Modeler®

C L,ca x ca ˜ C L (B.2.28)

Ve is the effective volume in L/mol:


B ca  0.002
Ve for electrolytes involving univalent ions (B.2.29)
2.60
B ca  0.011
Ve for other electrolytes (B.2.30)
5.06
B ca b c,1  b c,2 ˜ T  b a,1  b a,2 ˜ T (B.2.31)

xca is the apparent electrolyte mole fraction, CL is the liquid phase molar density, and T is the system
absolute temperature.

Databank values for the parameters bc,1, bc,2, ba,1 and ba,2 were available in Aspen Properties® for a
number of cations and anions, but not for the K+, HCO3-, CO32-, HS- and S2- ions. Given the relatively
negligible quantities of HS- and S2- ions present in the liquid phase for the process of interest, it was
considered reasonable to neglect the effect of these two ions on the liquid phase viscosity. The
parameter values for the K+, HCO3- and CO32- ions were regressed from the literature data in Table
B.2.8 using the Aspen Properties® DRS, and are given in Table B.2.9. Figure B.2.1 compares the
solution viscosities predicted by equations (B.2.24) to (B.2.31) against the literature values. The
average absolute deviation between the predicted and literature values is 3.1%.

Table B.2.7: Coefficients for the Andrade liquid viscosity equation (Reid et al., 1977).

NOTE:
This table is included on page A-24 of the print copy of
the thesis held in the University of Adelaide Library.

A-24
Appendix B: Property Models for Aspen Custom Modeler®

Table B.2.8: Atmospheric solution viscosity data.

Solution Concentration CO2 Loading Temperature Range


K2CO3 a 20 – 40 wt% K2CO3 - 35 – 90°C
K2CO3 b 20 – 25 wt% K2CO3 - 75 – 90°C
K2CO3 c 28 – 40 wt% K2CO3 - 30 – 60°C
K2CO3 d 22 – 30 wt% K2CO3 - 70 – 130°C
KHCO3 a 10 – 30 wt% KHCO3 - 35 – 75°C
K2CO3–KHCO3 e 30 wt% K2CO3 f 0.6 70 – 130°C
a b c
Chernen’kaya and Revenko (1975) Correia and co-workers (1980) Gonçalves and Kestin (1981)
d e f
Bocard and Mayland (1962) UOP Gas Processing (1998) This concentration is in terms of
equivalent K2CO3 weight percent .

Table B.2.9: Parameter values for the Jones-Dole viscosity equation.

K+ HCO3- CO32-
Parameter
Value Std. Dev. Value Std. Dev. Value Std. Dev.
a
bc,1 13.0705 4.7165 - - - -
bc,2 b -0.0295 0.0113 - - - -
ba,1 a - - -12.7297 4.7312 -25.1305 9.3835
ba,2 b - - 0.02893 0.01138 0.0576 0.0225
a 3 b 3
Parameter in m /kmol. Parameter in m /kmol·K.

3.5
Potassium carbonate
(Chernen'kaya & Revenko, 1975)
Potassium carbonate
3.0 (Correia et al, 1980)
Potassium carbonate
(Gonçalves & Kestin, 1981)
Potassium carbonate
Predicted Solution Viscosity (cP)

2.5 (Bocard & Mayland, 1962)


Potassium bicarbonate
(Chernen'kaya & Revenko, 1975)
Potassium carbonate-bicarbonate
2.0 (UOP, 1998)

1.5

1.0

0.5

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

Experimental Solution Viscosity (cP)

Figure B.2.2: Comparison between the predicted and experimental solution viscosities. The dashed lines
(---) represent the ± 5% lines.

A-25
Appendix B: Property Models for Aspen Custom Modeler®

B.2.5 Surface Tension


In Aspen Properties®, the Onsager-Samaras model (Onsager and Samaras, 1934) is used to
determine the liquid phase surface tension V L for an electrolyte system:

VL V L,solvent  ¦x ca
ca ˜ 'V ca (B.2.32)

where VL,solvent is the surface tension of the solvent, which is determined from the mole-fraction-
weighted sum of the component surface tensions VL,j for H2O and any molecular species present in the
liquid phase:
NC
V L,solvent ¦x
j 1
j ˜ V L, j (B.2.33)

x is the liquid phase mole fraction and NC is the number of molecular species present in the liquid
phase. The component surface tensions VL,j are calculated from the DIPPR surface tension equation
(Rowley et al., 1998):
2 3
§ · § ·
B j C j ˜ T D j ˜¨¨ T ¸¸ E j ˜¨¨ T ¸¸
§ T ·¸
Tc , j © Tc , j ¹ © Tc , j ¹
V L, j A j ˜ ¨1  (B.2.34)
¨ Tc, j ¸¹
©

where VL,j is in N/m, the temperature T and the component critical temperature Tc,j are in K, and the
relevant coefficients are given in Table B.2.10.

The contribution term 'Vca is defined as (Horvath, 1985):

§ 1.13 u 10 13 ˜ H H O ˜ T 3 ·¸
˜ C L,ca ˜ log ¨¨
80.0
'V ca 2
¸ (B.2.35)
H H2O C L,ca
© ¹

where H H2O is the dielectric constant of water (Harned and Owen, 1958):

§1 1 ·
H H2O 78.540  31989.380 ˜ ¨  ¸ (B.2.36)
© T 298.15 ¹

CL,ca is the molar concentration of the apparent electrolyte ca, and T is the system temperature in K.

When equations (B.2.32) to (B.2.36) were applied to carbonate-bicarbonate solutions, the predicted
surface tensions were found to be inconsistent with the literature data in Table B.2.11. Since there
were no adjustable parameters which could be regressed to give a better fit to the data, it was decided
to extrapolate the literature data to develop an empirical surface tension correlation of the form:

A-26
Appendix B: Property Models for Aspen Custom Modeler®

VL A  A ˜ wf
0 1 K 2CO3  A 2 ˜ wf K 2CO3
2

˜ > B  B ˜ T  B 2  C 0  C1 ˜ T @
(B.2.37)
0 1 2 ˜ T 2 ˜ FCO2  0.35

where VL is in N/m, T is the temperature in °C, wf K 2CO3 is the equivalent K2CO3 weight fraction, and

F CO2 is the CO2 loading. The correlation coefficients were determined via the simple unweighted

least squares method in a Microsoft® Excel spreadsheet, and the resulting values are given in Table
B.2.12. Since the surface tension of water is not affected by pressure within the pressure range of
interest (Haar et al., 1984), VL was assumed to be similarly unaffected by pressure.

The above correlation is valid for temperatures between 20° and 130°C, equivalent K2CO3
concentrations of 20 to 40 wt% and CO2 loadings between 0 and 1. These conditions encompass the
typical absorber and regenerator operating conditions for the Moomba CO2 trains. It was considered
reasonable to neglect the effect of the H2S loading on the surface tension since relatively negligible
quantities of H2S are absorbed in the CO2 trains. The typical H2S loading is of the order of 10-4, which
is several orders of magnitude less than the CO2 loading for the CO2 trains. Similarly, it was also
considered reasonable to disregard the presence of other gases (such as N2, CH4 and other
hydrocarbons) as negligible amounts of these gases are dissolved into the liquid phase.

Figure B.2.3 compares the surface tension values predicted by equation (B.2.37) against the literature
values. The average absolute deviation is 0.4%, compared to 7.6% for the values predicted by the
Onsager-Samaras model and the DIPPR equation. As a result, the empirical correlation was used in
this work.

Table B.2.10: Coefficients for the DIPPR surface tension equation (Rowley et al., 1998).

NOTE:
This table is included on page A-27 of the print copy of
the thesis held in the University of Adelaide Library.

A-27
Appendix B: Property Models for Aspen Custom Modeler®

Table B.2.11: Atmospheric solution surface tension data.

Solution Concentration CO2 Loading Temperature Range


K2CO3 a 0 – 50 wt% K2CO3 - 20°C
K2CO3–KHCO3 b 27 – 30 wt% K2CO3 d 0 – 0.35 20 – 130°C
Na2CO3–NaHCO3 c 20 wt% Na2CO3 e 0 – 0.9 25°C
a b c d
Armand Products Company (1998) UOP Gas Processing (1998) Bedekar (1955) These
e
concentrations are in terms of equivalent K2CO3 weight percent. This concentration is in terms of
equivalent Na2CO3 weight percent.

Table B.2.12: Surface tension correlation coefficients.

Coefficient Value
A0 8.7129×10-1
A1 0
A2 1.4250
B0 2.3859×10-2
B1 4.9160×10-5
B2 6.0020×10-7
C0 8.2563×10-2
C1 -1.4361×10-4

0.12
Predicted Solution Surface Tension (N/m)

Potassium carbonate
0.11
Potassium carbonate-bicarbonate

0.10

0.09

0.08

0.07

0.06
0.06 0.07 0.08 0.09 0.10 0.11 0.12
Experimental Solution Surface Tension (N/m)

Figure B.2.3: Comparison between the predicted and experimental solution surface tensions. The dashed
lines (---) represent the ± 1% lines.

A-28
Appendix B: Property Models for Aspen Custom Modeler®

B.2.6 Thermal Conductivity


B.2.6.1 Vapour Phase Thermal Conductivity
Like the vapour phase viscosity, the vapour phase thermal conductivity OG was determined in two

steps in this work. First, the low pressure vapour phase thermal conductivity OLP
G was determined from

the Wassiljewa-Mason-Saxena method (Mason and Saxena, 1958):


NC y j ˜ OLP
¦
G, j
OLP
G NC
(B.2.38)
j 1
¦y
k 1
k ˜ I jk

2
ª § LP ·
0.5 0.25 º
«1  ¨ P G, j ¸ ˜
§ MWk
¨
·
¸ »
« ¨© P LP ¸
G,k ¹ © MW j ¹ »
¬ ¼
I jk 0 .5
(B.2.39)
ª § MW j ·º
«8 ˜ ¨1  MW
¸»
¬ © k ¹¼

where OLP
G, j is the component low pressure thermal conductivity, yj and yk are the component vapour

phase mole fractions, P LP LP


G, j and P G,k are the component low pressure vapour phase viscosities, MWj

and MWk are the component molecular weights, and NC is the number of vapour phase components.

G to give OG:
The Stiel-Thodos pressure correction (Stiel and Thodos, 1964) was then applied to OLP

§ 0.535˜Vc,G ·
O G  OLP
G ˜ * ˜ Z c,G
5
1.22 u 10  2 ˜ ¨ e
¨
VG
 1¸
¸
for
Vc,G
VG
 0. 5 (B.2.40)
© ¹

§ 0.67˜Vc,G ·
O G  OLP
G ˜ * ˜ Z c,G
5
1.14 u 10 2 ˜ ¨ e
¨
VG
 1.069 ¸
¸
for 0.5 
Vc,G
VG
 2. 0 (B.2.41)
© ¹

§ 1.155˜Vc ,G ·
O G  OLP
G ˜ * ˜ Z c,G
5
2.60 u 10 3 ˜ ¨ e
¨
VG
 2.016 ¸ for 2.0 
¸
Vc,G
VG
 2. 8 (B.2.42)
© ¹
1/ 6
§ Tc,G ˜ MWG 3 ·
* 210 ˜ ¨ ¸ (B.2.43)
¨ Pc,G
4 ¸
© ¹

Zc,G, Vc,G, Tc,G and Pc,G are the vapour phase critical compressibility, critical molar volume, critical
temperature in K and critical pressure in bar, respectively, and the parameter * is in m·K/W . VG is the
vapour phase molar volume, and MWG is the vapour phase molecular weight in kg/kmol. The vapour
phase critical properties were obtained as follows (Poling et al., 2001):
NC NC
Vc,G ¦¦ y
j 1 k 1
j ˜ y k ˜ Vc, jk (B.2.44)

A-29
Appendix B: Property Models for Aspen Custom Modeler®

NC NC

¦¦ y
j 1 k 1
j ˜ y k ˜ Vc, jk ˜ Tc, jk
Tc,G (B.2.45)
Vc,G

NC
ZG ¦y
j 1
j ˜ Zj (B.2.46)

Z c,G 0.291  0.08 ˜ Z G (B.2.47)

Z c,G ˜ R ˜ Tc,G
Pc,G (B.2.48)
Vc,G

Vc, jk
V c, j
1/ 3
 Vc,k
1/ 3 3
(B.2.49)
8

Tc, jk Tc,j ˜ Tc,k 0.5 (B.2.50)

where ZG is the vapour phase acentricity, R is the gas constant, Vc,j and Vc,k are the component critical
molar volumes, Tc,j and Tc,k are the component critical temperatures, Zj is the component acentricity,
and Zc,j is the component critical compressibility.

The component low pressure vapour phase thermal conductivities OLP


G, j are calculated using the

DIPPR vapour thermal conductivity model (Rowley et al., 1998):


Bj
Aj ˜T
OLP
G, j (B.2.51)
Cj Dj
1 
T T2

where OLP
G, j is in W/m·K and T is the system temperature in K. The relevant coefficients are given in

Table B.2.13.

Table B.2.13: Coefficients for the DIPPR vapour thermal conductivity model (Rowley et al., 1998).

NOTE:
This table is included on page A-30 of the print copy of
the thesis held in the University of Adelaide Library.

A-30
Appendix B: Property Models for Aspen Custom Modeler®

B.2.6.2 Liquid Phase Thermal Conductivity


In Aspen Properties®, the liquid phase thermal conductivity OL for an electrolyte system is determined
from the Riedel equation (Riedel, 1951):

§
O L,solvent ·
OL ˜ ¨ O L,solvent (293K ) 
O L,solvent (293K ) ¨© ¦ a
ca
c  a a ˜ C L,ca ¸
¸
¹
(B.2.52)

where OL,solvent is the solvent thermal conductivity at the system temperature T, OL(293K) is the solvent
thermal conductivity at 293 K, CL,ca is the molar concentration of the apparent electrolyte ca, and ac
and aa are the cationic and anionic Riedel parameters.

The solvent thermal conductivity is determined from the component thermal conductivities OL,j of H2O
and any molecular species present in the liquid phase via the Vredeveld mixing rule (Reid et al.,
1977):
0.5
§ ·
¦ wf
¨ 2 ¸
O L,solvent ˜ O L, j (B.2.53)
¨ j ¸
© j ¹

where wfj is the component weight fraction and NC is the number of molecular species present in the
liquid phase. The values of OL,j are calculated from the DIPPR equation (Rowley et al., 1998) with the
coefficients given in Table B.2.14:
O L, j A j  B j ˜ T  C j ˜ T2  D j ˜ T3  E j ˜ T 4 (B.2.54)

where OL,j is in W/m·K and T is in K.

Using the Aspen Properties® databank values for the parameters ac and aa, equations (B.2.52) to
(B.2.54) were applied to potassium carbonate-bicarbonate solutions. The resulting liquid phase
thermal conductivity values were found to be inconsistent with the literature data in Table B.2.15. The
ac and aa values were re-regressed from the literature data using the Aspen Properties® DRS, but this
did not improve the performance of equations (B.2.52) to (B.2.54). Consequently, it was decided to
extrapolate the literature data to develop an empirical liquid phase thermal conductivity correlation of
the form:

OL A 0  A 1 ˜ wf K 2CO3  A 2 ˜ FCO2  A 3 ˜ T  A 4 ˜ T 2 (B.2.55)


˜ B 0  B1 ˜ T  B 2 ˜ T 2  B 3 ˜ P

where OL is in W/m·K, T is the temperature in °C, P is the pressure in bar, wf K 2CO3 is the equivalent

K2CO3 weight fraction, and F CO2 is the CO2 loading. The correlation coefficients were determined via

the simple unweighted least squares method in a Microsoft® Excel spreadsheet, and the resulting
values are given in Table B.2.16. The dependence of OL on pressure was assumed to be the same as

A-31
Appendix B: Property Models for Aspen Custom Modeler®

that for the thermal conductivity of water, which increases linearly with pressure within the pressure
range of interest (Haar et al., 1984). The correlation is valid for temperatures between 20° and 130°C,
pressures between 1 and 75 bar, equivalent K2CO3 concentrations of 20 to 40 wt% and CO2 loadings
between 0 and 1.

Figure B.2.4 compares the solution thermal conductivity values predicted by equation (B.2.56) against
the literature values. The average absolute deviation is 0.5%, compared to 10.6% for the values
predicted by the Riedel and DIPPR equations. As a result, the empirical correlation was used in this
work.

Table B.2.14: Coefficients for the DIPPR liquid thermal conductivity equation (Rowley et al., 1998).

NOTE:
This table is included on page A-32 of the print copy of
the thesis held in the University of Adelaide Library.

Table B.2.15: Atmospheric solution thermal conductivity data.

Solution Concentration CO2 Loading Temperature Range


K2CO3 a 20 – 40 wt% K2CO3 - 20 – 90°C
K2CO3–KHCO3 b 27 – 30 wt% K2CO3 c 0 – 0.75 70 – 130°C
a b c
Chernen’kaya and Revenko (1973) UOP Gas Processing (1998) These concentrations are in terms
of equivalent K2CO3 weight percent.

Table B.2.16: Liquid phase thermal conductivity correlation coefficients.

Coefficient Value
A0 6.2933×10-1
A1 -1.8732×10-1
A2 -2.3768×10-2
A2 1.2144×10-3
A4 -4.9031×10-6
B0 9.9983×10-1
B1 1.1860×10-5
B2 1.0952×10-7
B3 8.0807×10-5

A-32
Appendix B: Property Models for Aspen Custom Modeler®

0.68

Predicted Solution Thermal Conductivity


Potassium carbonate
0.66
Potassium carbonate-bicarbonate

0.64

(W/m·K)
0.62

0.60

0.58

0.56

0.54
0.54 0.56 0.58 0.60 0.62 0.64 0.66 0.68

Experimental Solution Thermal Conductivity


(W/m·K)

Figure B.2.4: Comparison between the predicted and experimental solution thermal conductivities. The
dashed lines (---) represent the ± 1% lines.

B.2.7 Diffusivities
B.2.7.1 Vapour Phase Diffusivities
In this work, the vapour phase component diffusivities Dc,j were calculated using Blanc’s Law (Reid et
al., 1977):
§ ·
NC ¨ NC D G, jk ¸
D G, j ¦ ¦
y k ˜¨
¨ k 1 yk
¸
¸
(B.2.56)
k 1 ¨ kz j ¸
kz j © ¹

where yj and yk are the component vapour phase mole fractions, NC is the number of vapour phase
components, and Dg,jk is the vapour phase binary diffusivity for the component pair j-k, which was
determined from the corresponding binary diffusivity D LP
G, jk at atmospheric pressure using the Dawson-

Khoury-Kobayashi expression (Reid et al., 1977):


D G, jk ˜ C G 2 3
1  0.053432 ˜ C R  0.030182 ˜ C R  0.029725 ˜ C R (B.2.57)
D LP
G, jk ˜ C LP
G

CG and C LP
G are the vapour phase molar densities at the system pressure P and at atmospheric

pressure, respectively. CR is the reduced molar density, which was determined from the relation:
§ y j ˜ Vc, j  y k ˜ Vc,k ·
CR C G ˜ Vc, jk CG ˜ ¨ ¸ (B.2.58)
¨ y j  yk ¸
© ¹

where Vc,j and Vc,k are the component critical volumes.

A-33
Appendix B: Property Models for Aspen Custom Modeler®

The atmospheric vapour phase binary diffusivities D LP


G, jk were obtained from the Chapman-Enskog-

Brokaw model (Poling et al., 2001):

0.00266 ˜ T 3 / 2
D LP
G, jk 1/ 2 2
(B.2.59)
P ˜ MW jk ˜ V jk ˜ : D, jk
1
§ 1 1 ·¸
MW jk 2˜¨  (B.2.60)
¨ MW j MWk ¸
© ¹

2
where D LP
G, jk is in cm /s, T is the system temperature in K, P is the system pressure in bar, MWj and

MWk are the component molecular weights in kg/kmol, Vjk is the characteristic length in Å, and :D,jk is
the diffusion collision integral.

The characteristic length Vjk was obtained from the following relations:

V jk V j ˜ V k 1/ 2 (B.2.61)

1/ 3
§ 1.585 ˜ Vb, j ·
Vj ¨ ¸ (B.2.62)
¨ 1  1 .3 ˜ G 2 ¸
© p, j ¹

G p, jk G p,j ˜ G p,k 1/ 2 (B.2.63)

2
1.94 u 10 3 ˜ G j
G p, j (B.2.64)
Vb, j ˜ Tb, j

where Tb,j is the normal boiling point in K, Vb,j is the component liquid molar volume at Tb,j in cm3/mol,
and Gp,j is the component polar parameter, Gj is the component dipole moment in Debye. Values of Vb,j
and Tb,j are given in Table B.2.17.

The diffusion collision integral :D,jk was determined from the correlation:
2
A C E G 0.19 ˜ G p, jk
: D, jk  *
 *
 *
 (B.2.65)
e D˜T e F˜T e H˜T
B
T* T*

where A is 1.06036, B is 0.15610, C is 0.19300, D is 0.47635, E is 1.03587, F is 1.52996, G is


1.76474 and H is 3.89411. The dimensionless temperature T* was calculated from:
k˜T
T* (B.2.66)
H jk
1/ 2
H jk § H j Hk ·
¨ ˜ ¸ (B.2.67)
k ¨k k ¸
© ¹
Hj
k

1.18 ˜ 1  1.3 ˜ G p, j
2
˜ T b, j (B.2.68)

A-34
Appendix B: Property Models for Aspen Custom Modeler®

where Hj and Hk are the component characteristic energies and k is the Boltzmann constant.

Table B.2.17: Normal boiling points and the corresponding liquid molar volumes (Poling et al., 2001).

NOTE:
This table is included on page A-35 of the print copy of
the thesis held in the University of Adelaide Library.

B.2.7.2 Liquid Phase Diffusivities


For ionic and electrolyte species, the liquid phase species diffusivities DL,j were determined in this
work using the Stokes-Einstein relation (Poling et al., 2001):

D L, j ˜ P L D fw, j ˜ P L25,H O
2
(B.2.69)
T 298.15K

where T is the system temperature in K, PL is the liquid phase viscosity, D fw, j is the species diffusivity

at infinite dilution in water at 25°C, and P L25,H2O is the viscosity of water at 25°C. P L25,H2O was calculated

from (Bingham, 1922):

0.021482 ˜ ª« T  8.435  8078.4  T  8.435 º»  1.2


1 2
(B.2.70)
P L,H2O ¬ ¼

where P L,H2O is in cP and T is in °C. The Nernst equation (Horvath, 1985) was used to obtain the

values of D fw, j :

Ofj ˜ R ˜ T
D fw, j for j = ionic species (B.2.71)
z j ˜ F2

z c  z a ˜ Ofc ˜ Ofa ˜ R ˜ T
D fw, j

z c ˜ z a ˜ Ofc  Ofa ˜ F 2
for j = electrolyte species ca (B.2.72)

A-35
Appendix B: Property Models for Aspen Custom Modeler®

where Ofj is the ionic conductivity at infinite dilution, zj is the charge number (or the absolute value of

the species ionic charge), R is the gas constant, T is 298.15 K, and F is Faraday’s constant. The
subscripts a and c refer to the anions and cations comprising the electrolyte species ca. Values of Ofj

are listed in Table B.2.18.

For molecular species, the liquid phase species diffusivities DL,j were determined using the correlation
suggested by Ratcliff and Holdcroft (1963):

§ D w, j · § PL ·
log¨ ¸ 0.637 ˜ log¨ ¸ (B.2.73)
¨ D L, j ¸ ¨ P L,H O ¸
© ¹ © 2 ¹

where PL is the liquid phase viscosity, Dw,j is the species diffusivity in water, and P L,H2O is the viscosity

of water.

The values of Dw,j for CO2 and H2S were calculated from the following correlations, which were derived
from the data presented by Versteeg and van Swaaij (1988) and Tamimi and co-workers (1994):
2184 .8
D w,CO2 2.9294 u 10 2 ˜ e T (B.2.74)

1839 .7
D w,H2S 8.7161 u 10 3 ˜ e T (B.2.75)

where D w,CO2 and D w,CO2 are in cm2/s and T is the temperature in K. For the other molecular

species present in the liquid phase, Dw,j was determined using the Stokes-Einstein relation:

D w, j ˜ P L,H2O D 25 25
w , j ˜ P L,H 2O
(B.2.76)
T 298.15K

where D 25
w, j is the species diffusivity in water at 25°C, values of which are given in Table B.2.19, and

PL25,H2O is the viscosity of water at 25°C.

A-36
Appendix B: Property Models for Aspen Custom Modeler®

Table B.2.18: Ionic conductivities at infinite dilution.

Of (m /·kmol)
2
Ionic Species
H3O+ 35.010 a
K+ 7.352 a
OH- 19.830 a
HCO3- 4.450 a
CO32- 6.930 a
HS- 6.500 b
S2- 9.990 b
a b
Robinson and Stokes (1965) Onda and
co-workers (1971)

Table B.2.19: Diffusivities in water at 25°C.


5 2
Component D 25
w , j ×10 (cm /s)

H2O 2.57 a
N2 2.01 b
CH4 1.88 c
C2H6 1.52 c
C3H8 1.21 c
n-C4H10 1.05 d
i-C4H10 1.04 d
n-C5H12 0.93 d
i-C5H12 0.93 d
n-C6H14 0.84 d
n-C7H16 0.77 d
a b
Wang (1965) Ferrell and Himmelblau
c d
(1967) Witherspoon and Saraf (1965)
Determined from the equation proposed by
Hayduk and Laudie (1974).

A-37
Appendix C: Electrolyte NRTL Adjustable Parameters

APPENDIX C

ELECTROLYTE NRTL
ADJUSTABLE PARAMETERS

This appendix outlines the procedure followed for the regression of the model parameters for the
Electrolyte NRTL model. The results of the data regression are analysed and the complete set of
model parameters is included within.

A-38
Appendix C: Electrolyte NRTL Adjustable Parameters

C.1 Parameter Values


The Electrolyte NRTL adjustable parameters for determining the energy parameter W from equation
(3.2.4) are presented in Tables C.1.1 and C.1.2. The listed values were taken from the Aspen
Properties® databanks, unless indicated otherwise. The default values for W and the non-randomness
factor D were set as 0 and 0.2, respectively.

Table C.1.1: The Electrolyte NRTL adjustable parameters used in this work.

Component i Component j A B C D
CO2 (H3O+,OH-) 15.000 0 0 0.1
(H3O+,OH-) CO2 -8.000 0 0 0.1
CO2 (H3O+,CO32-) 15.000 0 0 0.1
(H3O+,CO32-) CO2 -8.000 0 0 0.1
CO2 (H3O+,HCO3-) 15.000 0 0 0.1
(H3O+,HCO3-) CO2 -8.000 0 0 0.1
CO2 (H3O+,HS-) 15.000 0 0 0.1
(H3O+,HS-) CO2 -8.000 0 0 0.1
CO2 (H3O+,S2-) 15.000 0 0 0.1
(H3O+,S2-) CO2 -8.000 0 0 0.1
CO2 (K+,OH-) 10.000 0 0 0.2
(K+,OH-) CO2 -2.000 0 0 0.2
CO2 (K+,CO32-) 10.000 0 0 0.2
(K+,CO32-) CO2 -2.000 0 0 0.2
CO2 (K+,HCO3-) 10.000 0 0 0.2
+
(K ,HCO3-) CO2 -2.000 0 0 0.2
CO2 (K+,HS-) 10.000 0 0 0.2
(K+,HS-) CO2 -2.000 0 0 0.2
CO2 (K+,S2-) 10.000 0 0 0.2
(K+,S2-) CO2 -2.000 0 0 0.2
H2S (H3O+,OH-) 15.000 0 0 0.1
(H3O+,OH-) H2S -8.000 0 0 0.1
H2S (H3O+,CO32-) 15.000 0 0 0.1
(H3O+,CO32-) H2S -8.000 0 0 0.1
H2S (H3O+,HCO3-) 15.000 0 0 0.1
(H3O+,HCO3-) H2S -8.000 0 0 0.1
H2S (H3O+,HS-) 15.000 0 0 0.1
(H3O+,HS-) H2S -8.000 0 0 0.1
H2S (H3O+,S2-) 15.000 0 0 0.1
(H3O+,S2-) H2S -8.000 0 0 0.1
H2S (K+,OH-) 10.000 0 0 0.2
(K+,OH-) H2S -2.000 0 0 0.2
H2S (K+,CO32-) 10.000 0 0 0.2
(K+,CO32-) H2S -2.000 0 0 0.2
H2S (K+,HCO3-) 10.000 0 0 0.2
(K+,HCO3-) H2S -2.000 0 0 0.2
H2S (K+,HS-) 10.000 0 0 0.2
(K+,HS-) H2S -2.000 0 0 0.2
H2S (K+,S2-) 10.000 0 0 0.2
(K+,S2-) H2S -2.000 0 0 0.2

A-39
Appendix C: Electrolyte NRTL Adjustable Parameters

Table C.1.2: The Electrolyte NRTL adjustable parameters used in this work.

Component i Component j A B C D
H2O (H3O+,OH-) 8.045 0 0 0.2
(H3O+,OH-) H2O -4.072 0 0 0.2
H2O (H3O+,CO32-) 8.045 0 0 0.2
(H3O+,CO32-) H2O -4.072 0 0 0.2
H2O (H3O+,HCO3-) 8.045 0 0 0.2
(H3O+,HCO3-) H2O -4.072 0 0 0.2
H2O (H3O+,HS-) 8.045 0 0 0.2
(H3O+,HS-) H2O -4.072 0 0 0.2
H2O (H3O+,S2-) 8.045 0 0 0.2
(H3O+,S2-) H2O -4.072 0 0 0.2
H2O (K+,OH-) 7.841 773.360 0 0.2
(K+,OH-) H2O -4.259 -305.651 0 0.2
H2O (K+,CO32-) -5.020 a -250.640 a 0a 0.2
(K+,CO32-) H2O -0.176 a -864.400 a 0a 0.2
H2O (K+,HCO3-) 6.250 a 0a 0a 0.2
(K+,HCO3-) H2O -3.728 a 0a 0a 0.2
H2O (K+,HS-) 3.076 a 0a 0a 0.2
(K+,HS-) H2O -3.253 a -226.148 a 0a 0.2
H2O (K+,S2-) 3.438 a 4206.772 a 0a 0.2
(K+,S2-) H2O -6.305 a -267.739 a 0a 0.2
CO2 H2O 10.064 -3268.135 0 0.2
H2O CO2 10.064 -3268.135 0 0.2
H2S H2O -3.674 1155.9 0 0.2
H2O H2S -3.674 1155.9 0 0.2
a These values were regressed from literature data.

A-40
Appendix C: Electrolyte NRTL Adjustable Parameters

C.2 Data Regression Procedure


The Electrolyte NRTL adjustable parameter values not taken from the Aspen Properties® databanks
were determined using the Aspen Properties® Data Regression System (DRS). The adjustable
+ - +
parameter values for the H2O-(K ,HCO3 ) and H2O-(K ,CO32-) binary interaction pairs were first
determined from Tosh and co-workers’ (1959) data for CO2 solubility in aqueous potassium carbonate
solutions. Adjustable parameter values for the H2O-(K+,HS-) and H2O-(K+,S2-) binary interaction pairs
were then determined using Tosh and co-workers’ (1960) H2S solubility data in aqueous potassium
carbonate solutions. Both sets of data regressions were performed using the general procedure
outlined below.

1. Where available, the Aspen Properties® databank values were set as the initial values for
the adjustable parameters to be regressed from the data. Otherwise, the following default
values were applied:
Binary Interaction Pair A B C
Water-Electrolyte 8 0 0
Electrolyte-Water -4 0 0

2. The full set of twelve adjustable parameters was regressed from the data. The resulting set
of parameter values was labelled “Full”.

3. One of the four C parameters (designated as C1) was excluded from regression (i.e. it was
fixed at its initial value) and the remaining adjustable parameters were regressed from the
data. The resulting set of parameter values was labelled “C1”. Another C parameter (C2)
was then excluded instead of C1 to give the parameter value set “C2”. This step was
repeated with the remaining C parameters (C3 and C4) to give the parameter values sets
“C3” and “C4”.

4. C1 and C2 were excluded from regression and the remaining adjustable parameters were
regressed from the data. The resulting set of parameter values was labelled “C12”. C3 was
then excluded instead of C2 to give the parameter value set “C13”. C4 was then excluded
instead of C3 to give the parameter value set “C14”.

5. C2 and C3 were excluded from regression and the remaining adjustable parameters were
regressed from the data. The resulting set of parameter values was labelled “C23”. C4 was
then excluded instead of C3 to give the parameter value set “C24”.

6. C3 and C4 were excluded from regression and the remaining adjustable parameters were
regressed from the data. The resulting set of parameter values was labelled “C34”.

A-41
Appendix C: Electrolyte NRTL Adjustable Parameters

7. C1, C2 and C3 were excluded from regression and the remaining adjustable parameters
were regressed from the data. The resulting set of parameter values was labelled “C123”.
C4 was then excluded instead of C3 to give the parameter value set “C124”. C3 was then
excluded instead of C2 to give the parameter value set “C134”.

8. C2, C3 and C3 were excluded from regression and the remaining adjustable parameters
were regressed from the data. The resulting set of parameter values was labelled “C234”.

9. All four C parameters were excluded from regression and the remaining adjustable
parameters were regressed from the data. The resulting set of parameter values was
labelled “C”.

10. Excluding all four C parameters from regression, steps 3 to 9 were repeated with the four B
parameters (B1, B2, B3 and B4) to give the parameter value sets “B1C”, “B2C”, “B3C”,
“B4C”, “B12C”, “B13C”, “B14C”, “B123C”, “B124C”, “B134C”, “B234C” and “BC”.

11. Excluding all C and B parameters from regression, steps 3 to 9 were repeated with the four
A parameters (A1, A2, A3 and A4) to give the parameter value sets “A1BC”, “A2BC”,
“A3BC”, “A4BC”, “A12BC”, “A13BC”, “A14BC”, “A123BC”, “A124BC”, “A134BC”, “A234BC”
and “ABC”.

12. Each simplified set of parameter values obtained from steps 3 to 11 was compared against
the “Full” parameter value set via an F-Test to determine the reliability of the simplified set.
The corresponding one-tail p-value was also calculated to determine the probability of
observing the F-Test value if the null hypothesis was true.
Null hypothesis: The simplified set fits the data better than the “Full” set.
F  1 or p > 0.05

where
WSSQ Simplified  WSSQ Full
df Simplified  df Full
F (C.2.1)
WSSQ Full
dfFull

p
f F, df Simplified  dfFull , dfFull (C.2.2)

df NData  NParameters (C.2.3)

WSSQFull and WSSQSimplified are the weighted sum of squares for the full and simplified
parameter value sets, respectively, dfFull and dfSimplified are the corresponding degrees of
freedom, NData is the number of data points used in the regression, and NParameters is the
number of parameters regressed.

A-42
Appendix C: Electrolyte NRTL Adjustable Parameters

13. The simplified parameter value sets which passed the above F-Test were sorted according
to their weighted sum of squares. A logic test was then performed on each parameter set to
tabulate the number of parameters with a standard error greater than the associated
regressed value. The parameter value set with the lowest logic test result and the lowest
weighted sum of squares was labelled the optimal set.

A-43
Appendix C: Electrolyte NRTL Adjustable Parameters

C.3 Data Regression Results


The statistical results for the data regression runs for the CO2-K2CO3-KHCO3-H2O system are given in
Table C.3.1. The weighted sum of squares (WSSQ), residual root mean square error (RRMSQE),
degrees of freedom (df), F-value and p-value are listed for each parameter set. The data set used in
the regression runs consisted of 120 data points, and the number of parameters regressed ranged
from 0 to 12, depending on the parameter set. Using the F-Test, 25 simplified parameter sets were
identified as being more suitable than the “Full” set. These are sorted in ascending order according to
their WSSQ in Table C.3.1. The corresponding logic test results are also presented.

Of the 25 simplified parameter sets in Table C.3.1, set “B34C” was the only set to have all its
regressed parameter values greater than the associated standard error. Consequently, it was
selected as the optimal parameter value set and was used in the next series of data regression runs
for the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system. The regressed parameter values and standard
errors associated with this optimal set are given in Table 3.2.2.

Table C.3.3 presents the statistical results for the data regression runs for the CO2-H2S-K2CO3-
KHCO3-KHS-K2S-H2O system. The data set used in this series of regression runs consisted of 127
data points, and the number of parameters regressed ranged from 0 to 12, depending on the
parameter set. Only 6 simplified parameter sets were determined to be more suitable than the “Full”
set via the F-Test, and these are sorted in ascending order according to their WSSQ in Table C.3.4.
The corresponding logic test results are also presented.

Of the 6 simplified parameter sets in Table C.3.4, set “B1C” was the only set to have all its regressed
parameter values greater than the associated standard error. It was therefore selected as the optimal
parameter value set for the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system. The regressed parameter
values and standard errors associated with this optimal set are given in Table 3.2.3.

A-44
Appendix C: Electrolyte NRTL Adjustable Parameters

Table C.3.1: Statistical results for the CO2-K2CO3-KHCO3-H2O system data regression runs.

Parameter Set WSSQ RRMSQE df F p Null Hypothesis


Full 1143.4 3.254 108 - - -
C1 1378.6 3.573 109 22.21 7.3×10-6 False
C2 1223.8 3.366 109 7.59 6.9×10-3 False
C3 1129.1 3.233 109 -1.36 n/a a True
C4 1160.7 3.278 109 1.63 2.0×10-1 True
C12 1230.6 3.376 110 4.12 1.9×10-2 False
C13 1279.8 3.442 110 6.44 2.3×10-3 False
C14 1230.8 3.376 110 4.12 1.9×10-2 False
C23 1199.3 3.332 110 2.64 7.6×10-2 True
C24 1100.3 3.192 110 -2.03 n/a a True
C34 1198.3 3.331 110 2.59 8.0×10-2 True
C123 1178.2 3.303 111 1.10 3.5×10-1 True
C124 1214.5 3.353 111 2.24 8.8×10-2 True
C134 1192.0 3.322 111 1.53 2.1×10-1 True
C234 1191.1 3.321 111 1.50 2.2×10-1 True
C 1188.6 3.317 112 1.07 3.8×10-1 True
B1C 1105.0 3.199 113 -0.73 n/a a True
B2C 1178.2 3.303 113 0.66 6.6×10-1 True
B3C 1178.4 3.303 113 0.66 6.5×10-1 True
B4C 1206.7 3.343 113 1.19 3.2×10-1 True
B12C 1142.6 3.253 114 -0.01 n/a a True
B13C 1180.1 3.306 114 0.58 7.5×10-1 True
B14C 1198.3 3.331 114 0.86 5.2×10-1 True
B23C 1194.0 3.325 114 0.80 5.7×10-1 True
B24C 1207.2 3.343 114 1.00 4.3×10-1 True
B34C 1201.7 3.336 114 1.04 4.0×10-1 True
B123C 1194.9 3.326 115 0.69 6.8×10-1 True
B124C 1198.1 3.331 115 0.74 6.4×10-1 True
B134C 1204.7 3.340 115 0.83 5.7×10-1 True
B234C 1195.4 3.327 115 0.70 6.7×10-1 True
BC 1195.8 3.328 116 0.62 7.6×10-1 True
A1BC 1471.3 3.691 117 3.44 9.2×10-4 False
A2BC 1467.2 3.686 117 3.40 1.0×10-3 False
A3BC 7246.8 8.191 117 64.05 3.2×10-39 False
A4BC 3356.2 5.575 117 23.22 1.5×10-21 False
A12BC 1889.7 4.183 118 7.05 1.9×10-8 False
A13BC 8143.4 8.683 118 66.12 2.1×10-41 False
A14BC 4755.4 6.636 118 34.12 5.4×10-29 False
A23BC 1577.1 3.821 118 4.10 8.5×10-5 False
A24BC 6965.0 8.031 118 54.99 8.9×10-38 False
A34BC 8706.5 8.979 118 71.44 6.0×10-43 False
A123BC 9897.8 9.573 119 75.17 2.1×10-45 False
A124BC 10029.4 9.637 119 76.30 1.0×10-45 False
A134BC 9693.6 9.474 119 73.42 6.4×10-45 False
A234BC 9920.5 9.584 119 75.37 1.9×10-45 False
ABC 10032.7 9.638 120 69.97 3.2×10-45 False
a
No p-value was calculated as the F-value was negative due to the WSSQ being smaller than that of the “Full”
parameter set, indicating the null hypothesis is true.

A-45
Appendix C: Electrolyte NRTL Adjustable Parameters

Table C.3.2: Suitable parameter value sets for the CO2-K2CO3-KHCO3-H2O system.

Parameter Set WSSQ Logic Test


C24 1100.3 7
B1C 1105.0 3
C3 1129.1 9
B12C 1142.6 2
FULL 1143.4 8
C4 1160.7 8
C123 1178.2 6
B2C 1178.2 5
B3C 1178.4 5
B13C 1180.1 4
C 1188.6 5
C234 1191.1 6
C134 1192.0 6
B23C 1194.0 3
B123C 1194.9 2
B234C 1195.4 2
BC 1195.8 1
B124C 1198.1 3
C34 1198.3 7
B14C 1198.3 3
C23 1199.3 7
B34C 1201.7 0
B134C 1204.7 2
B4C 1206.7 4
B24C 1207.2 4
C124 1214.5 6

A-46
Appendix C: Electrolyte NRTL Adjustable Parameters

Table C.3.3: Statistical results for the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system data regression runs.

Parameter Set WSSQ RRMSQE df F p Null Hypothesis


Full 2006.6 4.177 115 - - -
C1 3221.7 5.293 116 69.63 1.8×10-13 False
C2 1962.6 4.131 116 -2.53 n/a a True
C3 3225.1 5.296 116 69.83 1.7×10-13 False
C4 3850.9 5.787 116 105.69 5.6×10-18 False
C12 1985.1 4.155 117 -0.62 n/a a True
C13 3860.6 5.794 117 53.12 4.6×10-17 False
C14 3860.6 5.794 117 53.12 4.6×10-17 False
C23 1981.6 4.151 117 -0.72 n/a a True
C24 3224.0 5.295 117 34.88 1.4×10-12 False
C34 3851.0 5.787 117 52.85 5.3×10-17 False
C123 1995.6 4.166 118 -0.21 n/a a True
C124 1991.8 4.162 118 -0.28 n/a a True
C134 3860.6 5.794 118 35.42 2.8×10-16 False
C234 3227.9 5.298 118 23.33 7.2×10-12 False
C 3259.7 5.324 119 17.95 1.8×10-11 False
B1C 2014.7 4.186 120 0.09 9.9×10-1 True
B2C 3314.8 5.369 120 14.99 2.6×10-11 False
B3C 3892.1 5.818 120 21.61 3.4×10-15 False
B4C 3892.1 5.818 120 21.61 3.4×10-15 False
B12C 3363.3 5.408 121 12.96 3.8×10-11 False
B13C 3980.0 5.883 121 18.85 3.5×10-15 False
B14C 3980.0 5.883 121 18.85 3.5×10-15 False
B23C 3319.9 5.373 121 12.54 7.7×10-11 False
B24C 3317.9 5.371 121 12.52 7.9×10-11 False
B34C 3892.1 5.818 121 18.01 1.2×10-14 False
B123C 4035.0 5.923 122 16.61 5.5×10-15 False
B124C 3366.4 5.410 122 11.13 1.1×10-10 False
B134C 3980.0 5.883 122 16.16 1.2×10-14 False
B234C 3689.0 5.664 122 13.77 7.5×10-13 False
BC 3718.5 5.686 123 12.26 1.4×10-12 False
A1BC 3898.5 5.822 124 12.05 3.1×10-13 False
A2BC 4906.3 6.532 124 18.46 1.1×10-18 False
A3BC 4027.1 5.918 124 12.87 5.4×10-14 False
A4BC 4031.1 5.921 124 12.89 5.1×10-14 False
A12BC 255230.6 47.110 125 1451.21 5.2×10-116 False
A13BC 4445.9 6.218 125 13.98 7.1×10-16 False
A14BC 4466.3 6.232 125 14.10 5.5×10-16 False
A23BC 5648.7 7.009 125 20.87 1.4×10-21 False
A24BC 5730.4 7.059 125 21.34 6.2×10-22 False
A34BC 22110.0 13.866 125 115.21 4.4×10-55 False
A123BC 255249.6 47.112 126 1319.39 1.8×10-115 False
A124BC 255255.3 47.113 126 1319.41 1.8×10-115 False
A134BC 24056.8 14.463 126 114.88 1.2×10-56 False
A234BC 23051.6 14.158 126 109.64 1.4×10-55 False
ABC 268481.8 48.318 127 1272.63 3.5×10-116 False
a
No p-value was calculated as the F-value was negative due to the WSSQ being smaller than that of the “Full”
parameter set, indicating the null hypothesis is true.

A-47
Appendix C: Electrolyte NRTL Adjustable Parameters

Table C.3.4: Suitable parameter value sets for the CO2-H2S-K2CO3-KHCO3-KHS-K2S-H2O system.

Parameter Set WSSQ Logic Test


C2 1962.6 4
C23 1981.6 7
C12 1985.1 5
C124 1991.8 6
C123 1995.6 7
FULL 2006.6 12
B1C 2014.7 0

A-48
Appendix D: Aspen Custom Modeler® Simulation Results

APPENDIX D

ASPEN CUSTOM MODELER®


SIMULATION RESULTS

This appendix presents the results for the Aspen Custom Modeler® simulations. The column profiles
generated from the preliminary column simulations are provided; along with the column profiles
produced by the Aspen Custom Modeler® CO2 train process models. Also included is a table of
alternative correlations for estimating the mass transfer coefficients and effective interfacial area.

A-49
Appendix D: Aspen Custom Modeler® Simulation Results

D.1 The Different Modelling Approaches


The absorber CO2 and H2S vapour phase profiles and the regenerator liquid phase CO2 and H2S
loading profiles for CO2 trains #2 to #6 are presented below for the three different modelling
approaches.

Model 1 Model 2 Model 3 Data

Cond
1.2
1.0 WS

Fraction of Total Packed Height


Fraction of Total Packed Height

1.0

0.8
0.8

0.6 0.6

0.4 0.4

0.2
0.2

0.0
0.0 Reb
0 5 10 15 20 0 5 10 15 20 25 -0.2
CO2 (mol%) H2S (ppm) 0.3 0.4 0.5 0.6 0.7 0 10 20 30

CO2 Loading H2S Loading (×106)

(a) (d)
1.2
Cond
1.0 WS
Fraction of Total Packed Height

Fraction of Total Packed Height

1.0

0.8
0.8

0.6 0.6

0.4 0.4

0.2
0.2

0.0
0.0 Reb
0 5 10 15 20 0 5 10 15 20 25 -0.2
CO2 (mol%) H2S (ppm) 0.3 0.4 0.5 0.6 0.7 0 5 10 15 20 25 30

CO2 Loading H2S Loading (×106)

(b) (e)
Cond
1.2
1.0 WS
Fraction of Total Packed Height

Fraction of Total Packed Height

1.0
0.8
0.8

0.6
0.6

0.4 0.4

0.2 0.2

0.0
0.0 Reb
0 5 10 15 20 0 5 10 15 20 25
-0.2
CO2 (mol%) H2S (ppm) 0.3 0.4 0.5 0.6 0.7 0 5 10 15 20 25 30

CO2 Loading H2S Loading (×106)

(c) (f)
Figure D.1.1: Results of the different modelling approaches for CO2 trains #2 to #4. Absorber CO2 and
H2S vapour phase profiles for (a) train #2, (b) train #3 and (c) train #4. Regenerator liquid phase CO2 and
H2S loading profiles for (d) train #2, (e) train #3 and (f) train #4 (Cond = condenser, WS = wash section,
Reb = reboiler).

Model 1 Model 2 Model 3 Data

A-50
Appendix D: Aspen Custom Modeler® Simulation Results

Cond
1.2
1.0 WS

Fraction of Total Packed Height

Fraction of Total Packed Height


1.0
0.8
0.8

0.6
0.6

0.4 0.4

0.2
0.2

0.0
0.0 Reb
0 5 10 15 20 0 5 10 15 20 25 -0.2
CO2 (mol%) H2S (ppm) 0.3 0.4 0.5 0.6 0.7 0 5 10 15 20 25

CO2 Loading H2S Loading (×106)

(a) (c)
Cond
1.2
1.0 WS
Fraction of Total Packed Height

Fraction of Total Packed Height


1.0

0.8
0.8

0.6
0.6

0.4 0.4

0.2 0.2

0.0
0.0 Reb
0 5 10 15 20 0 5 10 15 20 25 -0.2
CO2 (mol%) H2S (ppm) 0.3 0.4 0.5 0.6 0.7 0 5 10 15 20 25 30

CO2 Loading H2S Loading (×106)

(b) (d)
Figure D.1.2: Results of the different modelling approaches for CO2 trains #5 and #6. Absorber CO2 and
H2S vapour phase profiles for (a) train #5 and (b) train #6. Regenerator liquid phase CO2 and H2S loading
profiles for (c) train #5 and (d) train #6 (Cond = condenser, WS = wash section, Reb = reboiler).

A-51
Appendix D: Aspen Custom Modeler® Simulation Results

D.2 Model Adjustments


D.2.1 Liquid Phase Enthalpy Correction
The temperature profiles predicted by Model 2 for the absorber and regenerator columns in CO2 trains
#2 to #6 are presented below. The effect of the liquid phase enthalpy correction on the absorber
temperature profiles is also included.

Vapour Phase Liquid Phase Vapour Data Liquid Data Uncorrected enthalpy Corrected enthalpy Data

1.2
Cond
WS 1.0
Fraction of Total Packed Height

Fraction of Total Packed Height


1.0

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
-0.2 30 50 70 90 110 130 100 110 120 130
30 50 70 90 110 130 60 70 80 90 100 110 120
Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)
Absorber Temperature (°C) Regenerator Temperature (°C)

(a) (d)
1.2
Cond
WS 1.0
Fraction of Total Packed Height

Fraction of Total Packed Height

1.0

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
-0.2
90 100 110 120 130 100 110 120 130
90 100 110 120 130 60 70 80 90 100 110 120
Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)
Absorber Temperature (°C) Regenerator Temperature (°C)

(b) (e)
1.2
Cond
WS 1.0
Fraction of Total Packed Height

Fraction of Total Packed Height

1.0

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
-0.2
90 100 110 120 130 100 110 120 130
90 100 110 120 130 60 70 80 90 100 110 120
Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)
Absorber Temperature (°C) Regenerator Temperature (°C)

(c) (f)
Figure D.2.1: Temperature profiles for CO2 trains #2 to #4. Temperature profiles predicted by Model 2 for
(a) train #2, (b) train #3 and (c) train #4 (Cond = condenser, WS = wash section, Reb = reboiler). Effect of
the liquid phase enthalpy correction on the absorber temperatures for (d) train #2, (e) train #3 and (f) train
#4.

A-52
Appendix D: Aspen Custom Modeler® Simulation Results

Vapour Phase Liquid Phase Vapour Data Liquid Data Uncorrected enthalpy Corrected enthalpy Data

Cond
1.2
WS 1.0
Fraction of Total Packed Height

Fraction of Total Packed Height


1.0

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
-0.2 90 100 110 120 130 100 110 120 130
90 100 110 120 130 60 70 80 90 100 110 120
Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)
Absorber Temperature (°C) Regenerator Temperature (°C)

(a) (c)
Cond
1.2
WS 1.0
Fraction of Total Packed Height

Fraction of Total Packed Height


1.0

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
-0.2 90 100 110 120 130 100 110 120 130
90 100 110 120 130 60 70 80 90 100 110 120
Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)
Absorber Temperature (°C) Regenerator Temperature (°C)

(b) (d)
Figure D.2.2: Temperature profiles for CO2 trains #5 and #6. Temperature profiles predicted by Model 2
for (a) train #5 and (b) train #6 (Cond = condenser, WS = wash section, Reb = reboiler). Effect of the
liquid phase enthalpy correction on the absorber temperatures for (c) train #5 and (d) train #6.

A-53
Appendix D: Aspen Custom Modeler® Simulation Results

D.2.2 Sensitivity to Model Parameters


Table D.2.1 lists some alternative correlations for estimating the mass transfer coefficients and the
effective interfacial area.

Table D.2.1: Alternative mass transfer coefficient and effective interfacial area correlations.

Mass Transfer Coefficient Correlations


0. 8 1
DGj § UG ˜ v G · § PG · 3
k Gj 0 .2 ˜ ˜¨ ¸ ˜¨ ¸
Van Krevelen and Hoftijzer dN ¨© a ˜ PG ¸¹ ¨ UG ˜ DGj ¸
© ¹
(D.2.1)
(1948) 1 2 1
§U 2 ˜ g· 3
§U ˜v · 3 § PL · 3
kLj 0.015 ˜ DLj ˜ ¨ L 2 ¸ ˜ ¨¨ L L ¸¸ ˜¨ ¸
¨ P ¸ © aI ˜ PL ¹ ¨ UL ˜ DLj ¸
© L ¹ © ¹
0. 8 2
§ dp ˜ UG ˜ v G · § UG ˜ DGj · 3
k Gj 1.195 ˜ v G ˜ ¨¨ ¸ ˜ ¨¨ ¸
© PG ˜ 1  I ¹
¸ ¸
© PG ¹
Shulman and co-workers (1955) 0 .5
(D.2.2)
0.45
DLj § dp ˜ UL ˜ vL · § PL ·
kLj 25.1 ˜ ˜¨ ¸ ˜¨ ¸
dp ¨© aI ˜ PL ¸¹ ¨ UL ˜ DLj ¸
© ¹
0 .5 0.75 1
§ a · §U ˜v · § PG · 3
k Gj EG ˜ DGj ˜ ¨¨ ¸ ˜ ¨¨ G G ¸¸ ˜¨ ¸
d
© h ˜ I  I
L ¹
¸
© a ˜ PG ¹ ¨ UG ˜ DGj ¸
© ¹
Billet and Schultes (1999) (D.2.3)
1 0 .5 1
§U ˜ g· 6 § DLj · §v · 3
kLj EL ˜ ¨¨ L ¸¸ ˜ ¨¨ ¸
¸ ˜¨ L ¸
© PL ¹ © dh ¹ © a ¹

Effective Interfacial Area Correlations


0.041 0.133 0.182
aI §U ˜v · §U ˜v 2 · §V ·
Puranik and Vogelpohl (1974) 1.045 ˜ ¨¨ L L ¸¸ ˜¨ L L ¸ ˜ ¨¨ c ¸¸ (D.2.4)
a © a ˜ PL ¹ ¨ a ˜ VL ¸ © VL ¹
© ¹
0.49 0.196
§ U ˜g · §v 2 ˜a·
Kolev (1976)
aI
0.583 ˜ ¨¨ 2L ¸
¸ ˜¨ L
¨ g ¸
¸
˜ a ˜ dp 0.42 (D.2.5)
a © a ˜ VL ¹ © ¹
0.392
aI §v ˜P U ˜v · VL
0.5
Bravo and Fair (1982) 19.76 ˜ ¨¨ L L ˜ G G ¸¸ ˜ (D.2.6)
© VL a ˜ PG ¹ 0.4
a Hc
0.2 0.75 0.45
aI 1 .5 §U ˜v ˜d · § U ˜ v 2 ˜ dh · § v 2 ·
Billet and Schultes (1999) ˜ ¨¨ L L h ¸¸ ˜¨ L L ¸ ˜¨ L ¸ (D.2.7)
a a ˜ dh 0.5 © PL ¹ ¨
© V L
¸
¹
¨ g ˜ dh ¸
© ¹
Note: All variables are in SI units. Notation: Dj is the component diffusivity;  is the mass density;  is the viscosity; a is the
packing specific surface area; v is the phase flow velocity; g is the gravitational constant; dN is the nominal packing size; dp
is the packing particle diameter; dh is the packing hydraulic diameter; I is the voidage; IL is the liquid phase volumetric
holdup; E is a packing-specific constant; VL is the surface tension; Vc is the critical surface tension parameter; Hc is the
column height; Dc is the column diameter; G and L are the vapour and liquid phase molar flow rates; Ct is the molar density;
and the subscripts G and L denote the vapour and liquid phases.

A-54
Appendix D: Aspen Custom Modeler® Simulation Results

D.2.3 Effective Interfacial Area Adjustment Factor


The effect of the effective interfacial area adjustment factor on the absorber CO2 and H2S vapour
phase profiles for CO2 trains #2 to #6 are presented below.

Unadjusted Model Adjusted Model Data

1.0 1.0
Fraction of Total Packed Height

Fraction of Total Packed Height


0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
0 5 10 15 20 0 5 10 15 20 25 0 5 10 15 20 0 5 10 15 20 25

CO2 (mol%) H2S (ppm) CO2 (mol%) H2S (ppm)

(a) (b)
1.0 1.0
Fraction of Total Packed Height

Fraction of Total Packed Height


0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
0 5 10 15 20 0 5 10 15 20 25 0 5 10 15 20 0 5 10 15 20 25
CO2 (mol%) H2S (ppm) CO2 (mol%) H2S (ppm)

(c) (d)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 5 10 15 20 25

CO2 (mol%) H2S (ppm)

(e)
Figure D.2.3: Effect of the effective interfacial area adjustment factor on the absorber CO2 and H2S vapour
phase profiles. (a) CO2 train #2. (b) CO2 train #3. (c) CO2 train #4. (d) CO2 train #5. (e) CO2 train #6.

A-55
Appendix D: Aspen Custom Modeler® Simulation Results

D.3 CO2 Train Model Validation


The vapour and liquid phase composition profiles and the temperature profiles for the absorber and
regenerator columns in the Aspen Custom Modeler® models for CO2 trains #2 to #6 are presented
below.

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
Cond
1.2
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8 0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
1.2
Cond
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0
Fraction of Total Packed Height

1.0
Fraction of Total Packed Height

0.8 0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0

0.0 Reb
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
Cond
1.2
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0
Fraction of Total Packed Height

1.0
Fraction of Total Packed Height

0.8 0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(c)
Figure D.3.1: CO2 and H2S vapour and liquid phase column profiles for the first set of plant data.
(a) CO2 train #2. (b) CO2 train #3. (c) CO2 train #4. (Cond = condenser, WS = wash section, Reb =
reboiler)

A-56
Appendix D: Aspen Custom Modeler® Simulation Results

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05 1.2
Cond
1.0 WS

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
0.0 Reb
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.3 0.4 0.5 0.6
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
1.2
Cond
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8 0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Figure D.3.2: CO2 and H2S vapour and liquid phase column profiles for the first set of plant data.
(a) CO2 train #5. (b) CO2 train #6. (Cond = condenser, WS = wash section, Reb = reboiler)

A-57
Appendix D: Aspen Custom Modeler® Simulation Results

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05

0 5 10 15 20 0.E+00 1.E-05 2.E-05 3.E-05 4.E-05 1.2


Cond
1.0 WS

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0

0.0 Reb
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05

0 5 10 15 20 0.E+00 2.E-05 4.E-05 6.E-05 1.2


Cond
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0

0.0 Reb
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05
Cond
1.2
0 5 10 15 20 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0
Fraction of Total Packed Height

1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(c)
Figure D.3.3: CO2 and H2S vapour and liquid phase column profiles for the second set of plant data.
(a) CO2 train #2. (b) CO2 train #3. (c) CO2 train #4. (Cond = condenser, WS = wash section, Reb =
reboiler)

A-58
Appendix D: Aspen Custom Modeler® Simulation Results

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05

0 5 10 15 20 0.E+00 1.E-05 2.E-05 3.E-05 4.E-05 1.2


Cond
1.0 WS

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0

0.0 Reb
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05
1.2
Cond
0 5 10 15 20 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8 0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Figure D.3.4: CO2 and H2S vapour and liquid phase column profiles for the second set of plant data.
(a) CO2 train #5. (b) CO2 train #6. (Cond = condenser, WS = wash section, Reb = reboiler)

A-59
Appendix D: Aspen Custom Modeler® Simulation Results

Vapour Phase Liquid Phase Vapour Data Liquid Data

1.2
Cond 1.2
Cond
WS WS

Fraction of Total Packed Height


Fraction of Total Packed Height 1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
30 50 70 90 110 130 60 70 80 90 100 110 120 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(a) (b)
1.2
Cond 1.2
Cond
WS WS
Fraction of Total Packed Height

Fraction of Total Packed Height


1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
90 100 110 120 130 60 70 80 90 100 110 120 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(c) (d)
1.2
Cond
WS
Fraction of Total Packed Height

1.0

0.8

0.6

0.4

0.2

0.0
Reb
-0.2
90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C)

(e)
Figure D.3.5: Column temperature profiles for the first set of plant data. (a) CO2 train #2. (b) CO2 train #3.
(c) CO2 train #4. (d) CO2 train #5. (e) CO2 train #6. (Cond = condenser, WS = wash section, Reb =
reboiler)

A-60
Appendix D: Aspen Custom Modeler® Simulation Results

Vapour Phase Liquid Phase Vapour Data Liquid Data

1.2
Cond Cond
1.2
WS WS

Fraction of Total Packed Height


Fraction of Total Packed Height 1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
30 50 70 90 110 130 60 70 80 90 100 110 120 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(a) (b)
1.2
Cond 1.2
Cond
WS WS
Fraction of Total Packed Height

Fraction of Total Packed Height


1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
90 100 110 120 130 60 70 80 90 100 110 120 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(c) (d)
1.2
Cond
WS
Fraction of Total Packed Height

1.0

0.8

0.6

0.4

0.2

0.0
Reb
-0.2
90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C)

(e)
Figure D.3.6: Column temperature profiles for the second set of plant data. (a) CO2 train #2. (b) CO2 train
#3. (c) CO2 train #4. (d) CO2 train #5. (e) CO2 train #6. (Cond = condenser, WS = wash section, Reb =
reboiler)

A-61
Appendix E: Hypothetical K2CO3* HYSYS® Component Properties

APPENDIX E

HYPOTHETICAL K2CO3* HYSYS®


COMPONENT PROPERTIES

This appendix presents the various physical and thermodynamic properties required for the creation of
the hypothetical K2CO3* HYSYS® component.

A-62
Appendix E: Hypothetical K2CO3* HYSYS® Component Properties

E.1 Base Properties


E.1.1 Normal Boiling Point
When the normal boiling point is unknown for a hypothetical component, HYSYS® uses a proprietary
method to estimate this property (Hyprotech Ltd, 2001a). In the case of the hypothetical K2CO3*
component, the potassium carbonate melting point of 900.85°C (Chase, 1998) was used as the
hypothetical component’s normal boiling point since potassium carbonate does not boil, but instead
begins to decompose at temperatures close to its melting point.

E.1.2 Molecular Weight


HYSYS® has two default methods for estimating the molecular weight of hypothetical components:
the Bergman method (Bergman et al., 1975) and the Lee-Kesler model (Kesler and Lee, 1976). The
former method is applied only to components with boiling points less than 155°F (68.3°C) and is based
on tabulated molecular weights of gas condensate carbon number fractions from C5 to C45 that were
obtained from a study of US gas condensate mixtures by Bergman and co-workers (1975). The latter
estimation method is used for all other hypothetical components and gives the molecular weight MW
as a function of the boiling point Tb in °R and the specific gravity SG:
MW 12272.6  9486.4 ˜ SG  4.6523  3.3287 ˜ SG ˜T b

§
 1  0.77084 ˜ SG  0.02058 ˜ SG 2 ˜ ¨¨1.3437 
720.79 · 10 7
¸˜
Tb ¸¹ Tb
(E.1.1)
©

§
 1  0.80882 ˜ SG  0.02226 ˜ SG 2 ˜ ¨¨1.8828 
181.9 · 10 12
¸˜
Tb ¸¹ Tb 3
©

It should be noted that SG refers to 60°F/60°F specific gravity, i.e. the mass ratio of equal volumes of
the component of interest and water at 60°F.

For the hypothetical K2CO3* component, neither estimation method was used since the molecular
weight was taken to be that of potassium carbonate: 138.2058 kg/kmol (Chase, 1998).

E.1.3 Ideal Liquid Density


The default estimation method for determining the ideal liquid density of a hypothetical component in
HYSYS® is the following correlation proposed by Yen and Woods (1966):
j
Vc 4
§ T ·3
VLs
1 ¦
j 1
K j ˜ ¨¨1 
©
¸
Tc ¸¹
(E.1.2)

2 3
K1 17.4425  214.578 ˜ Z c  989.625 ˜ Z c  1522.06 ˜ Z c (E.1.3)
2 3
K2 3.28257  13.6377 ˜ Z c  107.4844 ˜ Z c  384.211 ˜ Z c if Z c d 0.26 (E.1.4)
2 3
K2 60.2091  402.063 ˜ Z c  501.0 ˜ Z c  641.0 ˜ Z c if Z c ! 0.26 (E.1.5)

A-63
Appendix E: Hypothetical K2CO3* HYSYS® Component Properties

K3 0 (E.1.6)

K4 0.93  K 2 (E.1.7)

where Vc is the critical molar volume, VLs is the saturated liquid molar volume, T is the temperature, Tc
is the critical temperature, and Zc is the critical compressibility.

For the hypothetical K2CO3* component, the ideal liquid density was not estimated by the above
method, but was instead taken to be 2423.11 kg/m3, the density of solid potassium carbonate (Perry
and Green, 1997).

E.1.4 Critical Temperature and Critical Pressure


Three default methods are available in HYSYS® for estimating the critical temperatures and critical
pressures of hypothetical components: the Lee-Kesler model (Kesler and Lee, 1976), the Bergman
method (Bergman, 1976) and the Cavett relations (Cavett, 1964). The first method is used for
components with liquid densities greater than 1067 kg/m3 or boiling points above 800 K, and consists
of the following correlations:

10 5
Tc 341.7  811 ˜ SG  0.4244  0.1174 ˜ SG ˜T b  0.4669  3.2623 ˜ SG ˜ (E.1.8)
Tb

0.0566 § 2.2898 0.11857 ·


ln Pc 8.3634   ¨¨ 0.24244   ¸¸ ˜ 10 3 ˜T b
SG © SG SG 2 ¹
(E.1.9)
§ 3.648 0.47227 · 2 § 1.6977 ·
¸ ˜ 10 7 ˜T b ¨¨ 0.42019  ¸ ˜ 10 10 ˜T b
3
 ¨¨1.4685   2 ¸ 2 ¸
© SG SG ¹ © SG ¹

where the critical temperature Tc and the boiling point Tb are in °R, Pc is the critical pressure in psia,
and SG is the 60°F/60°F specific gravity. The second method by Bergman (1976) is only applied to
components with low densities (< 850 kg/m3) and low boiling points (< 548.316 K), and was therefore
not of interest in this work. The last method by Cavett (1964) is used for all other hypothetical
components and is of the following form:

768.071  1.7134 ˜T b 0.10834 u 10 2 ˜T b 0.3889 u 10 6 ˜T b


2 3
Tc
§ 141.5 · 2 § 141.5 ·
 0.89213 u 10 2 ˜T b ˜¨  131.5 ¸  0.53095 u 10 6 ˜T b ˜¨  131.5 ¸ (E.1.10)
© SG ¹ © SG ¹
2
2 § 141.5 ·
 0.32712 u 10 7 ˜T b ˜¨  131.5 ¸
© SG ¹

2.829  0.9412 u 10 3 ˜T b 0.30475 u 10 5 ˜T b 0.15141 u 10 8 ˜T b


2 3
log Pc
§ 141.5 · 2 § 141 .5 ·
 0.20876 u 10  4 ˜T b ˜¨  131.5 ¸  0.11048 u 10 7 ˜T b ˜¨  131.5 ¸ (E.1.11)
© SG ¹ © SG ¹
2 2
2 § 141.5 · § 141.5 ·
 0.1395 u 10 9 ˜T b ˜¨  131.5 ¸  0.4827 u 10 7 ˜T b ˜¨  131.5 ¸
© SG ¹ © SG ¹

A-64
Appendix E: Hypothetical K2CO3* HYSYS® Component Properties

where Tc and Tb are in °F, Pc is in psia, and SG is the 60°F/60°F specific gravity.

Since the critical temperature and critical pressure of potassium carbonate were not available in
literature, the Lee-Kesler model was used to estimate these two critical properties for the hypothetical
K2CO3* component. The critical temperature was estimated to be 1645.10°C while the critical
pressure was estimated to be 4507.57 kPa.

E.1.5 Critical Volume and Acentricity


In HYSYS®, the correlations proposed by Pitzer and co-workers (1955) are the default estimation
method for determining the critical volume and acentricity of hypothetical components belonging to the
Miscellaneous class of components. These correlations take the following form:
Pc ˜ Vc § T P · § T P · T P
Zc Z ( 0 ) ¨¨ , ¸¸  Z ˜ Z (1) ¨¨ , ¸ at 1 and 1 (E.1.12)
R ˜ Tc ¸
© Tc Pc ¹ © Tc Pc ¹ Tc Pc

§ Ps · T
Z  log¨ ¸  1.000 at 0. 7 (E.1.13)
¨P ¸ Tc
© c ¹

where Zc is the critical compressibility, Pc is the critical pressure , Vc is the critical molar volume, Tc is
the critical temperature, P is the pressure, T is the temperature, Z is the acentricity, and Ps is the
vapour pressure. Z(0) and Z(1) are functions of T, P, Tc and Pc.

As literature values for the critical volume and acentricity of potassium carbonate were unavailable,
the Pitzer model was used to estimate these two properties for the hypothetical K2CO3* component.
The critical volume was estimated to be 0.997346 m3/kmol while the acentricity was estimated to be
0.107562.

A-65
Appendix E: Hypothetical K2CO3* HYSYS® Component Properties

E.2 Additional Point Properties


E.2.1 Heat of Formation
Two default methods are available for the estimation of the heat (or enthalpy) of formation at 25°C for
hypothetical HYSYS® components. If the component’s structure can be defined in terms of UNIFAC
groups, the method proposed by Joback (1984) is applied; otherwise, a simple ratio with respect to
octane (Hyprotech Ltd, 2001a) is used instead. The Joback method is a group contribution method
that takes the form:
f
'h 25 qC 68.29  ¦N
k
k ˜ 'hfk (E.2.1)

f
where 'h 25 qC is the heat of formation at 25°C in kJ/mol, Nk is the number of UNIFAC groups of type k

and 'hfk is the contribution for group k to the heat of formation in kJ/mol. The alternative HYSYS®
default method takes the form:
f
f 'h 25 qC,oc tan e ˜ MW
'h 25 qC (E.2.2)
MW oc tan e

where MW is the molecular weight.

For the hypothetical K2CO3* component, the heat of formation was not estimated by either of the
above two methods, but was instead taken to be that for potassium carbonate: -1144610 kJ/kmol
(Chase, 1998).

E.2.2 Dipole Moment


No default estimation method for the dipole moment of a hypothetical component is currently available
in HYSYS®. Instead, the dipole moments for such components are set equal to zero (Hyprotech Ltd,
2001a). Consequently, the dipole moment for the hypothetical K2CO3* component was set as 0.00
Debye.

E.2.3 Radius of Gyration


To estimate the radius of gyration for a hypothetical component, HYSYS® uses a proprietary method
(Hyprotech Ltd, 2001a). For the hypothetical K2CO3* component, the radius of gyration estimated by
this method was 2.64138 Å.

A-66
Appendix E: Hypothetical K2CO3* HYSYS® Component Properties

E.3 Temperature Dependent Properties


E.3.1 Ideal Gas Enthalpy
In HYSYS®, the ideal gas enthalpy hIG of a component is calculated from the following temperature
dependent relation:
hIG a  b ˜ T  c ˜ T2  d ˜ T3  e ˜ T4  f ˜ T5 (E.3.1)

where hIG is in kJ/kg, T is in K, and the reference point is an ideal gas at 0 K. If not provided, the
coefficients a to f are estimated from the ideal gas heat capacity correlation proposed by Cavett
(1964):
§ § T 13 · ·
¨ ¨ b ¸  0.4673722 ¸ ˜ MW
C pIG ¨ 0.036863384 ˜ ¨¨ SG ¸¸ ¸
¨ ¸
© © ¹ ¹
§ § T 13 · ·
¨
 ¨ 3.1865 u 10 5 ˜ ¨¨ b ¸  0.001045186 ¸ ˜ MW ˜ T (E.3.2)
¨ SG ¸¸ ¸
¨ ¸
© © ¹ ¹

  4.9572 u 10 7 ˜ MW ˜ T 2
where CpIG is the ideal gas heat capacity in Btu/lbmol·°R, Tb is the boiling point in °R, SG is the
60°F/60°F specific gravity, MW is the molecular weight, and T is the temperature in °R.

No thermochemical data were available for potassium carbonate in the ideal gas state. Consequently,
heat capacity and enthalpy data from the NIST-JANAF Thermochemical Tables (Chase, 1998) for
crystalline potassium carbonate were instead used to derive the coefficients for equation (E.3.1) for
the hypothetical K2CO3* component. After the data had been corrected to account for the different
reference temperatures (0 K for equation (E.3.1) and 25°C for the NIST-JANAF tables), the following
relationship between heat capacity Cp and enthalpy h was used to regress the coefficient values given
in Table E.3.1:
§ wh ·
Cp ¨ ¸ b  2 ˜ c ˜ T  3 ˜ d ˜ T2  4 ˜ e ˜ T3  5 ˜ f ˜ T4 (E.3.3)
© wT ¹ P

where Cp is in kJ/kg·K, h is in kJ/kg, and T is in K.

Table E.3.1: Temperature dependent property correlation coefficients.

Coefficient Ideal Gas Enthalpy Ideal Gas Gibbs Free Energy Vapour Pressure
a -19.8419 -1142180 22.0529
b 0.365850 246.572 -27345.9
c 9.80515×10-4 4.51346×10-2 0.00
d -5.86603×10-7 0.00 -0.515316
e 2.49532×10-10 0.00 2.08489
f -3.99650×10-14 – 0.00

A-67
Appendix E: Hypothetical K2CO3* HYSYS® Component Properties

E.3.2 Ideal Gas Gibbs Free Energy


The ideal gas Gibbs free energy GIG of a component in HYSYS® is calculated from the following
temperature dependent relation:
GIG a  b ˜ T  c ˜ T2  d ˜ T3  e ˜ T4 (E.3.4)

where GIG is in kJ/kmol, T is in K, and the reference point is an ideal gas at 25°C. If not provided, the
coefficients a to e are estimated from a HYSYS® proprietary method (Hyprotech Ltd, 2001a).

As mentioned above, no thermochemical data were available for potassium carbonate in the ideal gas
state. Consequently, Gibbs free energy data from the NIST-JANAF Thermochemical Tables (Chase,
1998) for crystalline potassium carbonate were instead used to regress the coefficients for equation
(E.3.4) for the hypothetical K2CO3* component. The resulting coefficient values are given in Table
E.3.1.

E.3.3 Vapour Pressure


In HYSYS®, the vapour pressure of a component is calculated from the Modified Antoine equation:
b
lnP s a  d ˜ ln T  e ˜ T f (E.3.5)
Tc

where Ps is the vapour pressure in kPa and T is the temperature in K. If the coefficients a to f are not
provided, HYSYS® estimates these based on the vapour pressures calculated from the correlation
proposed by Riedel (1954):
§ Ps · §T ·
ln¨ ¸ 2.933 ˜ 3.758  D c  3.0168 ˜ 3.758  D c ˜ ¨¨ c ¸¸
¨P ¸ © T ¹
© c ¹
6
(E.3.6)
§ T · § T ·
 3.5196 ˜ 3.758  D c  D c ˜ ln¨¨ ¸  0.0838 ˜ 3.758  D c ˜ ¨
¸ ¨T
¸
¸
© Tc ¹ © c ¹
0.315 ˜ \ b  ln Pc
Dc (E.2.7)
§T ·
0.0838 ˜ \ b  ln¨¨ b ¸
¸
© Tc ¹
6
§T · §T · § Tb ·
\b 35  36 ˜ ¨¨ c ¸  42 ˜ ln¨ b
¸ ¨T
¸¨
¸ ¨T
¸
¸ (E.3.8)
© Tb ¹ © c ¹ © c ¹

where Ps and the critical pressure Pc are in atm, and T, the critical temperature Tc and the boiling point
Tb are in K.

Since vapour pressure data for potassium carbonate was unavailable in literature, data for an
inorganic compound with a melting point similar to that for potassium carbonate was used instead to
derive the coefficients for equation (E.3.5). The compound selected was lead oxide PbO which has a

A-68
Appendix E: Hypothetical K2CO3* HYSYS® Component Properties

melting point of 890°C (c.f. 900.85°C for potassium carbonate) and the coefficient values in Table
E.3.1 were regressed from its vapour pressure data provided by Perry and Green (1997).

A-69
Appendix F: Property Models for HYSYS®

APPENDIX F

PROPERTY MODELS FOR


HYSYS®

This appendix presents the various thermodynamic and physical property models utilised in the
HYSYS® process models for the CO2 trains. Where necessary, model parameters are regressed from
literature data.

A-70
Appendix F: Property Models for HYSYS®

F.1 Thermodynamic Property Models


F.1.1 Enthalpy
For the HYSYS® simulations, the vapour and liquid phase enthalpies h were calculated from the
rigorous thermodynamic relation:
V
h  hIG 1 ª § wP · º
R˜T
Z  1 ³
˜ «T ˜ ¨ ¸  P» ˜ dV
R ˜ T ¬ © wT ¹ V
f ¼
(F.1.1)

which in terms of the enhanced PR equation of state is:

h  hIG 1 § da · §¨ V  2  1 ˜ b ·¸
R˜T
Z  1
21. 5
˜ ¨a  T ˜
˜b ˜R ˜ T ©
¸ ˜ ln
dt ¹ ¨© V  2  1 ˜ b ¸¹ (F.1.2)

hIG is the ideal gas enthalpy (the ideal gas enthalpy of formation at 25°C), R is the gas constant, T is
the absolute temperature, Z is the compressibility, V is the molar volume, P is the pressure, and a and
b are the PR parameters.

F.1.2 Heat Capacity


The heat capacity Cp was calculated from the rigorous relation:
2 2
§ dP · § dV · § dU · § dP · § dV ·
Cp Cv  T ˜ ¨ ¸ ˜¨ ¸ ¨ ¸  T˜¨ ¸ ˜¨ ¸ (F.1.3)
© dV ¹ T © dT ¹ P © dT ¹ V © dV ¹ T © dT ¹ P

where U is the total internal energy. If a solution could not be obtained for this equation, the ideal gas
method was applied instead:
Cp Cp
(F.1.4)
Cv Cp  R

A-71
Appendix F: Property Models for HYSYS®

F.2 Physical and Transport Property Models


F.2.1 Molecular Weight
For the HYSYS® simulations, the molecular weight MW of a phase was determined from the mole-
fraction-weighted sums of the individual component molecular weights:
NC
MW ¦x
j 1
j ˜ MW j (F.2.1)

where x is the mole fraction and NC is the number of components.

F.2.2 Density
F.2.2.1 Vapour Phase Density
The vapour phase mass density UG was determined from:
P ˜ MW G
UG (F.2.2)
Z ˜R ˜ T

where P is the pressure, MWG is the vapour phase molecular weight, Z is the vapour phase
compressibility factor calculated from the enhanced PR equation of state, R is the gas constant, and T
is the absolute pressure.

F.2.2.2 Liquid Phase Density


F.2.2.2.1 Sour PR Property Package
For the Sour PR property package, the liquid phase mass density UL was determined from the
corresponding states liquid density (COSTALD) equation by Hankinson and Thomson (1979):
1 1
UL (F.2.3)
NC xj NC x j ˜ VL, j
¦U
j 1 L, j
¦
j 1
MW j

VL, j
v *j ˜ VRo ˜ 1  Z j ˜ VRG (F.2.4)

1 2
§ T ·¸
3 § T ·¸
3 § T ·¸
VRo 1  1.52816 ˜ ¨1   1.43907 ˜ ¨1   0.81446 ˜ ¨1 
¨ Tc, j ¸¹ ¨ Tc, j ¸¹ ¨ Tc, j ¸¹
© © ©
(F.2.5)
4
§ T ·¸
3
 0.190454 ˜ ¨1 
¨ Tc, j ¸¹
©
2 2
§ T · § · § ·
 0.296123  0.386914 ˜ ¨ ¸  0.0427258 ˜ ¨ T ¸  0.0480645 ˜ ¨ T ¸
¨ Tc, j ¸ ¨ Tc, j ¸ ¨ Tc, j ¸
© ¹ © ¹ © ¹
VRG (F.2.6)
T
 1.00001
Tc, j

A-72
Appendix F: Property Models for HYSYS®

where x is the liquid phase mole fraction, VL is the liquid phase molar volume, MW is the molecular
weight, v* is the characteristic volume, Z is the acentricity, T is the absolute temperature, Tc is the
critical temperature, and NC is the number of components.

F.2.2.2.2 PR Property Package


In the PR property package, the HYSYS® tabular model was used instead of the default COSTALD
equation to facilitate the more accurate prediction of the liquid phase density. The temperature
dependence of the pure component liquid mass densities UL,j was described by polynomials of the
form:
U L, j A  B ˜ T  C ˜ T2  D ˜ T3  E ˜ T4 (F.2.7)

where UL,j is in kg/m3 and T is the absolute temperature in K, and the overall liquid phase density was
calculated from equation (F.2.3).

The default HYSYS® coefficient values for H2O were used for H2O and the hypothetical water
component H2O*. The effect of H2S and other gases (such as N2, CH4 and other hydrocarbons) on
the liquid phase density was disregarded due to their negligible liquid phase concentrations compared
to that of CO2, and these gases were assigned the same coefficient values as H2O. The coefficient
values for the hypothetical potassium carbonate component K2CO3* were regressed from the literature
data in Table B.2.3 for pure potassium carbonate solutions while the coefficient values for CO2 were
regressed from the data for pure potassium bicarbonate and potassium carbonate-bicarbonate
solutions. The K2CO3* and CO2 coefficients were determined via the simple unweighted least squares
method in a Microsoft® Excel spreadsheet, and the resulting values are listed in Table F.2.1.

Figure F.2.1 compares the liquid phase mass density values predicted by the tabular model against
the literature values. The average absolute deviation between the predicted and literature values is
0.8%, compared to 4.3% for the values predicted by the COSTALD equation.

Table F.2.1: Coefficient values for the HYSYS® liquid density tabular model.

Coefficient K2CO3* CO2 H2O and other components


A 3.4050×101 -7.5735×102 5.8305×101
B -2.1173×10-2 7.2698 -5.3842×10-3
C 0 -2.1830×10-2 -2.4881×10-5
D 0 2.1418×10-5 4.7481×10-8
E 0 0 -6.5125×10-11

A-73
Appendix F: Property Models for HYSYS®

1450

Predicted Solution Mass Density (kg/m3)


Potassium carbonate
Potassium bicarbonate
1350 Potassium carbonate-bicarbonate

1250

1150

1050

950
950 1050 1150 1250 1350 1450
3
Experimental Solution Mass Density (kg/m )

Figure F.2.1: Comparison between the predicted and experimental solution mass densities. The dashed
lines (---) represent the ± 1% lines.

F.2.3 Viscosity
F.2.3.1 Vapour Phase Viscosity
The vapour phase dynamic viscosity was determined from a proprietary modification of the model
proposed by Ely and Hanley (1981). In the original model, the viscosity P of a mixture at density U,
temperature T and composition x is equated to the viscosity Px of a hypothetical pure fluid:
P U, T P x U, T (F.2.8)

This enables the use of the following corresponding states argument to determine P from the viscosity
P0 of a reference fluid:

MWx fx,0
P x U, T P0 U0 , T0 ˜ ˜ (F.2.9)
MW0 h 2 3
x,0

T
T0 (F.2.10)
f x,0

U0 U ˜ h x,0 (F.2.11)

MW is the molecular weight. The subscript x refers to the fluid of interest while the subscript 0 refers
to the reference fluid. fx,0 and hx,0 are functions of the critical parameters and the acentricities of the
reference fluid and the fluid of interest.

A-74
Appendix F: Property Models for HYSYS®

F.2.3.2 Liquid Phase Viscosity


F.2.3.2.1 Sour PR Property Package
In the Sour PR property package, the liquid phase viscosity was determined in one of three ways: via
the proprietary modified Ely-Hanley model for light hydrocarbons; via the Twu model (Twu, 1985) for
heavy hydrocarbons; or via the proprietary modified Letsou-Stiel model (Letsou and Stiel, 1973) for all
other chemicals.

In the original model proposed by Letsou and Stiel (1973), the liquid phase dynamic viscosity PL is
determined from the following equations:
3
§ NC 1 ·
PL ¨
¨ ¦ x j ˜ P L, j 3 ¸
¸
(F.2.12)
©j1 ¹
1 2
Tc, j 6 T § T ·
P L, j ˜ 0.015174  0.02135 ˜  0.0075 ˜ ¨ ¸
2 Tc, j ¨ Tc, j ¸
Pc, j 3 ˜ MW j © ¹
(F.2.13)
§ § T ·
2 ·
¨ T ¸
 Z j ˜ ¨ 0.042552  0.07674 ˜  0.0340 ˜ ¨ ¸
¸
¨ Tc, j ¨ Tc, j ¸ ¸
© © ¹ ¹

where PL is in cP, x is the mole fraction, NC is the number of components, Tc is the critical temperature
in K, Pc is the critical pressure in atm, MW is the molecular weight, and Z is the acentricity.

The Twu model is more complex, first requiring the calculation of the component kinematic viscosities
at 100°F and 210°F from their boiling points Tb and specific gravities SG:
2
§ 450 ·¸ § 450 ·¸ §¨ 1  2 ˜ f T, j ·
ln¨ Q L, j,T  ln¨ Q Lo, j,T  ˜ ¸ for T = 100°F, 210°F (F.2.14)
¨ Tb, j ¸¹ ¨ Tb, j ¸¹ ¨© 1  2 ˜ f T, j ¸
© © ¹
2
56.7394 21.1141 ˜ 'SG j
f100qF, j 1.33932 ˜ 1.99873  ˜ 'SG j  (F.2.15)
Tb, j Tb, j

2
56.7394 21.1141 ˜ 'SG j
f 210qF, j 1.99873  ˜ 'SG j  (F.2.16)
Tb, j Tb, j

2
§1 2 ˜ hj ·
'SG j SG j  SG oj
˜ 1.49546  SG j ˜ ¨
¨1 2 ˜ hj
¸
¸
(F.2.17)
© ¹

§ ·
hj
¨
¨¨  21 .6364 
844.687 ¸ o

¸ ˜ SG j  SG j ˜ 1.49546  SG j
Tb, j ¸
© ¹
(F.2.18)
§ ·
¨
 ¨ 458.199 
¨
7543.00 ¸
¸ >
˜ SG j  SG oj ˜ 1.49546  SG j
2
@
Tb, j ¸
© ¹

A-75
Appendix F: Property Models for HYSYS®

where QL is the kinematic viscosity in cSt , T is the temperature in °F, and Tb is in °R. The reference
variables, denoted by the superscript o, are obtained from:
2
§ Tbo, j · § o ·
ln Q Lo,210qF, j  1 .5 4.73227  27.0975 ˜ ¨1  o
¨
¸  49.4491 ˜ ¨1  Tb, j
¸ ¨
¸
¸
© Tc, j ¹ © Tco, j ¹
(F.2.19)
4
§ Tbo, j ·
¨
 50.4706 ˜ 1  o ¸
¨ Tc, j ¸
© ¹


ln Q Lo,100qF, j 0.801621  1.37179 ˜ ln Q Lo,210qF, j (F.2.20)

3
§ Tbo, j · § o ·
SG oj ¨
0.843593  0.128624 ˜ 1  o ¸  3.36129 ˜ ¨1  Tb, j ¸
¨ Tc, j ¸ ¨ Tco, j ¸
© ¹ © ¹
(F.2.21)
12
§ Tbo, j ·
 13749 .5 ˜ ¨1  o ¸
¨ Tc, j ¸
© ¹

Tbo, j
Tco, j (F.2.22)
§ 0.533272  0.191017 u 10 3 ˜ T o  0.779681 u 10 7 ˜ T o 2 ·
¨ b, j b, j ¸
¨ 3 13 ¸
¨  0.284376 u 10 10
˜ T o
 0 . 959468 u 10  28
˜ T o ¸
© b , j b , j ¹

The two calculated kinematic viscosities are then substituted into the following equations to determine
the component dynamic viscosities PL,j at temperature T, which are then used to calculate the liquid
phase dynamic viscosity from equation (F.2.12):

1.47 1.84˜Q L,T, j 0.51˜Q L,T, j 2


Z T, j Q L,T, j  0.7  e for T = 100°F, 210°F (F.2.23)


ln ln Z 100qF, j  ln ln Z 210qF, j

ln ln Z j
ln ln Z 100qF, j  ln 100qF  ln 210qF
˜ ln T  ln 100qF (F.2.24)


0.7487 3.295˜ Z j 0.7  6.119˜ Z j 0.7 2 0.3193˜ Z j 0.7 3
Q L, j Z j  0. 7  e (F.2.25)

P L, j U L, j ˜ Q L, j (F.2.26)

It should be noted that in equation (F.2.26), PL,j is in kg/m·s, QL,j is in m2/s and the density UL,j is in
kg/m3.

F.2.3.2.2 PR Property Package


In the PR property package, the HYSYS® tabular model was used, instead of the above-mentioned
default HYSYS® models, to facilitate the more accurate prediction of the liquid phase viscosity. The
temperature dependence of the pure component liquid dynamic viscosities PL,j was described by
polynomials of the form:
B
ln P L, j A  C ˜ ln T  D ˜ T E (F.2.27)
T

A-76
Appendix F: Property Models for HYSYS®

where PL,j is in cP and T is the absolute temperature in K, and the overall liquid phase viscosity PL was
calculated from equation (F.2.12).

As for the liquid density tabular model, the default HYSYS® coefficient values for H2O were used for
H2O, the hypothetical water component H2O* and H2S and other gases (such as N2, CH4 and other
hydrocarbons). The coefficient values for the hypothetical potassium carbonate component K2CO3*
were regressed from the literature data in Table B.2.8 for pure potassium carbonate solutions while
the coefficient values for CO2 were regressed from the data for pure potassium bicarbonate and
potassium carbonate-bicarbonate solutions. The K2CO3* and CO2 coefficients were determined via
the simple unweighted least squares method in a Microsoft® Excel spreadsheet, and the resulting
values are listed in Table F.2.2.

Figure F.2.2 compares the liquid phase dynamic viscosity values predicted by the tabular model
against the literature values. The average absolute deviation between the predicted and literature
values is 4.6%, compared to 7.5% for the values predicted by the default HYSYS® models.

Table F.2.2: Coefficient values for the HYSYS® liquid viscosity tabular model.

Coefficient K2CO3* CO2 H2O and other components


A -2.3878×102 -2.7937E+03 -1.1213×10-1
B 9.7635E×103 8.0287E+04 -2.4004×10-4
C 3.8379E×101 4.8288E+02 -3.1123
D -4.2743×10-2 -7.3978E-01 0
E 1.0000 1.0000 0

3.5

Potassium carbonate
3.0 Potassium bicarbonate
Predicted Solution Viscosity (cP)

Potassium carbonate-bicarbonate

2.5

2.0

1.5

1.0

0.5

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

Experimental Solution Viscosity (cP)

Figure F.2.2: Comparison between the predicted and experimental solution viscosities. The dashed lines
(---) represent the ± 5% lines.

A-77
Appendix F: Property Models for HYSYS®

F.2.4 Surface Tension


F.2.4.1 Sour PR Property Package
In the Sour PR property package, the liquid phase surface tension VL was determined in one of two
ways: via a proprietary modification of the correlation proposed by Brock and Bird (1955) for
hydrocarbon systems or via a proprietary polynomial for aqueous systems. The original form of the
Brock-Bird model is as follows:
1
VL NC
(F.2.28)
xj
¦V
j 1 L, j

§ § Tb, j · ·
¨ ¨ ˜ ln Pc, j ¸ ¸ 11
2 1 ¨ ¨ Tc, j ¸ ¸ § T · 9
V L, j Pc, j 3 ˜ Tc, j 3 ˜ ¨ 0.1207 ˜ ¨1  ¸  0.281¸ ˜ ¨¨1  ¸ (F.2.29)
¨ ¨ Tb, j ¸ ¸ © Tc, j ¸¹
¨¨ ¨ 1 ¸ ¸¸
© Tc, j ¹
© ¹

where VL is in dyne/cm, x is the mole fraction, NC is the number of components, Pc is the critical
pressure in atm, Tc is the critical temperature in K, and Tb is the boiling point in K.

F.2.4.2 PR Property Package


In the PR property package, the HYSYS® tabular model was used, instead of the above-mentioned
default HYSYS® models, to facilitate the more accurate prediction of the liquid phase surface tension.
The temperature dependence of the pure component liquid surface tensions VL,j was described by
polynomials of the form:
V L, j A  B ˜ T  C ˜ T2  D ˜ T3  E ˜ T4 (F.2.30)

where VL,j is in dyne/cm and T is the absolute temperature in K, and the overall liquid phase surface
tension VL was calculated from equation (F.2.28).

Like the liquid density and liquid viscosity tabular models, the default HYSYS® coefficient values for
H2O were used for H2O, the hypothetical water component H2O* and H2S and other gases (such as
N2, CH4 and other hydrocarbons). The coefficient values for the hypothetical potassium carbonate
component K2CO3* and for CO2 were regressed from a set of values generated using the empirical
surface tension correlation (B.2.37). The K2CO3* and CO2 coefficients were determined via the simple
unweighted least squares method in a Microsoft® Excel spreadsheet, and the resulting values are
listed in Table F.2.3.

Figure F.2.3 compares the liquid phase surface tension values predicted by the tabular model against
those generated by the empirical correlation. The average absolute deviation between the two sets of
values is 3.8%, compared to 94.4% for the values predicted by the modified Brock-Bird model.

A-78
Appendix F: Property Models for HYSYS®

Table F.2.3: Coefficient values for the HYSYS® liquid surface tension tabular model.

Coefficient K2CO3* CO2 H2O and other components


A 2.5209×102 -3.0588×102 1.8277×104
B -1.4077 1.2235 -9.3402×101
C 4.4242×10-3 -1.2345×10-3 1.1945×10-1
D -6.9349×10-6 0 0
E 3.9069×10-9 0 0

100
Tabular Model Solution Surface Tension (dyne/cm)

Potassium carbonate
Potassium bicarbonate
Potassium carbonate-bicarbonate
90

80

70

60
60 70 80 90 100
Empirical Correlation Solution Surface Tension
(dyne/cm)

Figure F.2.3: Comparison between the solution surface tensions predicted by the tabular model and the
empirical correlation. The dashed lines (---) represent the ± 5% lines.

F.2.5 Thermal Conductivity


F.2.5.1 Vapour Phase Thermal Conductivity
The vapour phase thermal conductivity was determined from a proprietary modification of the model
proposed by Ely and Hanley (1983). In the original model, the thermal conductivity O of the fluid of
interest (a pure fluid or a mixture) at density U and temperature T is divided into two contributory terms:
O U, T O' U, T  O" T (F.2.31)

The first term O’ represents the transfer of energy from purely collisional or translational effects, while
the second term O” is due to the transfer of energy via the internal degrees of freedom. The
translational contribution O’ for a mixture of composition x is equated to the contribution O’x for a
hypothetical pure fluid:
O' U, T O' x U, T (F.2.32)

A-79
Appendix F: Property Models for HYSYS®

Consequently, from the corresponding states argument, O’x is obtained from the translational
contribution O’0 of a reference fluid:
3
ª§ · Z º 2
MW 0 f x,0 T § wf x,0 · ¸ ˜ c,0 »
O' x U, T O' 0 U 0 , T0 ˜ ˜ ˜ «¨¨1  ˜¨ ¸
¸ (F.2.33)
MW x h 2 3 « f x,0 ¨© wT ¹ Vx
¸ Z »
x,0 ¬© ¹ c, x ¼

MW is the molecular weight and Zc is the critical compressibility. The subscript x refers to the fluid of
interest while the subscript 0 refers to the reference fluid. T0 and U0 are given by equations (F.2.10)
and (F.2.11). fx,0 and hx,0 are functions of the critical parameters and the acentricities of the reference
fluid and the fluid of interest.

The internal contribution O” is given by:

§ 5 ˜R · P
O" T 1.32 ˜ ¨ C pIG  ¸˜ (F.2.34)
© 2 ¹ MW

where CpIG is the ideal gas heat capacity, R is the gas constant, and P is the viscosity from equation
(F.2.8).

Two other methods were also used to calculate the vapour phase thermal conductivities in the
HYSYS® simulations: the simple set of correlations proposed by Misic and Thodos (1961, 1963) and
the more complex model by Chung and co-workers (1988). The Misic-Thodos correlations are based
on dimensional analysis and assume that the vapour phase thermal conductivity OG is described by a
function with the general form:

OG B D E
f MW A ˜ Tc ˜ T C ˜ Pc ˜ Vc ˜ Cp ˜ RG
F

§§ T 5 ·
2 C F
· § Cp · (F.2.35)
˜ f ¨¨ ¨¨ ¸ ˜ Z c 6  G F ˜ R 6 ¸
Pc 3 5
¸ ˜¨
¸ ¨ R ¸ ¸¸
Tc 6 ˜ MW ¨© © Tc
1
¹ © ¹ ¹

where OG is in cal/cm·s·K, MW is the molecular weight in g/mol, Tc is the critical temperature in K, T is


the temperature in K, Pc is the critical pressure in atm, Vc is the critical volume in cm3/mol, Zc is the
critical compressibility, Cp is the heat capacity in cal/mol·K, and R is the gas constant in
atm·cm3/mol·K. Several functions, based on the equation (F.2.35), were obtained by Misic and
Thodos (1961, 1963) for different types of compounds over various reduced temperature ranges.

While the Misic-Thodos correlations are suitable only for low pressures, the Chung model is applicable
to both high and low pressures:

31.2 u 10 3 ˜ P LP
G ˜< § V ·
OG ˜ ¨¨ F1  F2 ˜ c ¸¸
MW © 6˜V¹
2
(F.2.36)
10 3 ˜ Tc 2 § V · T
 3.586 u 10 3 ˜ ˜ Vc 3 ˜ F1 ˜ F3 ˜ ¨¨ c ¸¸ ˜
MW ©6˜V¹ Tc

A-80
Appendix F: Property Models for HYSYS®

< f T, Tc , C v , Z (F.2.37)

F1 f Tc , V, Vc , Z, G, N (F.2.38)

F2 f Tc , Vc , Z, G, N (F.2.39)

F3 f Tc , Vc , Z, G, N (F.2.40)

where OG is in W/m·K, P LP
G is the low pressure vapour phase viscosity in kg/m·s, MW is the molecular

weight in kg/kmol, Vc is the critical volume in cm3/mol, V is the molar volume in cm3/mol, T is the
temperature in K, Tc is the critical temperature in K, Cv is the heat capacity at constant volume in
J/mol·K, Z is the acentricity, G is the dipole moment in Debye, and N is the Chung association factor.

F.2.5.2 Liquid Phase Thermal Conductivity


F.2.5.2.1 Sour PR Property Package
In the Sour PR property package, the liquid phase thermal conductivity was determined in one of five
ways: via the proprietary modified Ely-Hanley model; via a proprietary polynomial for selected
components including H2O, CH4, C2H6 and C3H8; via a proprietary modification of the Missenard-
Riedel equation (Reid et al., 1977) for hydrocarbons with molecular weights greater than 140 kg/kmol
at reduced temperatures greater than 0.8; via the Latini method (Baroncini et al., 1980) for alcohols,
esters and all other hydrocarbons; or via the Sato-Riedel equation (Reid et al., 1977) for all other
compounds.

The liquid phase thermal conductivity OL is determined from the original Missenard-Riedel equation as
follows:
3
§ NC 1 ·
OL ¨
¨¦ x j ˜ O L, j 3 ¸
¸
(F.2.41)
©j1 ¹
2
§ T ·¸
3
3  20 ˜ ¨1 
84 u 10  6 ˜ Tb, j ˜ C L,0qC, j ˜ C p,0qC, j ¨ Tc, j ¸¹
©
O L, j 1
˜ 2
(F.2.42)
MW j ˜ N j 4 § 273 ·¸
3
3  20 ˜ ¨1 
¨ Tc, j ¸¹
©

where OL is in cal/cm·s·K, x is the mole fraction, Tb is the normal boiling point in K, CL,0°C is the liquid
phase density at 0°C in mol/cm3, CpL,0°C is the liquid heat capacity at 0°C in cal/mol·K, MW is the
molecular weight in g/mol, N is the number of atoms in the molecule, and Tc is the critical temperature
in K.

The Sato-Reidel equation takes a similar form to the Missenard-Riedel equation. The liquid phase
thermal conductivity is determined from equation (F.2.41) and the component thermal conductivities
are obtained from:

A-81
Appendix F: Property Models for HYSYS®

2
§ T ·¸
3
3  20 ˜ ¨1 
¨ Tc, j ¸¹
2.64 u 10 3 ©
O L, j ˜ 2
(F.2.43)
MW j § Tb, j · 3
3  20 ˜ ¨1  ¸
¨ Tc, j ¸
© ¹

The Latini method also uses equation (F.2.41) to calculate the liquid phase thermal conductivity from
component thermal conductivities given by:
0.38
§ ·
¨1  T ¸
A * ˜ Tb, j
D ¨ Tc, j ¸¹
©
O L, j E J
˜ 1
(F.2.44)
MW j ˜ Tc, j § T · 6
¨ ¸
¨ Tc, j ¸
© ¹

where OL is in W/m·K, and A*, D, E and J are component specific parameters.

F.2.5.2.2 PR Property Package


In the PR property package, the HYSYS® tabular model was used, instead of the above-mentioned
default HYSYS® models, to facilitate the more accurate prediction of the liquid phase thermal
conductivity. The temperature dependence of the pure component liquid thermal conductivities OL,j
was described by polynomials of the form:
O L, j A  B ˜ T  C ˜ T2  D ˜ T3  E ˜ T4 (F.2.45)

where OL,j is in W/m·K and T is the absolute temperature in K, and the overall liquid phase surface
tension OL was calculated from:
1
OL 2
(F.2.46)
§ NC y j ·
¨
¨¦ 2
¸
¸
© j 1 O L, j ¹

where y is the mass fraction and NC is the number of components.

Like the other liquid property tabular models, the default HYSYS® coefficient values for H2O were
used for H2O, the hypothetical water component H2O* and H2S and other gases (such as N2, CH4 and
other hydrocarbons). The coefficient values for the hypothetical potassium carbonate component
K2CO3* and for CO2 were regressed from a set of values generated using the empirical thermal
conductivity correlation (B.2.55). The K2CO3* and CO2 coefficients were determined via the simple
unweighted least squares method in a Microsoft® Excel spreadsheet, and the resulting values are
listed in Table F.2.4.

A-82
Appendix F: Property Models for HYSYS®

Figure F.2.4 compares the liquid phase thermal conductivity values predicted by the tabular model
against those generated by the empirical correlation. The average absolute deviation between the two
sets of values is 1.2%, compared to 59.1% for the values predicted by the default HYSYS® methods.

Table F.2.4: Coefficient values for the HYSYS® liquid thermal conductivity tabular model.

Coefficient K2CO3* CO2 H2O and other components


-1 -1
A -4.3200×10 5.1309×10 1.7269
B 5.7255×10-3 2.4746×10-4 -3.0912×10-3
C -8.0780×10-6 0 0
D 1.8610×10-9 0 0
E 0 0 0

0.70
Tabular Model Solution Thermal Conductivity

Potassium carbonate
Potassium bicarbonate
Potassium carbonate-bicarbonate

0.65
(W/m·K)

0.60

0.55
0.55 0.60 0.65 0.70
Empirical Correlation Solution Thermal Conductivity
(W/m·K)

Figure F.2.4: Comparison between the solution thermal conductivities predicted by the tabular model and
the empirical correlation. The dashed lines (---) represent the ± 1% lines.

A-83
Appendix G: Enhanced PR Binary Interaction Parameters

APPENDIX G

ENHANCED PR BINARY
INTERACTION PARAMETERS

This appendix outlines the procedure followed for the regression of the model parameters for the
enhanced PR equation of state. The results for the data regression are analysed and presented
within.

A-84
Appendix G: Enhanced PR Binary Interaction Parameters

G.1 Data Regression Procedure


The enhanced PR equation of state binary interaction parameter values that were not taken from the
HYSYS® property library were determined from a series of data regressions. As discussed in
Sections 6.2.2.1 and 6.2.2.2, binary interaction parameters for the H2O*-K2CO3*, CO2-H2O* and CO2-
K2CO3* component pairs were first obtained from the CO2 solubility data presented by Tosh and co-
workers (1959). The interaction parameters for the H2S-H2O* and H2S-K2CO3* component pairs were
then determined using the H2S solubility data presented by Tosh and co-workers (1960). Both sets of
data regressions were performed using the following general procedure:

1. The initial values for the binary interaction parameters were taken as 0 (the HYSYS® default
value), except for the CO2-H2O* and H2S-H2O* component pairs. These were set equal to the
default HYSYS® values for the CO2-H2O and H2S-H2O component pairs.

2. Parameter bounds were set according to findings from the preliminary simulations discussed
in Section 6.1.1. It was determined that negative values were required for the H2O*-K2CO3*,
CO2-H2O*, CO2-K2CO3* and H2S-K2CO3* component pair parameters in order to ensure
sufficient absorption of CO2 and H2S in the absorber, while the H2S-H2O* component pair
parameter had to be positive to ensure sufficient H2S desorption in the regenerator.
Consequently, the upper bounds for the H2O*-K2CO3*, CO2-H2O*, CO2-K2CO3* and H2S-
K2CO3* component pair parameters were set as 0, while the lower bound for the H2S-H2O*
component pair parameter was set as 0.

3. The full set of interaction parameters was regressed from the data, and the resulting set of
parameter values was labelled “Full”.

4. One of the interaction parameters was excluded from regression (i.e. it was fixed at its initial
value) and the remaining parameters were regressed from the data. This step was repeated
with the remaining parameters.

5. Step 3 was repeated with an additional interaction parameter being excluded from regression
until all the parameters were fixed at their initial values.

6. Each simplified set of parameter values obtained from steps 3 and 4 was compared against
the “Full” parameter value set via an F-Test to determine the reliability of the simplified set.
The corresponding one-tail p-value was also calculated to determine the probability of
observing the F-Test value if the null hypothesis was true.
Null hypothesis: The simplified set fits the data better than the “Full” set.
F  1 or p > 0.05

where

A-85
Appendix G: Enhanced PR Binary Interaction Parameters

SSQ Simplified  SSQ Full


df Simplified  dfFull
F (G.1.1)
SSQ Full
df Full

p
f F, df Simplified  df Full , df Full (G.1.2)

df NData  NParameters (G.1.3)

SSQFull and SSQSimplified are the sum of squares for the full and simplified parameter value
sets, respectively, dfFull and dfSimplified are the corresponding degrees of freedom, NData is the
number of data points used in the regression, and NParameters is the number of parameters
regressed.

7. The simplified parameter value sets which passed the above F-Test were sorted according to
their sum of squares. A logic test was then performed on each parameter set to tabulate the
number of parameters with a standard error greater than the associated regressed value. The
parameter value set with the lowest logic test result and the lowest weighted sum of squares
was labelled the optimal set.

A-86
Appendix G: Enhanced PR Binary Interaction Parameters

G.2 Data Regression Results


The statistical results for the data regression runs for the CO2-K2CO3-H2O system are given in Table
G.2.1. The sum of squares (SSQ), residual root mean square error (RRMSQE), degrees of freedom
(df), F-value and p-value are listed for each parameter set. The data set used in the regression runs
consisted of 120 data points, and the number of parameters regressed ranged from 0 to 3, depending
on the parameter set. From the F-Test, none of the simplified parameter sets were identified as being
more suitable than the “Full” set. Consequently, the “Full” set was selected as the optimal parameter
value set and was used in the next series of data regression runs for the CO2-H2S-K2CO3-H2O system.
The regressed parameter values and standard errors associated with this optimal set are given in
Table 6.2.2.

Table G.2.2 presents the statistical results for the data regression runs for the CO2-H2S-K2CO3-H2O
system. The data set used in this series of regression runs consisted of 127 data points, and the
number of parameters regressed ranged from 0 to 2, depending on the parameter set. As for the
previous set of regressions, none of the simplified parameter sets were determined to be more
suitable than the “Full” set via the F-Test. Consequently, the “Full” set was selected as the optimal
parameter value set for the CO2-H2S-K2CO3-H2O system, and its regressed parameter values and
standard errors are given in Table 6.2.3.

Table G.2.1: Statistical results for the CO2-K2CO3-H2O system data regression runs.

Parameter Set SSQ RRMSQE df F p Null Hypothesis


Full 150.9 1.290 117 - - -
Fixed k H O* K CO * 185.9 1.255 118 27.18 8.1×10-7 False
2 2 3

Fixed k CO K CO * 211.5 1.339 118 47.04 3.5×10 -10


False
2 2 3

Fixed k CO H O* 375.2 1.783 118 173.93 6.8×10 -25


False
2 2

Fixed k H O* K CO *
2 2 3
220.9 1.363 119 27.16 2.0×10-10 False
and k CO K CO *
2 2 3

Fixed k H O* K CO *
2 2 3
385.4 1.800 119 90.93 1.5×10-24 False
and k CO H O*
2 2

Fixed k CO K CO *
2 2 3
1160.3 3.123 119 391.37 1.5×10-52 False
and k CO H O*
2 2

All Fixed 2704.5 4.747 120 660.07 4.0×10-73 False

Table G.2.2: Statistical results for the CO2-H2S-K2CO3-H2O system data regression runs.

Parameter Set SSQ RRMSQE df F p Null Hypothesis


Full 1611.9 3.591 125 -
Fixed kH S K CO * 2361.1 4.329 126 58.10 5.4×10-12 False
2 2 3

Fixed kH S K CO * 2971.0 4.856 126 105.39 2.6×10-18 False


2 2 3
-32
All Fixed 5113.9 6.346 127 135.78 4.6×10 False

A-87
Appendix H: HYSYS® Simulation Results

APPENDIX H

HYSYS® SIMULATION RESULTS


This appendix presents the results of the HYSYS® simulations for CO2 trains #2 to #6. Included are
the column profiles generated by the preliminary column simulations and the column profiles predicted
by the full CO2 trains process models. Also presented are the process flow diagrams for the HYSYS®
process models for CO2 trains #2 to #6.

A-88
Appendix H: HYSYS® Simulation Results

H.1 Preliminary Column Model Simulations


H.1.1 Equilibrium Stage Model Approach
The composition and temperature profiles predicted by the equilibrium stage models for the absorber
and regenerator columns in CO2 trains #2 to #6 are presented below.

CO2
CO2 H2S
H2S VapourPhase
Vapour Phase LiquidPhase
Liquid Phase
CO2 Data
CO2 Data H2S Data
H2S Data VapourData
Vapour Data LiquidData
Liquid Data

Regenerator H2S Loading


0.E+00 1.E-05 2.E-05 3.E-05
Absorber H2S (ppm) Cond
1.2
0 10 20 30 WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4
0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 30 50 70 90 110 130 0.0 0.2 0.4 0.6 0.8 60 70 80 90 100 110 120

Absorber CO2 (mol%) Absorber Temperature (°C) Regenerator CO2 Loading Regenerator Temperature (°C)

(a)
Regenerator H2S Loading
0.E+00 1.E-05 2.E-05 3.E-05
Absorber H2S (ppm) Cond
1.2
0 10 20 30 WS
Fraction of Total Packed Height

1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4
0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 90 100 110 120 130 0.0 0.2 0.4 0.6 0.8 60 70 80 90 100 110 120

Absorber CO2 (mol%) Absorber Temperature (°C) Regenerator CO2 Loading Regenerator Temperature (°C)

(b)
Regenerator H2S Loading
0.E+00 1.E-05 2.E-05 3.E-05
Absorber H2S (ppm) Cond
1.2
0 10 20 30 WS
Fraction of Total Packed Height

1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4

0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 90 100 110 120 130 0.0 0.2 0.4 0.6 0.8 60 70 80 90 100 110 120

Absorber CO2 (mol%) Absorber Temperature (°C) Regenerator CO2 Loading Regenerator Temperature (°C)

(c)
Figure H.1.1: Equilibrium stage simulation results for the absorber and regenerator columns.
(a) CO2 train #2. (b) CO2 train #3. (c) CO2 train #4. (Cond = condenser, WS = wash section, Reb =
reboiler)

A-89
Appendix H: HYSYS® Simulation Results

CO2
CO2 H2S
H2S VapourPhase
Vapour Phase LiquidPhase
Liquid Phase
CO2 Data
CO2 Data H2S Data
H2S Data VapourData
Vapour Data LiquidData
Liquid Data

Regenerator H2S Loading


0.E+00 1.E-05 2.E-05 3.E-05
Absorber H2S (ppm) Cond
1.2
0 10 20 30 WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4

0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 90 100 110 120 130 0.0 0.2 0.4 0.6 0.8 60 70 80 90 100 110 120

Absorber CO2 (mol%) Absorber Temperature (°C) Regenerator CO2 Loading Regenerator Temperature (°C)

(a)
Regenerator H2S Loading
0.E+00 1.E-05 2.E-05 3.E-05
Absorber H2S (ppm) Cond
1.2
0 10 20 30 WS

Fraction of Total Packed Height


1.0 1.0
Fraction of Total Packed Height

0.8
0.8

0.6
0.6
0.4
0.4
0.2

0.2
0.0
Reb
0.0 -0.2
0 5 10 15 20 90 100 110 120 130 0.0 0.2 0.4 0.6 0.8 60 70 80 90 100 110 120

Absorber CO2 (mol%) Absorber Temperature (°C) Regenerator CO2 Loading Regenerator Temperature (°C)

(b)
Figure H.1.2: Equilibrium stage simulation results for the absorber and regenerator columns.
(a) CO2 train #5. (b) CO2 train #6. (Cond = condenser, WS = wash section, Reb = reboiler)

A-90
Appendix H: HYSYS® Simulation Results

H.1.2 Column Stage Efficiency Correlations


The composition and temperature profiles predicted using the correlated overall stage efficiencies for
the steady-state and dynamic absorber columns in CO2 trains #2 to #6 are presented below.

Default Efficiencies Correlated Efficiencies Data

1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 30 50 70 90 110 130 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(a)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(b)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(c)
Figure H.1.3: Effect of the correlated overall stage efficiencies on the steady-state absorber columns.
(a) CO2 train #2. (b) CO2 train #3. (c) CO2 train #4.

A-91
Appendix H: HYSYS® Simulation Results

Default Efficiencies Correlated Efficiencies Data

1.0

Fraction of Total Packed Height 0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(a)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(b)
Figure H.1.4: Effect of the correlated overall stage efficiencies on the steady-state absorber columns.
(a) CO2 train #5. (b) CO2 train #6.

A-92
Appendix H: HYSYS® Simulation Results

Default Efficiencies Steady-State Efficiencies Dynamic Efficiencies Data

1.0

Fraction of Total Packed Height 0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 30 50 70 90 110 130 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(a)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(b)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(c)
Figure H.1.5: Effect of the correlated overall stage efficiencies on the steady-state behaviour of the
dynamic absorber columns. (a) CO2 train #2. (b) CO2 train #3. (c) CO2 train #4.

A-93
Appendix H: HYSYS® Simulation Results

Default Efficiencies Steady-State Efficiencies Dynamic Efficiencies Data

1.0

Fraction of Total Packed Height


0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(a)
1.0
Fraction of Total Packed Height

0.8

0.6

0.4

0.2

0.0
0 5 10 15 20 0 10 20 30 90 100 110 120 90 100 110 120 130
CO2 (mol%) H2S (ppm) Vapour Phase Temperature (°C) Liquid Phase Temperature (°C)

(b)
Figure H.1.6: Effect of the correlated overall stage efficiencies on the steady-state behaviour of the
dynamic absorber columns. (a) CO2 train #5. (b) CO2 train #6.

A-94
Appendix H: HYSYS® Simulation Results

H.2 Steady-State CO2 Train Models


The process flow diagrams for the steady-state HYSYS® models of CO2 trains #2 to #6 are provided
on the following pages.

A-95
Appendix H: HYSYS® Simulation Results
A-96

Figure H.2.1: Process flow diagram for the steady-state model of CO2 train #2.
Appendix H: HYSYS® Simulation Results
A-97

Figure H.2.2: Process flow diagram for the steady-state model of CO2 train #3.
Appendix H: HYSYS® Simulation Results
A-98

Figure H.2.3: Process flow diagram for the steady-state model of CO2 train #4.
Appendix H: HYSYS® Simulation Results
A-99

Figure H.2.4: Process flow diagram for the steady-state model of CO2 train #5.
Appendix H: HYSYS® Simulation Results
A-100

Figure H.2.5: Process flow diagram for the steady-state model of CO2 train #6.
Appendix H: HYSYS® Simulation Results

H.3 CO2 Train Model Validation


The vapour and liquid phase composition profiles and the temperature profiles for the absorber and
regenerator columns in the steady-state HYSYS® models for CO2 trains #2 to #6 are presented below.

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05 Cond
1.2
1.0 WS

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05 Cond
1.2
1.0 WS
Fraction of Total Packed Height

1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
Cond
1.2
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0
Fraction of Total Packed Height

1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(c)
Figure H.3.1: CO2 and H2S vapour and liquid phase column profiles for the first set of plant data.
(a) CO2 train #2. (b) CO2 train #3. (c) CO2 train #4. (Cond = condenser, WS = wash section, Reb =
reboiler)

A-101
Appendix H: HYSYS® Simulation Results

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
Cond
1.2
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 25 50 75 100 0.E+00 1.E-05 2.E-05 3.E-05
Cond
1.2
0 10 20 30 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Figure H.3.2: CO2 and H2S vapour and liquid phase column profiles for the first set of plant data.
(a) CO2 train #5. (b) CO2 train #6. (Cond = condenser, WS = wash section, Reb = reboiler)

A-102
Appendix H: HYSYS® Simulation Results

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05
0 5 10 15 20 0.E+00 1.E-05 2.E-05 3.E-05 4.E-05 Cond
1.2
1.0 WS

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05
0 5 10 15 20 0.E+00 2.E-05 4.E-05 6.E-05 Cond
1.2
1.0 WS

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05
Cond
1.2
0 5 10 15 20 0.E+00 2.E-05 4.E-05 6.E-05
WS
1.0
Fraction of Total Packed Height

1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(c)
Figure H.3.3: CO2 and H2S vapour and liquid phase column profiles for the second set of plant data.
(a) CO2 train #2. (b) CO2 train #3. (c) CO2 train #4. (Cond = condenser, WS = wash section, Reb =
reboiler)

A-103
Appendix H: HYSYS® Simulation Results

CO2
CO2 CO2
CO2 Data
Data H2S
H2S H
H2S
2S Data
Data

Regenerator H2S (ppm) Regenerator H2S Loading


Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05
Cond
1.2
0 5 10 15 20 0.E+00 1.E-05 2.E-05 3.E-05 4.E-05
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(a)
Regenerator H2S (ppm) Regenerator H2S Loading
Absorber H2S (ppm) Absorber H2S Loading 0 20 40 60 80 0.E+00 5.E-06 1.E-05 2.E-05 2.E-05
Cond
1.2
0 5 10 15 20 0.E+00 1.E-05 2.E-05 3.E-05 4.E-05
WS
1.0

Fraction of Total Packed Height


1.0
Fraction of Total Packed Height

0.8
0.8

0.6 0.6

0.4
0.4

0.2
0.2
0.0
Reb
0.0
0 5 10 15 20 0.2 0.4 0.6 0.8 1.0 -0.2
10 30 50 70 90 0.2 0.4 0.6 0.8
Absorber CO2 (mol%) Absorber CO2 Loading
Regenerator CO2 (mol%) Regenerator CO2 Loading

(b)
Figure H.3.4: CO2 and H2S vapour and liquid phase column profiles for the second set of plant data.
(a) CO2 train #5. (b) CO2 train #6. (Cond = condenser, WS = wash section, Reb = reboiler)

A-104
Appendix H: HYSYS® Simulation Results

Vapour Phase Liquid Phase Vapour Data Liquid Data

1.2
Cond 1.2
Cond
WS WS

Fraction of Total Packed Height

Fraction of Total Packed Height


1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
30 50 70 90 110 130 60 70 80 90 100 110 120 80 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(a) (b)
1.2
Cond 1.2
Cond
WS WS
Fraction of Total Packed Height

Fraction of Total Packed Height


1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
80 90 100 110 120 130 60 70 80 90 100 110 120 80 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(c) (d)
1.2
Cond
WS
Fraction of Total Packed Height

1.0

0.8

0.6

0.4

0.2

0.0
Reb
-0.2
80 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C)

(e)
Figure H.3.5: Column temperature profiles for the first set of plant data. (a) CO2 train #2. (b) CO2 train #3.
(c) CO2 train #4. (d) CO2 train #5. (e) CO2 train #6. (Cond = condenser, WS = wash section, Reb =
reboiler)

A-105
Appendix H: HYSYS® Simulation Results

Vapour Phase Liquid Phase Vapour Data Liquid Data

1.2
Cond 1.2
Cond
WS WS

Fraction of Total Packed Height

Fraction of Total Packed Height


1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
30 50 70 90 110 130 60 70 80 90 100 110 120 80 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(a) (b)
1.2
Cond 1.2
Cond
WS WS
Fraction of Total Packed Height

Fraction of Total Packed Height


1.0 1.0

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
Reb Reb
-0.2 -0.2
80 90 100 110 120 130 60 70 80 90 100 110 120 80 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C) Absorber Temperature (°C) Regenerator Temperature (°C)

(c) (d)
1.2
Cond
WS
Fraction of Total Packed Height

1.0

0.8

0.6

0.4

0.2

0.0
Reb
-0.2
80 90 100 110 120 130 60 70 80 90 100 110 120

Absorber Temperature (°C) Regenerator Temperature (°C)

(e)
Figure H.3.6: Column temperature profiles for the second set of plant data. (a) CO2 train #2. (b) CO2 train
#3. (c) CO2 train #4. (d) CO2 train #5. (e) CO2 train #6. (Cond = condenser, WS = wash section, Reb =
reboiler)

A-106
Appendix I: Process Control Studies of the CO2 Trains

APPENDIX I

PROCESS CONTROL STUDIES


OF THE CO2 TRAINS

This appendix provides examples of the MATLAB® scripts used in the process control studies for CO2
trains #1 and #7. Also included are the step response curves generated from the closed-loop testing
of the optimal diagonal control structures for the two CO2 trains.

A-107
Appendix I: Process Control Studies of the CO2 Trains

I.1 Selection of Diagonal Control Structure


An example of the MATLAB® scripts used to calculate the various sensitivity analysis indices is
provided below.

% Program sensitivity.m
% Compares alternative diagonal control structures for CO2 Train #1
% at high gas throughput
% Calculates CN, MRI, DCN and DC
%
% Notation:
% alt1 = SGC-LSF
% alt2 = SGC-RSF
% alt3 = SGC-RLL
% y1 = Sweet gas CO2 content
% y2 = Sweet gas flow rate
% d1 = Raw gas CO2 content
%
clear
%
% Process gain matrices
KpH1alt1=[1.1695 -0.0635; 1.0881 -0.0026];
KpH1alt2=[1.1695 -0.9885; 1.0881 -0.0262];
KpH1alt3=[1.1695 0.0412; 1.0881 0.0032];
%
% Process time constant matrices
TpH1alt1=[8.2231 2.2825; 1.6168 3.3742];
TpH1alt2=[8.2231 13.1691; 1.6168 16.0241];
TpH1alt3=[8.2231 10.2357; 1.6168 9.4975];
%
% Process dead time matrices
DTH1alt1=[0.1667 0.1667; 0.1667 0.1667];
DTH1alt2=[0.1667 0.6667; 0.1667 0.8333];
DTH1alt3=[0.1667 51.9408; 0.1667 69.1521];
%
% Disturbance gain, time constant and dead time vectors
KdH1=[0.5757; -0.0735];
TpdH1=[6.1206; 3.4381];
DTdH1=[0.1667; 0.1667];

%
% Unit magnitude disturbance in d1 and unit magnitude setpoint changes
% in y1 and y2
Dd1=[1];
Dy=eye(2);
%
for i=1:2
for j=1:2
% 3rd order Pade approximations for dead times
[DTH1alt1num(i,:,j),DTH1alt1den(i,:,j)]=pade(DTH1alt1(i,j),3);
[DTH1alt2num(i,:,j),DTH1alt2den(i,:,j)]=pade(DTH1alt2(i,j),3);
[DTH1alt3num(i,:,j),DTH1alt3den(i,:,j)]=pade(DTH1alt3(i,j),3);
%
[DTdH1num(i,:),DTdH1den(i,:)]=pade(DTdH1(i),3);
%
% Process transfer functions
GpH1alt1(i,j)=tf(KpH1alt1(i,j)*DTH1alt1num(i,:,j),
conv([TpH1alt1(i,j) 1],DTH1alt1den(i,:,j)));
GpH1alt2(i,j)=tf(KpH1alt2(i,j)*DTH1alt2num(i,:,j),
conv([TpH1alt2(i,j) 1],DTH1alt2den(i,:,j)));
GpH1alt3(i,j)=tf(KpH1alt3(i,j)*DTH1alt3num(i,:,j),
conv([TpH1alt3(i,j) 1],DTH1alt3den(i,:,j)));
%
% Disturbance transfer functions
GdH1(i,1)=tf(KdH1(i)*DTdH1num(i,:),conv([TpdH1(i) 1],DTdH1den(i,:)));

A-108
Appendix I: Process Control Studies of the CO2 Trains

end
end
%
% Steady-state analysis
% Singular value decomposition
SVDH1(:,1)=svd(KpH1alt1);
SVDH1(:,2)=svd(KpH1alt2);
SVDH1(:,3)=svd(KpH1alt3);
%
% MRI values
MRIH1(1)=min(SVDH1(:,1));
MRIH1(2)=min(SVDH1(:,2));
MRIH1(3)=min(SVDH1(:,3))
%
% CN values
CNH1(1)=cond(KpH1alt1);
CNH1(2)=cond(KpH1alt2);
CNH1(3)=cond(KpH1alt3)
%
% DCN and DC values
DCH1(1,1)=norm(inv(KpH1alt1)*KdH1*Dd1);
DCH1(1,2)=norm(inv(KpH1alt2)*KdH1*Dd1);
DCH1(1,3)=norm(inv(KpH1alt3)*KdH1*Dd1);
%
DCNH1(1,1)=DCH1(1,1)/norm(KdH1*Dd1)*max(SVDH1(:,1));
DCNH1(1,2)=DCH1(1,2)/norm(KdH1*Dd1)*max(SVDH1(:,2));
DCNH1(1,3)=DCH1(1,3)/norm(KdH1*Dd1)*max(SVDH1(:,3));
%
for i=1:2
DCH1(i+1,1)=norm(inv(KpH1alt1)*Dy(:,i));
DCH1(i+1,2)=norm(inv(KpH1alt2)*Dy(:,i));
DCH1(i+1,3)=norm(inv(KpH1alt3)*Dy(:,i));
%
DCNH1(i+1,1)=DCH1(i+1,1)/norm(Dy(:,i))*max(SVDH1(:,1));
DCNH1(i+1,2)=DCH1(i+1,2)/norm(Dy(:,i))*max(SVDH1(:,2));
DCNH1(i+1,3)=DCH1(i+1,3)/norm(Dy(:,i))*max(SVDH1(:,3));
end
DCNH1, DCH1
%
% Frequency analysis
w=logspace(-5,2,100);
%
% Evaluate process and disturbance transfer functions over frequency range
GpH1alt1_w=freqresp(GpH1alt1,w);
GpH1alt2_w=freqresp(GpH1alt2,w);
GpH1alt3_w=freqresp(GpH1alt3,w);
GdH1_w=freqresp(GdH1,w);
%
% Calculate indices over frequency range and print to text file
for i=1:length(w)
% Singular value decomposition
SVDH1_w(:,1,i)=svd(GpH1alt1_w(:,:,i));
SVDH1_w(:,2,i)=svd(GpH1alt2_w(:,:,i));
SVDH1_w(:,3,i)=svd(GpH1alt3_w(:,:,i));
%
% MRI values
for j=1:3
MRIH1_w(j,i)=min(SVDH1_w(:,j,i));
end
%
% CN values
CNH1_w(1,i)=cond(GpH1alt1_w(:,:,i));
CNH1_w(2,i)=cond(GpH1alt2_w(:,:,i));
CNH1_w(3,i)=cond(GpH1alt3_w(:,:,i));
%
% DCN and DC values
DCH1alt1_w(1,i)=norm(inv(GpH1alt1_w(:,:,i))*GdH1_w(:,:,i)*Dd1);
DCH1alt2_w(1,i)=norm(inv(GpH1alt2_w(:,:,i))*GdH1_w(:,:,i)*Dd1);

A-109
Appendix I: Process Control Studies of the CO2 Trains

DCH1alt3_w(1,i)=norm(inv(GpH1alt3_w(:,:,i))*GdH1_w(:,:,i)*Dd1);
%
DCNH1alt1_w(1,i)=DCH1alt1_w(1,i)/norm(GdH1_w(:,:,i)*Dd1)*max(SVDH1_w(:,1,i));
DCNH1alt2_w(1,i)=DCH1alt2_w(1,i)/norm(GdH1_w(:,:,i)*Dd1)*max(SVDH1_w(:,2,i));
DCNH1alt3_w(1,i)=DCH1alt3_w(1,i)/norm(GdH1_w(:,:,i)*Dd1)*max(SVDH1_w(:,3,i));
%
for j=1:2
DCH1alt1_w(j+1,i)=norm(inv(GpH1alt1_w(:,:,i))*Dy(:,j));
DCH1alt2_w(j+1,i)=norm(inv(GpH1alt2_w(:,:,i))*Dy(:,j));
DCH1alt3_w(j+1,i)=norm(inv(GpH1alt3_w(:,:,i))*Dy(:,j));
%
DCNH1alt1_w(j+1,i)=DCH1alt1_w(j+1,i)/norm(Dy(:,j))*max(SVDH1_w(:,1,i));
DCNH1alt2_w(j+1,i)=DCH1alt2_w(j+1,i)/norm(Dy(:,j))*max(SVDH1_w(:,2,i));
DCNH1alt3_w(j+1,i)=DCH1alt3_w(j+1,i)/norm(Dy(:,j))*max(SVDH1_w(:,3,i));
end
end
y=[w; MRIH1_w;];
fid=fopen('MRIH1_w.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f %12.8f\n',y);
fclose(fid);
y=[w; CNH1_w];
fid=fopen('CNH1_w.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f %12.8f\n',y);
fclose(fid);
y=[w; DCNH1alt1_w; DCNH1alt2_w; DCNH1alt3_w];
fid=fopen('DCNH1_w.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f %12.8f %12.8f %12.8f %12.8f %12.8f %12.8f
%12.8f\n',y);
fclose(fid);
y=[w; DCH1alt1_w; DCH1alt2_w; DCH1alt3_w];
fid=fopen('DCH1_w.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f %12.8f %12.8f %12.8f %12.8f %12.8f %12.8f
%12.8f\n',y);
fclose(fid);

A-110
Appendix I: Process Control Studies of the CO2 Trains

I.2 Selection of Diagonal Control Structure Configuration


I.2.1 MATLAB® Script for Stability Analysis
An example of the MATLAB® scripts used to analyse system stability is provided below. The
calculated system poles and zeros are given in Table I.2.1.

% Program Stability.m
% Determines the system poles and zeros for CO2 train #1 at high gas throughput
%
clear
%
% Process gain, time constant and dead time matrices
KpH1=[1.1695 -0.9885; 1.0881 -0.0262];
TpH1=[8.2231 13.1691; 1.6168 16.0241];
DTH1=[0.1667 0.6667; 0.1667 0.8333];
%
for i=1:2
for j=1:2
% Find 3rd order Pade approximations for dead times
[DTH1num(i,:,j),DTH1den(i,:,j)]=pade(DTH1(i,j),3);
%
% Create system transfer function matrices
GpH1(i,j)=tf(KpH1(i,j)*DTH1num(i,:,j),conv([TpH1(i,j) 1],DTH1den(i,:,j)));
end
end
%
% Obtain state-space model coefficient matrices
% and calculate system poles and zeros
[AH1,BH1,CH1,DH1]=ssdata(GpH1);
PH1=eig(AH1)
ZH1=tzero(AH1,BH1,CH1,DH1)

Table I.2.1: System poles and zeros for the SGC-RSF control structure.

CO2 Train #1 CO2 Train #7


High Gas Medium Gas Low Gas High Gas Medium Gas Low Gas
Throughput Throughput Throughput Throughput Throughput Throughput
System Poles -0.12 -0.07 -0.04 -0.23 -0.10 -0.05
-0.62 -0.43 -0.36 -0.48 -0.36 -0.34
-22.06±21.05i -22.06±21.05i -22.06±21.05i -22.06±21.05i -22.06±21.05i -22.06±21.05i
-27.86 -27.86 -27.86 -27.86 -27.86 -27.86
-5.52±5.26i -5.52±5.26i -5.52±5.26i -5.52±5.26i -5.52±5.26i -5.52±5.26i
-6.97 -6.97 -6.97 -6.97 -6.97 -6.97
-0.08 -0.06 -0.05 -0.10 -0.07 -0.05
-4.41±4.21i -4.41±4.21i -4.41±4.21i -11.03±10.53i -11.03±10.53i -11.03±10.53i
-5.57 -5.57 -5.57 -13.93 -13.93 -13.93
-0.06 -0.05 -0.05 -0.10 -0.07 -0.06

System Zeros -4.41±4.20i -4.41±4.20i -4.41±4.21i -15.33 -11.27±10.34i -11.09±10.48i


-5.57 5.52±5.26i -5.57 -11.83±9.70i -14.25 -14.01
5.53±5.26i 6.97 5.52±5.26i 5.89±4.90i 5.61±5.16i 5.54±5.24i
6.97 -5.57 6.97 6.18 6.77 6.92
-0.06 -0.05 -0.05 -0.09 -0.07 -0.06
-0.12 -0.07 -0.04 -0.21 -0.09 -0.05
22.06±21.05i 22.06±21.05i 22.06±21.05i 22.06±21.05i 22.06±21.05i 22.06±21.05i
27.86 27.86 27.86 27.86 27.86 27.86

A-111
Appendix I: Process Control Studies of the CO2 Trains

I.2.2 MATLAB® Script for RGA Calculations


An example of the MATLAB® scripts used for the RGA calculations is provided below.

% Program RGA.m
% Calculates the RGA for CO2 train #1 at high gas throughput
%
clear
%
% Process gain, time constant and dead time matrices for SGC-RSF control structure
KpH1=[1.1695 -0.9885; 1.0881 -0.0262];
TpH1=[8.2231 13.1691; 1.6168 16.0241];
DTH1=[0.1667 0.6667; 0.1667 0.8333];
%
for i=1:2
for j=1:2
% 3rd order Pade approximations for dead times
[DTH1num(i,:,j),DTH1den(i,:,j)]=pade(DTH1(i,j),3);
%
% Process transfer functions
GpH1(i,j)=tf(KpH1(i,j)*DTH1num(i,:,j),conv([TpH1(i,j) 1],DTH1den(i,:,j)));
end
end
%
% Steady-state RGA
RGAH1=KpH1.*inv(KpH1).'
%
% Calculate RGA over frequency range and print to text file
w = logspace(-5,2,100);
%
GpH1_w=freqresp(GpH1,w);
%
for i=1:length(w)
RGAH1_w(:,:,i)=GpH1_w(:,:,i).*inv(GpH1_w(:,:,i)).';
end
%
for i=1:2
RGAH1r1(i,:)=sign(RGAH1(1,i))*abs(RGAH1_w(1,i,:));
RGAH1r2(i,:)=sign(RGAH1(2,i))*abs(RGAH1_w(2,i,:));
end
%
y=[w; RGAH1r1; RGAH1r2];
fid=fopen('RGAH1_w.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f %12.8f %12.8f\n',y);
fclose(fid);

A-112
Appendix I: Process Control Studies of the CO2 Trains

I.2.3 MATLAB® Script for Stability Index Calculations


An example of the MATLAB® scripts used for the stability index calculations is provided below.

% Program StabilityIndices.m
% Completes variable pairing analysis for CO2 Train #1 at high gas throughput
% Calculates NI and MIC stability indices
%
clear
%
% Reordered process gain matrices for SGC-RSF control structure
% i.e. paired variables according to RGA along diagonal
KpH1=[-0.9885 1.1695; -0.0262 1.0881];
%
%
% Reordered process gain matrices with positive diagonal elements
KpH1p=[0.9885 1.1695; 0.0262 1.0881];
%
% Calculate NI values
NIH1=det(KpH1)/(KpH1(1,1)*KpH1(2,2))
%
% Determine eigenvectors and eigenvalues for positive diagonal gain matrices
[KpH1pEVec, KpH1pEVal] = eig(KpH1p);
%
% Calculate MIC values
MICH1=diag(KpH1pEVal)

A-113
Appendix I: Process Control Studies of the CO2 Trains

I.3 Analysis of Diagonal Control Structure Performance


An example of the MATLAB® scripts used for the PRGA and CLDG calculations is provided below.

% Program PerformanceAnalysis.m
% Analyses performance of control structure for CO2 Train #1 at high gas throughput
% Calculates PRGA and CLDG
%
clear
%
% Reordered process gain, time constant and dead time matrices
% i.e. paired variables according to RGA along diagonal
KpH1=[-0.9885 1.1695; -0.0262 1.0881];
TpH1=[13.1691 8.2231; 16.0241 1.6168];
DTH1=[0.6667 0.1667; 0.8333 0.1667];
%
% Disturbance gain, time constant and dead time vectors
KdH1=[0.5757; -0.0735];
TpdH1=[6.1206; 3.4381];
DTdH1=[0.1667; 0.1667];
%
for i=1:2
for j=1:2
% 3rd order Pade approximations for dead times
[DTH1num(i,:,j),DTH1den(i,:,j)]=pade(DTH1(i,j),3);
[DTdH1num(i,:),DTdH1den(i,:)]=pade(DTdH1(i),3);
%
% Process and disturbance transfer functions
GpH1(i,j)=tf(KpH1(i,j)*DTH1num(i,:,j),conv([TpH1(i,j) 1],DTH1den(i,:,j)));
GdH1(i,1)=tf(KdH1(i)*DTdH1num(i,:),conv([TpdH1(i) 1],DTdH1den(i,:)));
end
end
%
% Steady-state PRGA and CLDG
PRGAH1=diag(diag(KpH1))*inv(KpH1)
CLDGH1=PRGAH1*KdH1
%
% Frequency analysis
w=logspace(-5,2,100);
%
% Evaluate process and disturbance transfer functions over frequency range
GpH1_w=freqresp(GpH1,w);
GdH1_w=freqresp(GdH1,w);
%
% Calculate PRGA and CLDG over frequency range and print to text file
for i=1:length(w)
% PRGA
PRGAH1_w(:,:,i)=diag(diag(GpH1_w(:,:,i)))*inv(GpH1_w(:,:,i));
%
for j=1:2
PRGAH1r1_w(j,i)=PRGAH1_w(1,j,i);
PRGAH1r2_w(j,i)=PRGAH1_w(2,j,i);
end
%
% CLDG
CLDGH1_w(:,i)=PRGAH1_w(:,:,i)*GdH1_w(:,:,i);
end
y=[w; PRGAH1r1_w; PRGAH1r2_w
fid=fopen('PRGAH1_w.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f %12.8f %12.8f\n',y);
fclose(fid);
y=[w; CLDGH1_w];
fid=fopen('CLDGH1_w.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f\n',y);
fclose(fid);

A-114
Appendix I: Process Control Studies of the CO2 Trains

I.4 BLT Tuning


An example of the MATLAB® scripts used to perform BLT tuning is provided below.

% Program BLT.m
% Performs BLT tuning for CO2 Train #1 at high gas throughput
%
clear
%
% Reordered process gain, time constant and dead time matrices
% i.e. paired variables according to RGA along diagonal
KpH1=[-0.9885 1.1695; -0.0262 1.0881];
TpH1=[13.1691 8.2231; 16.0241 1.6168];
DTH1=[0.6667 0.1667; 0.8333 0.1667];
%
% Ziegler-Nichols PI settings
KcZNH1=diag([-14.3942 7.2173]);
TiZNH1=diag([1.1883 0.2914]);
%
% Set up tuning parameters
i=sqrt(-1);
w=logspace(-3,2,500);
s=i*w;
f=2.1;
df=0.01;
loop=0;
flagm=-1;
flagp=-1;
dbmax=-100;
%
for i=1:2
for j=1:2
% 3rd order Pade approximations for dead times
[DTH1num(i,:,j),DTH1den(i,:,j)]=pade(DTH1(i,j),3);
[DTdH1num(i,:),DTdH1den(i,:)]=pade(DTdH1(i),3);
%
% Process transfer functions
GpH1(i,j)=tf(KpH1(i,j)*DTH1num(i,:,j),conv([TpH1(i,j) 1],DTH1den(i,:,j)));
end
end
%
% BLT tuning: vary f until dbmax=4 (because 2x2 system)
while abs(dbmax-4)>0.01
KcH1=KcZNH1/f;
TiH1=TiZNH1*f;
%
% Controller transfer function
for i=1:2
GcH1(i,i)=tf(KcH1(i,i)*[TiH1(i,i) 1],[TiH1(i,i) 0]);
end
%
% Evaluate Gp and Gc over frequency range
GpH1_w=freqresp(GpH1,w);
GcH1_w=freqresp(GcH1,w);
%
% Calculate W function
for i=1:length(w)
WnyquistH1(i)=-1+det(eye(2)+GpH1_w(:,:,i)*GcH1_w(:,:,i));
lc(i)=WnyquistH1(i)/(1+WnyquistH1(i));
dbcl(i)=20*log10(abs(lc(i)));
end
%
% Identify peak in closed-loop log modulus
[dbmax,nmax]=max(dbcl);
wmax=w(nmax);
%

A-115
Appendix I: Process Control Studies of the CO2 Trains

loop=loop+1;
if loop>50, break, end
%
% Test if +4 dB criterion is satisfied, if not get new value for f
if dbmax>4
if flagp>0
df=df/2;
end
flagm=1;
f=f+df;
else
if flagm>0
df=df/2;
end
flagp=1;
f=f-df;
if f<1
f=1;
end
end
end
%
% Output to screen
loop, dbmax, f, KcH1, TiH1
%
% Output to text file
y=[w; real(WnyquistH1); imag(WnyquistH1)];
fid=fopen('WnyquistH1.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f\n',y);
fclose(fid);

A-116
Appendix I: Process Control Studies of the CO2 Trains

I.5 Diagonal Control Structure Dynamic Simulations


I.5.1 MATLAB® Script for Dynamic Simulations
An example of the MATLAB® scripts used to perform the dynamic simulations of the RGF-RSF
diagonal control structure is provided below.

% DynSim1.m
% Performs dynamic simulations for CO2 train #1 at high gas throughput
%
clear
%
% Reordered process gain, time constant and dead time matrices
% i.e. paired variables according to RGA along diagonal
KpH1=[-0.9885 1.1695; -0.0262 1.0881];
TpH1=[13.1691 8.2231; 16.0241 1.6168];
DTH1=[0.6667 0.1667; 0.8333 0.1667];
%
% Disturbance gain, time constant and dead time vectors
KdH1=[0.5757; -0.0735];
TpdH1=[6.1206; 3.4381];
DTdH1=[0.1667; 0.1667];
%
% BLT tuned controller settings for reordered system
% Include settings for medium and low gas throughput
KcH1=diag([-6.7578 3.3884]);
KcM1=diag([-7.5029 4.1056]);
KcL1=diag([-8.1480 4.3767]);
TiH1=diag([2.5311 0.6207]);
TiM1=diag([2.9964 0.7399]);
TiL1=diag([3.3639 0.8320]);
%
for i=1:2
for j=1:2
% 3rd order Pade approximations for dead times
[DTH1num(i,:,j),DTH1den(i,:,j)]=pade(DTH1(i,j),3);
[DTdH1num(i,:),DTdH1den(i,:)]=pade(DTdH1(i),3);
%
% Process, disturbance and controller transfer functions
GpH1(i,j)=tf(KpH1(i,j)*DTH1num(i,:,j),conv([TpH1(i,j) 1],DTH1den(i,:,j)));
GdH1(i,1)=tf(KdH1(i)*DTdH1num(i,:),conv([TpdH1(i) 1],DTdH1den(i,:)));
GcH1(i,i)=tf(KcH1(i,i)*[TiH1(i,i) 1],[TiH1(i,i) 0]);
GcM1(i,i)=tf(KcM1(i,i)*[TiM1(i,i) 1],[TiM1(i,i) 0]);
GcL1(i,i)=tf(KcL1(i,i)*[TiL1(i,i) 1],[TiL1(i,i) 0]);
end
end
%
w=logspace(-5,2,100);
%
% Evaluate process, disturbance and controller transfer functions over frequency
range
GpH1_w=freqresp(GpH1,w);
GdH1_w=freqresp(GdH1,w);
GcH1_w=freqresp(GcH1,w);
GcM1_w=freqresp(GcM1,w);
GcL1_w=freqresp(GcL1,w);
%
% Calculate system loop transfer functions and output to text file
for i=1:length(w)
for j=1:2
% Using H controller settings
LH1H_w(j,i)=GcH1_w(j,j,i)*GpH1_w(j,j,i);
%
% Using M controller settings
LH1M_w(j,i)=GcM1_w(j,j,i)*GpH1_w(j,j,i);

A-117
Appendix I: Process Control Studies of the CO2 Trains

%
% Using L controller settings
LH1L_w(j,i)=GcL1_w(j,j,i)*GpH1_w(j,j,i);
end
end
y=[w; LH1H_w];
fid=fopen('LH1H_w_v4.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f\n',y);
fclose(fid);
y=[w; LH1M_w];
fid=fopen('LH1M_w_v4.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f\n',y);
fclose(fid);
y=[w; LH1L_w];
fid=fopen('LH1L_w_v4.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f\n',y);
fclose(fid);
%
% Consider step response to unit setpoint change
% Closed-loop transfer functions with H controller settings
GclH1H=inv(eye(2)+GpH1*GcH1)*GpH1*GcH1;
%
% Closed-loop transfer functions with M controller settings
GclH1M=inv(eye(2)+GpH1*GcM1)*GpH1*GcM1;
%
% Closed-loop transfer functions with L controller settings
GclH1L=inv(eye(2)+GpH1*GcL1)*GpH1*GcL1;
%
T=[0:0.05:60];
%
% H controller settings
[YH1H,T]=step(GclH1H,T);
%
% M controller settings
[YH1M,T]=step(GclH1M,T);
%
% L controller settings
[YH1L,T]=step(GclH1L,T);
%
for i=1:length(T)
for j=1:2
YH1Hr1(j,i)=YH1H(i,1,j);
YH1Hr2(j,i)=YH1H(i,2,j);
%
YH1Mr1(j,i)=YH1M(i,1,j);
YH1Mr2(j,i)=YH1M(i,2,j);
%
YH1Lr1(j,i)=YH1L(i,1,j);
YH1Lr2(j,i)=YH1L(i,2,j);
end
end
%
% Output to text file
y=[T'; YH1Hr1; YH1Hr2];
fid=fopen('StepH1H.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f %12.8f %12.8f\n',y);
fclose(fid);
y=[T'; YH1Mr1; YH1Mr2];
fid=fopen('StepH1M.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f %12.8f %12.8f\n',y);
fclose(fid);
y=[T'; YH1Lr1; YH1Lr2];
fid=fopen('StepH1L.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f %12.8f %12.8f\n',y);
fclose(fid);
%
% Consider step response to unit disturbance
% Closed-loop transfer functions with H controller settings

A-118
Appendix I: Process Control Studies of the CO2 Trains

GcldH1H=inv(eye(2)+GpH1*GcH1)*GdH1;
%
% Closed-loop transfer functions with M controller settings
GcldH1M=inv(eye(2)+GpH1*GcM1)*GdH1;
%
% Closed-loop transfer functions with L controller settings
GcldH1L=inv(eye(2)+GpH1*GcL1)*GdH1;
%
T=[0:0.05:60];
%
% H controller settings
[YdH1H,T]=step(GcldH1H,T);
%
% M controller settings
[YdH1M,T]=step(GcldH1M,T);
%
% L controller settings
[YdH1L,T]=step(GcldH1L,T);
%
for i=1:length(T)
for j=1:2
YdH1Hr1(j,i)=YdH1H(i,j);
YdH1Mr1(j,i)=YdH1M(i,j);
YdH1Lr1(j,i)=YdH1L(i,j);
end
end
%
% Output to text file
y=[T'; YdH1Hr1];
fid=fopen('StepdH1H.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f\n',y);
fclose(fid);
y=[T'; YdH1Mr1];
fid=fopen('StepdH1M.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f\n',y);
fclose(fid);
y=[T'; YdH1Lr1];
fid=fopen('StepdH1L.txt','w');
fprintf(fid,'%12.8f %12.8f %12.8f\n',y);
fclose(fid);

A-119
Appendix I: Process Control Studies of the CO2 Trains

I.5.2 Step Response Curves


The step response curves for the RGF-RSF diagonal control structure for CO2 trains #1 and #7 at the
medium and low gas throughput conditions are given below.

A-120
Appendix I: Process Control Studies of the CO2 Trains
High gas throughput controller settings Medium gas throughput controller settings Low gas throughput controller settings

0.120 1.500 0.300

1.200
Sweet Gas CO2 Content

Sweet Gas CO2 Content

Sweet Gas CO2 Content


0.080 0.200
0.900

0.040 0.600 0.100

0.300
0.000 0.000
0.000

-0.040 -0.300 -0.100


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60
A-121

Time (min) Time (min) Time (min)


0.000 0.006 1.400

0.005 1.200
-0.001

Sweet Gas Flow Rate


Sweet Gas Flow Rate

Sweet Gas Flow Rate

0.004 1.000
-0.002
0.003 0.800
-0.003
0.002 0.600
-0.004
0.001 0.400

-0.005 0.000 0.200

-0.006 -0.001 0.000


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60

Time (min) Time (min) Time (min)

(a) (b) (c)


Figure I.5.1: CO2 train #1 closed-loop step response curves at the medium gas throughput conditions. (a) Unit magnitude change in the raw gas CO2 content at
0 min. (b) Unit magnitude change in the sweet gas CO2 content setpoint at 0 min. (c) Unit magnitude change in the sweet gas flow rate setpoint at 0 min.
Appendix I: Process Control Studies of the CO2 Trains
High gas throughput controller settings Medium gas throughput controller settings Low gas throughput controller settings

0.300 1.500 1.000

1.200 0.800
Sweet Gas CO2 Content

Sweet Gas CO2 Content

Sweet Gas CO2 Content


0.200
0.900 0.600

0.100 0.600 0.400

0.300 0.200
0.000
0.000 0.000

-0.100 -0.300 -0.200


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60
A-122

Time (min) Time (min) Time (min)


0.001 0.016 1.400

0.000 0.012 1.200


Sweet Gas Flow Rate

Sweet Gas Flow Rate

Sweet Gas Flow Rate


-0.001 1.000
0.008
-0.002 0.800
0.004
-0.003 0.600
0.000
-0.004 0.400

-0.005 -0.004 0.200

-0.006 -0.008 0.000


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60

Time (min) Time (min) Time (min)

(a) (b) (c)


Figure I.5.2: CO2 train #7 closed-loop step response curves at the medium gas throughput conditions. (a) Unit magnitude change in the raw gas CO2 content at
0 min. (b) Unit magnitude change in the sweet gas CO2 content setpoint at 0 min. (c) Unit magnitude change in the sweet gas flow rate setpoint at 0 min.
Appendix I: Process Control Studies of the CO2 Trains
High gas throughput controller settings Medium gas throughput controller settings Low gas throughput controller settings

0.100 1.500 0.250

0.080 0.200
1.200
Sweet Gas CO2 Content

Sweet Gas CO2 Content

Sweet Gas CO2 Content


0.060 0.150
0.900
0.040 0.100
0.600
0.020 0.050
0.300
0.000 0.000

0.000
-0.020 -0.050

-0.040 -0.300 -0.100


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60
A-123

Time (min) Time (min) Time (min)


0.000 0.006 1.400

1.200
-0.001
Sweet Gas Flow Rate

Sweet Gas Flow Rate

Sweet Gas Flow Rate


0.004
1.000

-0.002
0.800
0.002
0.600
-0.003

0.400
0.000
-0.004
0.200

-0.005 -0.002 0.000


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60

Time (min) Time (min) Time (min)

(a) (b) (c)


Figure I.5.3: CO2 train #1 closed-loop step response curves at the low gas throughput conditions. (a) Unit magnitude change in the raw gas CO2 content at 0 min.
(b) Unit magnitude change in the sweet gas CO2 content setpoint at 0 min. (c) Unit magnitude change in the sweet gas flow rate setpoint at 0 min.
Appendix I: Process Control Studies of the CO2 Trains
High gas throughput controller settings Medium gas throughput controller settings Low gas throughput controller settings

0.200 1.500 0.350

1.200 0.280
Sweet Gas CO2 Content

Sweet Gas CO2 Content

Sweet Gas CO2 Content


0.150

0.900 0.210
0.100

0.600 0.140

0.050
0.300 0.070

0.000
0.000 0.000

-0.050 -0.300 -0.070


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60
A-124

Time (min) Time (min) Time (min)


0.001 0.008 1.400

0.000 1.200
0.006
Sweet Gas Flow Rate

Sweet Gas Flow Rate

Sweet Gas Flow Rate


-0.001 1.000

0.004
-0.002 0.800

-0.003 0.600
0.002

-0.004 0.400
0.000
-0.005 0.200

-0.006 -0.002 0.000


0 10 20 30 40 50 60 0 10 20 30 40 50 60 0 10 20 30 40 50 60

Time (min) Time (min) Time (min)

(a) (b) (c)


Figure I.5.4: CO2 train #7 closed-loop step response curves at the low gas throughput conditions. (a) Unit magnitude change in the raw gas CO2 content at 0 min.
(b) Unit magnitude change in the sweet gas CO2 content setpoint at 0 min. (c) Unit magnitude change in the sweet gas flow rate setpoint at 0 min.

You might also like