You are on page 1of 246

Development of Advanced Amine

Systems with Accurate Vapour-liquid


Equilibrium Measurement

Danlu Tong

A thesis submitted for the degree of


Doctor of Philosophy of the University of London and
Diploma of Membership of Imperial College

Energy Engineering Group


Department of Chemical Engineering and Chemical Technology
Imperial College London
London SW7 2AZ, United Kingdom

2012

Declaration of Originality
I, Danlu Tong, hereby certify that this thesis has been written by me, based on
the original work conducted by me in the Department of Chemical Engineering
at Imperial College London between 2008 and 2012, except where specifically
acknowledged in the text.

Abstract
Imperative pressure on containing anthropogenic greenhouse gas (GHG) emissions,
particularly CO2, has led to the development and deployment of low-carbon technologies
across the globe. Carbon capture and storage (CCS) is widely considered as an indispensible
part of the technology portfolio alongside energy efficiency, renewables and nuclear. Due to
its operational flexibility and technology maturity, post-combustion capture (PCC) is most
likely to be one of the first adopted capture processes. However, the associated capital
expenditure (CAPEX) and operating cost (OPEX) of the current state-of-the-art PCC process
are seemed as the major barriers for its commercialisation.
This thesis investigated the relationship between the chemical structure of the amines and
their CO2 cyclic loading capacity through solubility measurements with two types of
apparatuses: a static-analytical apparatus and a static-synthetic setup. The former was first
validated by measuring the solubility of CO2 in 30 mass% aqueous monoethanolamine (MEA)
solutions at

= (313 and 393) K and a total pressure range between (11 and 415) kPa. After

successful validation of the apparatus, new CO2 solubility data for a sterically-hindered
amine, 2-Amino-2-Methyl-1-Propanol (AMP) were obtained at temperatures between (313
and 393) K and a total pressure range of (23 to 983) kPa. CO2 solubility data for a tertiary
amine, 2-Dimethylaminoethanol (DMMEA) were also measured with the analytical apparatus
from

= 313 K to

= 393 K and a total pressure of (6 to 616) kPa. Apart from the three

single amines studied, this work investigated the blended amine systems and the influence of
an activator (piperazine or PZ) on the promoted amines. For AMP + PZ blends, solubility
measurements of CO2 in 25 mass% AMP + 5 mass% PZ and 20 mass% AMP + 10 mass%
PZ were conducted from (313 to 393) K; while for the DMMEA + PZ mixture, 25 mass%
DMMEA + 5 mass% PZ system was studied.
The static-synthetic setup employed in this work was first validated through measuring the
solubility of CO2 in water from (313 to 393) K. A calculation algorithm implemented in
Microsoft Excel as VBA was used to convert the experimental data to Henrys constants
which were then compared with literature data. After the validation, CO2 solubility of 25
mass% MDEA was measured at 313 K.

4
Another aspect of this work involved building quasi-chemical thermodynamic models to
interpret the experimental data obtained in both apparatuses. This work used two different
methods including Deshmukh-Mather model based on activity ()-fugacity () approach and
a more simplified Kent-Eisenberg model to represent the single and mixed amine systems
respectively. Despite the compromise in the correlation quality compared to the - approach,
Kent-Eisenberg model has received wide popularity due to its simplicity. In this work, the
correlation results from the two models were compared. The applications of the models to
predict solution phase speciation and CO2 solubility behaviour not measured in this work
were also discussed.

Acknowledgements
Spending three-and-half years on a PhD project is a significant commitment. This experience
is an invaluable asset in my life and will not be so rich and colourful without the
contributions from numerous people, to whom I would like to express my sincere
appreciation.
First and foremost, my sincere gratitude undoubtedly goes to my supervisors: Paul, Martin
and Geoff. Your immense support and encouragement throughout my PhD were always the
sources of momentum that led me through the obstacles. Thank you also Jon, Niall, and
Mathieu, for the inspiring and thoughtful discussions on the CCS-related issues. Thank you
Nigel and John, for your patience and kind guidance on any query I had. And to all the fellow
colleagues in Pauls and Martins groups, thank you for the inspiring discussions as well as
fun social activities that made my life more colourful. Thank you very much the Granthams,
for providing the financial assistance that allow me to complete my PhD. Thank you all the
people from the Grantham Institute, Brian, Simon, Neil, and Ajay, for the invaluable
discussions on the climate change and energy-related issues. Finally, special thanks to Zheng,
your tremendous support is very much appreciated.

Success is the ability to go from one failure to another with no loss


of enthusiasm.

-Winston Churchill

Contents
Chapter 1 Introduction.......................................................................................................... 18
1.1 Motivation for carbon capture ........................................................................................ 18
1.2 Scope and objectives of this work .................................................................................. 21
1.3 Outline of thesis ............................................................................................................. 22
1.4 Key contributions of this work ....................................................................................... 23

Chapter 2 Literature Review: Amine Scrubbing in the Context of Carbon Capture ..... 25
2.1 Carbon capture technologies .......................................................................................... 26
2.1.1 Post-combustion capture.......................................................................................... 26
2.1.2 Pre-combustion capture ........................................................................................... 28
2.1.3 Oxy-combustion capture.......................................................................................... 31
2.1.4 Other capture technologies ...................................................................................... 32
2.2 Types of amines ............................................................................................................. 35
2.3 Experimental apparatus for VLE measurement ............................................................. 36
2.3.1 Static-analytical apparatus ....................................................................................... 38
2.3.2 Static-synthetic apparatus ........................................................................................ 43
2.3.3 Flow systems ........................................................................................................... 45
2.3.4 Other methods.......................................................................................................... 46
2.4 Thermodynamic models ................................................................................................. 47
2.4.1 Solution chemistry ................................................................................................... 48
2.4.2 Thermodynamic framework .................................................................................... 48
2.4.3 Important VLE model .............................................................................................. 50

8
Chapter 3 Solubility Measurements for CO2 in Aqueous Amines with a Static-analytical
Apparatus ............................................................................................................................... 59
3.1 Introduction .................................................................................................................... 60
3.2 Original design of the apparatus..................................................................................... 60
3.2.1 Layout of the static-analytical apparatus ................................................................. 60
3.2.2 Gas and liquid sampling systems............................................................................. 62
3.3 Revised design of apparatus ........................................................................................... 68
3.4 Experimental procedures ................................................................................................ 69
3.5 Validation of the apparatus ............................................................................................ 70
3.5.1 Validation of the pressure and temperature sensors ................................................ 70
3.5.2 Calibration of the GC system .................................................................................. 71

Chapter 4 Experimental Results from the Static-analytical Apparatus for CO2 solubility
in aqueous amines .................................................................................................................. 76
4.1 Introduction .................................................................................................................... 77
4.2 Experimental uncertainty ............................................................................................... 78
4.3 Validation of the static-analytical apparatus by measuring CO2 solubility in aqueous 30
mass% MEA solutions ......................................................................................................... 78
4.3.1 MEA in the context of carbon capture..................................................................... 78
4.3.2 Validation results ..................................................................................................... 82
4.4 Solubility of CO2 in aqueous AMP and AMP+PZ solutions ......................................... 84
4.4.1 Sterically-hindered amines in the context of carbon capture .................................. 84
4.4.2 Experimental results on CO2 solubility in aqueous AMP solutions ........................ 91
4.4.3 Experimental results on CO2 solubility in aqueous mixtures of AMP and PZ........ 93
4.5 Solubility of CO2 in DMMEA and DMMEA solutions ................................................. 95
4.5.1 Tertiary amines in the context of carbon capture .................................................... 96

9
4.5.2 Experimental results on CO2 solubility in aqueous DMMEA and DMMEA+PZ
solutions .......................................................................................................................... 100
4.6 Comparisons of the CO2 solubility in amines solutions............................................... 102
4.6.1 30 mass% MEA vs 30 mass% AMP ..................................................................... 103
4.6.2 30 mass% AMP vs 25 mass% AMP + 5 mass% PZ vs 20 mass% AMP + 10 mass%
PZ.................................................................................................................................... 104
4.6.3 30 mass% MEA vs 30 mass% AMP vs 30 mass% DMMEA ............................... 107
4.6.4 30 mass% DMMEA vs 25 mass% DMMEA + 5 mass% PZ ................................ 110

Chapter 5 Thermodynamic Models for the Solubility of CO2 in Single and Blended
Amine Systems ..................................................................................................................... 112
5.1 Introduction .................................................................................................................. 113
5.2 - model for CO2 solubility in single amine solutions ............................................... 114
5.2.1 Physical and chemical equilibria ........................................................................... 114
5.2.2 Phase nonideality ................................................................................................... 117
5.2.3 Balance equations .................................................................................................. 119
5.2.4 - model for MEA-H2O-CO2............................................................................... 119
5.2.5 - model for AMP-H2O-CO2 ............................................................................... 123
5.3 Kent-Eisenberg model for AMP-based and DMMEA-based blended amine systems 126
5.3.1 Physical and chemical equilibria ........................................................................... 126
5.3.2 Phase nonideality ................................................................................................... 128
5.3.3 Balance equations .................................................................................................. 129
5.3.4 Kent-Eisenberg model for AMP-PZ-H2O-CO2 ..................................................... 130
5.3.5 Kent-Eisenberg model for DMMEA-PZ-H2O-CO2 .............................................. 135
5.4 Thermodynamic models for CO2 solubility in aqueous MDEA solutions ................... 139
5.4.1 Kent-Eisenberg model for MDEA-H2O-CO2 ........................................................ 139
5.4.2 - model for MDEA-H2O-CO2 ............................................................................ 142

10
Chapter 6 Solubility Measurements for H2O-CO2 and MDEA-H2O-CO2 Systems with
Synthetic Apparatus ............................................................................................................ 145
6.1 Introduction .................................................................................................................. 146
6.2 Apparatus and experimental procedures ...................................................................... 149
6.2.1 Apparatus ............................................................................................................... 149
6.2.2 Experimental procedures ....................................................................................... 153
6.3 Results and discussions ................................................................................................ 156
6.3.1 Calculation of the Henrys constant with graphical interpretation ........................ 156
6.3.2 Results from the graphical interpretation method ............................................... 158
6.3.3 Calculation of the Henrys constant with theoretical modelling ........................... 159
6.3.4 Sources of uncertainties ......................................................................................... 165
6.3.5 Uncertainty analysis .............................................................................................. 165
6.4 Measurement of the solubility of CO2 in aqueous MDEA system with the synthetic
apparatus............................................................................................................................. 166
6.4.1 Methodology for the calculation of CO2 solubility from experimental
measurements ................................................................................................................. 167
6.4.2 Results ................................................................................................................... 170
6.4.3 Other sources of uncertainties and limitations ...................................................... 174

Chapter 7 Conclusions and Future Work ......................................................................... 175


7.1 Summary of this work .................................................................................................. 176
7.1.1 Static-analytical apparatus ..................................................................................... 176
7.1.2 Quasi-chemical modelling ..................................................................................... 178
7.1.3 Synthetic apparatus ................................................................................................ 179
7.2 Conclusions .................................................................................................................. 180
7.2.1 CO2 absorption in 30 mass% aqueous AMP and AMP+PZ blends ...................... 181
7.2.2 CO2 absorption in 30 mass% aqueous DMMEA and DMMEA+PZ blends ......... 183

11
7.2.3 Relationship between the structure of amine and the solubility of CO2 ................ 185
7.2.4 Conclusions from the synthetic apparatus ............................................................. 185
7.3 Future work .................................................................................................................. 185

Chapter 8 Nomenclature ..................................................................................................... 188

Bibliography ......................................................................................................................... 191

Appendices ............................................................................................................................ 220


I Summary of typical amines .......................................................................................... 221
II Sample gPROMS codes (MEA, DeshmukhMather model) ................................... 222
III Calibration of the synthetic apparatus ....................................................................... 233
IV Experimental procedure of the synthetic apparatus .................................................. 236
V Plots of experimental results from the synthetic apparatus ........................................ 239
VI Thermal expansion of Hastelloy................................................................................ 241
VII Sample VBA code for calculation of the Henrys constant from the synthetic
apparatus ......................................................................................................................... 242

12

List of Figures
Figure 1.1 Key technologies for reducing CO2 emissions under the BLUE Map scenario .... 19
Figure 2.1 Post-combustion CO2 capture................................................................................ 26
Figure 2.2 Pre-combustion CO2 capture ................................................................................. 29
Figure 2.3 Oxy-combustion CO2 capture................................................................................ 31
Figure 2.4 Schematic diagram of Ca looping process ............................................................ 33
Figure 2.5 Schematic diagram of CLC process ...................................................................... 34
Figure 2.6 FTIR coupled VLE apparatus ................................................................................ 43
Figure 2.7 Equilibrium vessels for medium-pressure measurements .................................... 46
Figure 3.1 Schematic diagram of the static-analytic apparatus. ............................................. 61
Figure 3.2 GC signal for H2O-acetone solution with the tubing unheated ............................. 63
Figure 3.3 Tubing connecting LSV and vaporisation chamber being wrapped with heating
tape ........................................................................................................................................... 63
Figure 3.4 GC signal for MEA-H2O with the tubing heated to 473 K ................................... 64
Figure 3.5 GC signal for CO2-H2O-MEA with the tubing heated to 523 K ........................... 65
Figure 3.6 GC signal for CO2-H2O-MEA through direct sampling with the manual syringe 66
Figure 3.7 GC signal for MEA-H2O-CO2 with the tubing unheated ...................................... 67
Figure 3.8 Schematic diagram of the static-analytic apparatus .............................................. 68
Figure 3.9 Comparison of water vapour pressure measurements from this work .................. 71
Figure 3.10 Relation between GC peak area and molar amount for CO2 ............................... 72
Figure 3.11 Relation between GC peak area and molar amount for H2O ............................... 73
Figure 3.12 GC diagram for CO2-H2O at

= 313 K, total

= 1884 kPa, and with tubing

heated to 423 K ........................................................................................................................ 74


Figure 4.1 Comparisons of solubility data of CO2 in 30 mass% aqueous MEA solutions ..... 82
Figure 4.2 Solubility of CO2 in 30 mass% aqueous AMP solutions....................................... 93

13
Figure 4.3 CO2 solubility in aqueous AMP and PZ blends .................................................... 95
Figure 4.4 CO2 solubility in aqueous DMMEA and DMMEA + PZ.................................... 102
Figure 4.5 CO2 solubility in aqueous MEA and AMP .......................................................... 103
Figure 4.6 CO2 solubility in aqueous MEA and AMP .......................................................... 104
Figure 4.7 CO2 solubility in aqueous AMP and AMP + PZ blends ..................................... 105
Figure 4.8 CO2 solubility in aqueous AMP and AMP + PZ blends ..................................... 106
Figure 4.9 CO2 solubility in aqueous MEA, AMP and DMMEA ......................................... 108
Figure 4.10 CO2 solubility in aqueous MEA, AMP and DMMEA ...................................... 109
Figure 4.11 CO2 solubility in aqueous DMMEA and DMMEA + PZ blend ........................ 110
Figure 4.12 CO2 solubility in aqueous DMMEA and AMP based systems ......................... 111
Figure 5.1 Schematic diagram for the CO2 dissolution into aqueous amine system ............ 114
Figure 5.2 Correlation of experimental results from this work and Jou et al. ...................... 120
Figure 5.3 Parity plot between experimental and model loadings ........................................ 121
Figure 5.4 Liquid phase speciation in 30 mass% MEA at 313 K ......................................... 122
Figure 5.5 Predictions from this work compared with Jou et al.s experimental results ...... 123
Figure 5.6 Correlation of experimental results for AMP-H2O-CO2 system from this work . 124
Figure 5.7 Parity plot between experimental and model loadings ........................................ 125
Figure 5.8 Liquid phase speciation in 30 mass% AMP at 373 K from model predictions ... 125
Figure 5.9 Correlation of experimental results for AMP-PZ-H2O-CO2 system.................... 131
Figure 5.10 Correlation of experimental results for AMP-H2O-CO2 system........................ 131
Figure 5.11 Parity plot between the experimental data from this work and the model data for
30 mass% AMP, 25 mass% AMP + 5 mass% PZ, and 20 mass% AMP + 10 mass% PZ
systems ................................................................................................................................... 132
Figure 5.12 Liquid phase speciation in aqueous 25 mass% AMP + 5 mass% PZ solution at
313 K from model predictions ............................................................................................... 133
Figure 5.13 Liquid phase speciation in aqueous 20 mass% AMP + 10 mass% PZ solution at
313 K from model predictions ............................................................................................... 133

14
Figure 5.14 Liquid phase speciation in aqueous 30 mass% AMP solution at 313 K from
model predictions. (a) normal scale for y-axis; (b) logarithmic scale for y-axis ................... 134
Figure 5.15 Correlation of experimental results for DMMEA-PZ-H2O-CO2 system from this
work ....................................................................................................................................... 136
Figure 5.16 Correlation of experimental results for DMMEA-H2O-CO2 system from this
work ....................................................................................................................................... 136
Figure 5.17 Parity plot between the experimental data from this work and the model data for
30 mass% DMMEA, and 25 mass% DMMEA + 5 mass% PZ systems ............................... 137
Figure 5.18 Liquid phase speciation in aqueous 30 mass% DMMEA + PZ solution at 313 K
from model predictions .......................................................................................................... 138
Figure 5.19 Liquid phase speciation in aqueous 30 mass% DMMEA solution at 313 K from
model predictions ................................................................................................................... 138
Figure 5.20 Correlation of experimental results for MDEA-H2O-CO2 system from SidiBoumedine et al ..................................................................................................................... 140
Figure 5.21 Correlation of experimental results for MDEA-H2O-CO2 system from this work
at 316 K .................................................................................................................................. 141
Figure 5.22 Liquid phase speciation in aqueous 25 mass% MDEA solution at 313 K from
model predictions ................................................................................................................... 142
Figure 5.23 Correlation of experimental results for MDEA-H2O-CO2 system from SidiBoumedine et al. .................................................................................................................... 143
Figure 5.24 Correlation of experimental results for MDEA-H2O-CO2 system from this work
at 316 K .................................................................................................................................. 144
Figure 5.25 Parity plot between the experimental data from this work and the model data for
30 mass% MDEA systems ..................................................................................................... 144
Figure 6.1 Schematic diagram for the synthetic apparatus ................................................... 150
Figure 6.2 Assembly view for the equilibrium vessel and heater shell ................................ 151
Figure 6.3 Schematic diagram of a high pressure generator ................................................. 152
Figure 6.4 Change of total pressure plotted against time...................................................... 155
Figure 6.5 Total pressure against cumulative moles of water injected at

= 313 K ........... 155

15
Figure 6.6 Total pressure against cumulative moles of solvent for MDEA-H2O-CO2 at

313 K...................................................................................................................................... 167


Figure 6.7 Comparison of CO2 solubility in 25 mass% MDEA at

= 313 K (total pressure).

................................................................................................................................................ 172
Figure 6.8 Comparison of CO2 solubility in 25 mass% MDEA at

= 313 K (partial pressure)

................................................................................................................................................ 173
Figure 7.1 pH of AMP, DMMEA, MDEA and PZ at 296 K ................................................ 184

16

List of Tables
Table 2.1 Typical apparatus and analytical methods employed by previous studies .............. 37
Table 3.1 Calibration results of relative response factor at 313 K .......................................... 75
Table 4.1 Experimental conditions conducted in this work .................................................... 77
Table 4.2 Summary of literature for CO2 solubility in aqueous MEA solution ...................... 81
Table 4.3 Comparison of CO2 (1) solubility data in a solution of water (2) (70 mass%) and
MEA (3) (30 mass%) at

= (313 and 393) K from Jou et al. and this work .......................... 83

Table 4.4 Summary of literature for CO2 solubility in aqueous AMP and AMP+PZ blends . 91
Table 4.5 CO2 (1) solubility data in a solution of water (2) (70 mass%) and AMP (3) (30
mass%) at

= (313, 333, 353, 373 and 393) K ....................................................................... 92

Table 4.6 CO2 (1) solubility data in a solution of water (2) (70 mass%), AMP (3) (25 mass%)
and PZ (4) (5 mass%) at

= (313, 333, 373 and 393) K ........................................................ 94

Table 4.7 CO2 (1) solubility data in a solution of water (2) (70 mass%), AMP (3) (20 mass%)
and PZ (4) (10 mass%) at

= (313, 333, 373 and 393) K ...................................................... 94

Table 4.8 CO2 (1) solubility data in a solution of water (2) (70 mass%), and DMMEA (3) (30
mass%) at

= (313, 333, 353, 373 and 393) K ..................................................................... 101

Table 4.9 CO2 (1) solubility data in a solution of water (2) (70 mass%), DMMEA (3) (25
mass%), and PZ (4) (5 mass% ) at

= (313, 333, 373 and 393) K ....................................... 101

Table 4.10 Comparisons of mole-ratio loadings for three aqueous AMP-based solutions ... 105
Table 4.11 Comparisons of mass-ratio loadings for three aqueous AMP-based solutions ... 107
Table 4.12 Comparisons of mass-ratio loadings for aqueous MEA, AMP and DMMEA
solutions ................................................................................................................................. 109
Table 4.13 Comparisons of mass-ratio loadings for AMP, AMP+PZ and DMMEA+PZ
systems ................................................................................................................................... 111
Table 5.1 Models applied to the amine systems in this work ( = was applied in this work;
= was not applied) .................................................................................................................. 114
Table 5.2 Equilibrium and Henrys constants used in this work for reactions I-VI ............. 116

17
Table 5.3 Parameters used in eqs. (5.9) and (5.10) to calculate vapour pressure and partial
molar volume of water from Saul and Wagner ...................................................................... 117
Table 5.4 Parameters of eq. (5.14) for pure components coefficient .................................... 117
Table 5.5 Parameters for cross coefficient used in eq. (5.15) ............................................... 117
Table 5.6 Parameters used in eq. (5.18) ................................................................................ 118
Table 5.7 Binary interaction parameters for MEA-H2O-CO2 system ................................... 120
Table 5.8 Binary interaction parameters for the AMP-H2O-CO2 system ............................. 124
Table 5.9 Equilibrium constants used in the Kent-Eisenberg model for reaction VII-X, and V
for DMMEA (all values are in mol/kg basis) ........................................................................ 128
Table 5.10 Adjustable parameters in eq. (5.35) for the AMP-PZ-H2O-CO2 system ............ 130
Table 5.11 Adjustable parameters in eq. (5.35) for the DMMEA-PZ-H2O-CO2 system ...... 135
Table 5.12 MDEA protonation constant described in eq. (5.8) ............................................ 139
Table 5.13 Regressed parameters in eq. (5.43) ..................................................................... 139
Table 5.14 Binary interaction parameters for MDEA-H2O-CO2 system in the - model ... 142
Table 6.1 Matrix for experimental conditions ....................................................................... 156
Table 6.2 Constants in eq. (6.11) .......................................................................................... 157
Table 6.3 Comparison between experimental results and literature values .......................... 158
Table 6.4 Constants in eq. (6.23) .......................................................................................... 160
Table 6.5 Constants in eq. (6.24) .......................................................................................... 160
Table 6.6 Comparison of the Henrys constants obtained using the model of section 6.3.3
with literature values .............................................................................................................. 164
Table 6.7 Constants in eq. (6.41) .......................................................................................... 168
Table 6.8 Constants in eq. (6.43) .......................................................................................... 169
Table 6.9 CO2 (1) solubility data in a solution of water (2) (75 mass%) and MDEA (3) (25
mass%) at

= 316 K ............................................................................................................. 170

Table 7.1 Literature data on the reaction kinetics of CO2 in aqueous amines at 298 K........ 182

1. Introduction

18

Chapter 1
Introduction
1.1 Motivation for carbon capture

The scientific debate over climate change, including the identification of CO2 as the major
greenhouse gas (GHG) and the principal cause of rising global average temperature, is widely
considered to have reached consensus [1]. It is also generally accepted that most of the
observed increase in global average temperature since the mid-20th century is very likely due
to the observed increase in anthropogenic GHG, particularly CO2, concentrations [2]. The
focus has now shifted from if we should to how to decarbonise our global energy system
[3] with numerous studies carried out around the world devising what our future energy
systems look like [4-9].

Of the many studies of energy scenarios, the most well known are those from the
International Energy Agency (IEA). In their biannual publication, Energy Technology
Perspectives (ETP) [4], a baseline scenario and a BLUE Map1 scenario for the world energy
system from present to year 2050 were conceived. The baseline scenario, which assumes that
no new energy or climate change policies, beyond those already in place, are introduced
during the scenario period, leads to almost a doubling in the CO2 emissions in 2050 compared
1

In contrast to the Baseline scenario which assumes governments introduce no new energy and climate policies,
the BLUE Map scenario is target-oriented: it sets the goal of halving global energy-related CO2 emissions by
2050 (compared to 2005 levels) and examines the least-cost means of achieving that goal through the
deployment of existing and new low-carbon technologies.

1. Introduction

19

to 2007. When effective policy incentives and technology advancement are envisaged, the
energy related CO2 in 2050 could be cut to half of the level in 2007, on track to achieve the
long-term global target of limiting the global average temperature rise to between 2 C and 3
C (CO2 concentration of 450-550 ppm).
Reducing the CO2 emissions in 2050 from 57 Gt in the baseline scenario by more than 3/4th
to 14 Gt in the BLUE Map scenario requires drastic actions to be taken in almost every
energy-related sector. As can be seen from graph 1.1, efficiency improvements in end-use
fuel and electricity utilisation is the biggest contributor, accounting for a 38% reduction.
Following that, carbon capture and storage (CCS) is expected to contribute 19% to the total
emissions cuts envisioned. More specifically, 10% is from CCS in power generation while
the remaining 9% is from CCS applied to industry and transformation. Increasing the use of
renewable energy and end-use fuel switching will potentially lead to 17% and 15%
reductions respectively. To a lesser extent, nuclear and power generation efficiency and fuel
switching each accounted for 6% and 5%.

Figure 1.1 Key technologies for reducing CO2 emissions under the BLUE Map scenario [4]
Undoubtedly, CCS is an essential part of the technology portfolio required to achieve the
target of limiting the global average temperature rise to 3 C, at least according to the BLUE
Map scenario from the IEA. Without it, the overall costs to reduce CO2 emissions increase by
as much as 70% [10]. However, the envisaged road map is ambitious and requires over 3 000
projects (equivalent to 10 Gt of CO2 captured) by 2050. To achieve this, it is estimated that
expenditure of over US$ 2.5-3 trillion will be needed between 2010 and 2050 [10].

1. Introduction

20

While it is clear that the individual stages of CCS are technically viable, the challenges lie in
integrating and scaling up these technologies from capture and compression, to transport and
storage; these can only be addressed through commercial-scale CCS projects. As summarised
by the global CCS Institute in their December 2011 update report on the global status of
large-scale integrated CCS projects (LSIPs) 2 [11], there are 74 LSIPs currently identified
around the world, including 15 that are currently operating or in construction, and a further
59 in the planning stages of development. All those combined results in a capture capacity of
more than 157 Mt a-1. Although this development showed the determination of government
and industry to advance CCS technology at a commercial scale, more substantial, timely and
stable policy support, including a carbon price signal is needed [12].
The successful operation of LSIPs around the globe indicates that the technology risks and
technical barriers are not insurmountable. What is hindering the deployment of the CCS
technology at the commercial scale is largely attributed to the cost, notably in the capture and
compression stage, which contribute to between 2/3rd 3/4th of the total CCS cost. Of the
many carbon capture technologies, amine scrubbing is the most mature and was first patented
in 1930 [13]. Fitting an amine scrubbing unit to the power plant will typically lead to a 10%
net reduction in the thermal efficiency or approximately 30% less efficiency of electricity
production compared to the original. For a typical amine plant, over 50% of the cost results
from the energy consumption, of which the reboiler heat and the compression work
requirements take equal shares [14]. In addition, the capital cost, which is largely dependent
upon the size of the absorber, is principally determined by the reaction kinetics between the
aqueous amine solvent and the CO2. While process integration and optimisation can play an
important role in the economic optimisation, it is the nature of the solvent which determines
the thermodynamic and kinetic limits of the CO2 capture process. Thus, the greatest potential
for reducing the cost resides in the development of efficient amine solvents or blends for
post-combustion capture by solvent scrubbing.

According to the Global CCS Institute, Large-scale integrated projects are defined as those which involve the capture,
transport and storage of CO2 at a scale of:
* not less than 800 000 tonnes of CO2 annually for a coal-based power plant; and
* not less than 400 000 tonnes of CO2 annually for other emission-intensive industrial facilities (including natural gas-based
power generation).

1. Introduction

21

1.2 Scope and objectives of this work


This work was focused around amine scrubbing technology, concentrating on the vapourliquid equilibria of various amine-H2O-CO2 systems both experimentally and via
thermodynamic modelling. The major objectives of this work are to develop an experimental
apparatus and method to measure accurately the solubility of CO2 in different aqueous amine
systems at typical post-combustion capture conditions. The equipment developed was a
static-analytic vapour-liquid equilibrium apparatus, with on-line gas chromatography for
composition determination. The amine systems studied in this work covered three types
including: monoethanolamine (MEA) which is a primary amine, 2-Amino-2-methyl-1propanol (AMP) known as a sterically-hindered amine, and 2-Dimethylaminoethanol which
is a tertiary amine, as well as the piperazine (PZ) activated blends of the latter two (AMP +
PZ and DMMEA + PZ). These amine systems are carefully selected through research on
published literature, according to the process economic potential and scarcity of the existing
database. More details relating to the rationale behind the solvents selection can be found in
Chapter 7 Conclusions and Future Work.
As a part of collaborating work, the experimental data from this study will also provide a
validation basis for the SAFT-VR model which was recently developed into rigorous
thermodynamic models for the MEA- and AMP-based systems [15]. In addition to this, this
work also developed its own thermodynamic models based on quasi-chemical theory
including: the Kent-Eisenberg model popularly used in industry; and a rigorous - model for
treating liquid and vapour phases non-ideality with separate theoretical formulations. The
models can be used to correlate the experimental results and provide speciation information
in the solution phase; besides, in certain cases it can also be used to predict CO2 solubility in
conditions other than those covered by the experimental data.
From the experimental and modelling results, this work aimed to gain insights into the
relations between the CO2 solubility and the chemical structures or more simply, types of
amines and amine blends. Eventually, the results from this work can be combined with the
SAFT-VR model, to form an integral approach in the design and development of efficient
amine solvent for carbon capture process [15].
This project also used a synthetic apparatus to measure the CO2 solubility in aqueous amine
systems. CO2-H2O was studied in the first place and the system was measured from 313 K to
393 K. A theoretical framework was employed to analyse the raw experimental data to

1. Introduction

22

determine Henrys constant, which allowed the comparison between results from this work
and literature values. This work subsequently explored the application of the synthetic
apparatus to an amine system, in this case aqueous MDEA solution.
1.3 Outline of thesis
This thesis is organised into four main sections covering the experiments conducted with the
static-analytical apparatus and the synthetic setup, as well as the models developed to
interpret the experimental data from these two apparatuses. The synthetic setup and the
modelling work are each covered in one chapter. More space is devoted to the static-analytic
apparatus because, as a newly constructed apparatus in this work, considerable efforts were
devoted to adjusting the design and operation including developing reliable methodology to
carry out the measurements. It is therefore worthwhile to elaborate on the methodology of
evolution towards the final version of the experimental design and procedure by separating
them as a separate section. The remaining chapters are summarised as follows.

In Chapter 2, a comprehensive literature review is presented in four parts. The first


part covers different carbon capture technologies including the three main categories
(post-, pre-, and oxy-combustion capture) and some advanced capture systems (Ca
looping and chemical looping). In the second part, the types of amines employed in
carbon capture are summarised. The next subsection is concerning the historical
development of VLE apparatus and three categories are used to group them: staticanalytical apparatus, static-synthetic apparatus and flow-type system. Finally, a
general thermodynamic framework for the amine-H2O-CO2 system and commonly
used vapour-liquid equilibrium models are discussed.

In Chapter 3, the design of the static-analytical apparatus is covered. Details of the


obstacles encountered in the initial stages and the subsequent modifications to the
setup are also included. In the end, the calibration methodology of the analytical
device (GC) is elaborated.

Continue from the previous chapter, Chapter 4 is devoted to the experimental results
from the static-analytical apparatus. This chapter is structured according to the types
of amine systems studied in this work with the final section summarises and compares
the cyclic loading of these amines and blends.

In Chapter 5, thermodynamic modelling is discussed including the general framework


widely employed to describe the phase and chemical equilibria. The two

1. Introduction

23

thermodynamic models, one based on the Kent-Eisenberg model [16] and the other
built upon the - approach and their applications in the amine systems studied in this
work are presented.

In Chapter 6, work relating to the synthetic setup is discussed. As an independent


work, this chapter can be read as standalone. In the beginning, literatures on the CO2H2O are summarised and the correlation equations of the Henrys constant derived by
Carroll et al. [17] and Crovetto and Mather [18] are given. This is followed by the
detailed description of the synthetic apparatus and procedures. The experimental
results produced from this setup for the CO2-H2O systems are also included which are
converted to Henrys constant with a model built in Excel VBA. Finally, experiments
carried out for the MDEA-H2O-CO2 system are presented.

In Chapter 7, we draw the conclusions of this work and provide recommendations for
further work.

1.4 Key contributions of this work


The main contributions of this work are listed below.
A static-analytical apparatus was constructed and subsequently employed to measure
accurately the solubility of CO2 in various types of amine and blended amine systems. An
innovative liquid-phase sampling system was devised in order to solve the amine adsorption
and carbamate precipitation issues which had led to the incomplete analysis of the liquid
phase components, particularly CO2 and amine. An in-situ calibration method was
developed in this work to allow quantitative determination of the solution loading.
We then developed two types of thermodynamic models to correlate the experimental data
obtained in this work. A rigorous - approach, using an extended Debye-Hckel expression
proposed by Guggenheim and Stokes [19] for the activity coefficients and a virial equation of
state truncated after the second term for the vapour phase non-ideality, was developed for the
single amine systems. Meanwhile, for blended amine systems, a modified version of the
Kent-Eisenberg model which has been widely deployed in industry was employed. In general,
both models are capable of providing reasonable correlations, although the - approach can
be more confidently applied in prediction beyond the experimental conditions. However, the
less rigorous Kent-Eisenberg model with its simplicity is easier to be incorporated into a
process model than the more computationally difficult - approach. In addition, as quasi-

1. Introduction

24

chemical models, both methods can provide phase speciation information, showing the
concentration profiles of different ions as the solution loading increased.
VLE data are vital for thermodynamic modelling, process design and amine selection;
nevertheless, owing to the difficulty in the VLE measurements for the amine systems, the
reported data from previous studies often suffer from discrepancies. For some amines, the
available solubility data are scattered. Therefore, the accurate measurements produced from
this work are valuable additions to the existing database.
We achieved the initial objective of identifying the influence of chemical structure on the
CO2 solubility for selected amine systems from our experimental measurements. Although
more studies are needed to rationalize the solvent selection process, the experimental
apparatus and methodology developed in this work laid the foundation for a more in-depth
collaboration with the SAFT-VR model in the pursuit of development and design of efficient
amine systems for the carbon capture process.
This work also successfully validated a novel synthetic apparatus and a theoretical method
used to treat the data via studying the phase behaviour of CO2-H2O system. Attempts have
been made to expand the application to amine systems.

2. Literature Review

25

Chapter 2

Literature Review: Amine


Scrubbing in the Context of
Carbon Capture
This chapter summarises previous studies on carbon capture, focusing on amine scrubbing
technology. A general overview of carbon capture technologies will be presented at the
beginning, including post-, pre-, oxy-combustion capture, high-temperature looping and
adsorption using amine functionalized materials. This will be followed by an introduction of
different types of amines which have been used in carbon capture processes and their
chemical properties. Next, a discussion on the evolution of experimental apparatus for
measuring vapour-liquid equilibrium, particularly for amine-H2O-CO2 systems, will be
presented. Finally, some of the thermodynamic models developed in the literatures will also
be discussed, especially in terms of their respective advantages and drawbacks.

2. Literature Review

26

2.1 Carbon capture technologies


Carbon capture, transport and storage (CCTS) refers to a series of technologies and processes
which capture CO2 gas from point sources, such as power stations and industrial plants, then
compress it to a supercritical phase and transport it via pipelines or ships, finally storing it
underground in depleted coal and oil reservoirs or saline aquifers. Apart from storing the CO2
permanently underground, it can also be used in enhanced oil recovery or as a feedstock for
chemical processes [15]. Generally, the capture processes are grouped into three categories,
namely: post-combustion, pre-combustion, and oxy-combustion.
2.1.1 Post-combustion capture
Post-combustion is an end-of-pipe process which separates CO2 from flue gas, mainly
consisting of N2, CO2, water vapour, and other impurities, e.g. SOx, NOx. Amine scrubbing is
the most commonly referred to and the most mature process (Fig. 2.1).

Figure 2.1 Post-combustion CO2 capture [20]


The flue gas, with dust and sulphur impurities removed, is sent to an absorber where CO2 is
selectively absorbed by amine solvent at a temperature between 313 K and 333 K. The
remaining flue gas, mostly N2, is then released from the top of the column, while the CO2
enriched solution is transferred to a second column, called a stripper. In the stripping column,
temperature is maintained at between 373 K and 393 K with the heat drawn from the power
plant. The reaction in the absorber is reversed at this higher temperature which results in the

2. Literature Review

27

release of CO2 (over 99% pure) and lean solvent (containing a small amount of CO2). This
regenerated solvent is recycled to the absorber while CO2 gas is ready for the next steps, i.e.
compression and transportation.
Amine scrubbing is a well-understood and widely employed technology. The first process
was invented by Bottoms in 1930 [13]. There are hundreds of plants relying on this
technology to remove CO2 from low oxygen gas streams, such as natural gas and hydrogen
[21]. Over 20 plants are using 20 mass% to 30 mass% MEA in an oxidising environment
such as pilot-scale coal-fired power plants [21]. Apart from MEA, other solvents such as
aqueous ammonia, amino acid salts, and proprietary amines (e.g. KS-1) have also been
successfully employed to capture CO2 from flue gas. The Global CCS Institute [22]
summarised several near-term industrial post-combustion technologies including: Fluors
Econamine FG PLUS [23], Mitsubishi Heavy Industries KS solvent, Cansolv Technologies,
Aker Clean Carbon, and Alstoms Chilled Ammonia process (ACAP). These post
combustion capture technologies will be discussed in more detail below.
Fluors Econamine FG PlusSM technology is based on aqueous monoethanolamine with
proprietary inhibitors to control corrosion, oxidative degradation and thermal degradation.
The addition of these inhibitors allows higher concentration of MEA to be used, i.e. 30-35
mass% compared to 18-20 mass% which combined with process optimisation (such as heat
integration, inter-cooling and split flow configuration) could reduce steam consumption by
over 30% compared to generic MEA technology. Econamine FG PlusSM is often used as a
representative technology for economic evaluation of post-combustion technologies.
Unlike Econamine FG PlusSM technology which has been mainly targeted to natural gas
combustion processes, MHI KM-COR process has been employed in coal-based flue gas
treatments. It uses a proprietary sterically-hindered solvent, KS-1, which is claimed to offer a
number of advantages compared to conventional MEA solvent including: exceptionally low
corrosiveness which means no corrosion inhibitor is needed; lower degradation, less solvent
loss, and greater cyclic capacity compared to MEA. All of these features combined with
process improvement could lead to a 30% reduction in steam consumption. [24]
A more recent technology is Alstoms Chilled Ammonia Process (CAP) which uses the
chemical equilibrium between ammonia, ammonia bicarbonate and ammonia carbonate [25].
As ammonia is a low-cost solvent and is resistant to degradation, this allows its exposure to
higher temperatures and exposure to sulphur impurities. This process has been employed in

2. Literature Review

28

AEPs Mountaineer project and CO2 is captured at 80-90% capture efficiency and at 99.9+ %
purity. Owing to the low temperature leading to slower reaction rate, the absorber in the CAP
is usually larger than others.
Cansolv, owned by Shell Global Solutions International B.V., has developed aqueous
diamine solvents to selectively remove CO2, SOx, NOx, and mercury. It has been chosen as a
part of a retrofit of a 150 MWe at SaskPowers Boundary Dam. The reported capture rate has
achieved 90%.
Other than the above-mentioned technologies, Just Catch technology developed by Aker
Clean Carbon, has been installed at Scottish Powers Longannet station and reported to
achieve an energy penalty reduction target of about one-third. Doosan Babcock Energy and
HTC Purenergy [22] are working together to commercialise a mixed amine solvent initially
developed at the University of Regina which has reduced regeneration energy to less than 1.0
tonne steam per tonne CO2 (about 30% improvement over conventional 30 mass% MEA).
There are also amino acid salt based processes jointly developed by Siemens Energy and
TNO. Amino acid salts are believed to possess a number of advantages compared to amines
including larger cyclic capacity, greater absorption rate, smaller heat of regeneration,
improved stability to oxygen, lower volatility and less environmental impact [26].
Apart from solvent-based absorption, post-combustion capture technologies also include solid
adsorption processes and membranes [27] which are at an earlier phase of development
compared to absorption processes. Since the focus of this study is on amine scrubbing, these
technologies will not be discussed further here.
2.1.2 Pre-combustion capture
In pre-combustion capture, the principal reaction involves that coal, natural gas or biomass is
partially oxidised with steam to form syngas, with reaction (2.2) referred to as steam
reforming reaction (other side reactions are also possible, such as combustion and pyrolysis):
(I)
(II)
Carbon monoxide can be further oxidised to produce more hydrogen, known as the water
gas-shift reaction:
(III)

2. Literature Review

29

The gasification reactions are conducted at around 2-7 MPa resulting in a gas mixture of CO2
and H2 at much higher pressure than the post-combustion counterpart (0.2 MPa) and the CO2
concentration is also considerably higher (15-60% by volume compared to 10-15% for postcombustion) [28]. This substantially higher partial pressure of CO2 allows the use of physical
solvents instead of chemical solvents to separate the two gases. Unlike chemical absorbents
which react with acid gases, physical solvents dissolve them preferentially without forming
new chemical bonds. Physical solvent-based processes are uneconomical for low pressure
processes owing to the large capital cost (a large absorber is required for the low solubility of
the gases). However it is more advantageous for high pressure systems due to the
significantly lower operating cost resulting from the weaker reaction of the solvents and acid
gases. Fig. 2.2 illustrates the pre-combustion capture process.

Figure 2.2 Pre-combustion CO2 capture [29]


Pre-combustion capture can be applied to natural gas reforming, coal gasification and IGCC
processes. Current and planned projects for these applications are summarised here.

2. Literature Review

30

IGCC involves converting coal into syngas, which is then combusted in a gas turbine. The
process achieves extremely high thermal efficiency and significantly improved environmental
performance compared to a conventional coal-fired power plant. Pre-combustion capture
applied with IGCC possesses some thermodynamic advantages over post-combustion capture
owing to the possibility of using physical solvents. This reduces the efficiency penalty from
10-11% for post-combustion capture to 7-8% [29]. Despite several IGCC plants being in
commercial operation in several countries, until now none of them has integrated CO2 capture.
However, there are a number of demonstration projects of IGCC plants with CCS around the
world, including: Texas Clean Energy Project led by Summit Power Inc. - a 400 MW IGCC
plant that captures CO2 for EOR [30]; a 900 MW IGCC plant with pre-combustion capture
currently being developed by Powerfuel Power in the UK [31]; GreenGen Project in China
using technology from Thermal Power Research Institute which seeks to be the first
commercial-scale IGCC (400 MW) integrated with CCS [32]; the ZeroGen project [33], a
400 MW IGCC plant in Australia based on Mitsubishi Heavy Industries technology, planned
to achieve commercial deployment by 2015.
CO2 capture from natural gas reforming generally uses commercially available acid gas
removal (AGR) processes. Most of the separated CO2 is currently vented, though sometimes
it is used to produce urea if the H2 is used to produce ammonia. A few projects are in
operation today including: the Weyburn-Midale Project in the USA and Canada [35], which
is a synfuels production plant with pre-combustion CO2 capture and subsequent utilisation of
the gas for EOR; a fertiliser plant conducted by Enid Fertilizer [36], USA, which also uses
the captured CO2 for EOR. There are also several natural gas processing plants deploying
pre-combustion capture in Norway, USA, and Algeria [29]. Most of these projects use the
CO2 gas for EOR, except two projects in Norway, Sleipner and Snhvit, which store the CO2
in saline formations. In addition to those above, several other projects based on precombustion technologies are still at the execution stage according to a report published by the
Global CCS Institute [12].
Removal of CO2 after coal gasification is a mature commercial process deployed in many
countries [29]. Coal, petroleum coke and heavy oils are gasified with steam or oxygen to
produce feedstock for chemical processes such as the production of ammonia, urea, methanol,
dimethyl ether, SNG, gasoline and other transport fuels via the Fischer Tropsch process. The
CO2 and H2S in the shifted syngas are usually separated from H2 using commercially
available AGR processes [29].

2. Literature Review

31

Adsorption, membrane, and cryogenic processes can also be applied to pre-combustion


capture, separating H2 from CO2 in shifted syngas [20, 29, and 36].
2.1.3 Oxy-combustion capture
Approximately 80% of air by volume is composed of N2, while the remaining is mostly O2.
In conventional combustion processes, fuels are combusted in air generating flue gas
containing N2 diluted CO2, which is then separated using post-combustion capture. Oxycombustion capture separates N2 from air using ASU technologies prior to burning of the
fuels, leaving few requirements for further processing before the compression stage, apart
from condensing the water vapour. In order to exploit the extensive engineering experience
on designing and operating air-fired combustion equipment/process and allow the possibility
of retrofitting, synthetic air made up of recycled CO2 and O2 is often used [37]. A schematic
diagram of the oxy-combustion process is shown in Fig. 2.3.

Figure 2.3 Oxy-combustion CO2 capture [20]


Vattenfalls lignite-fuelled 30 MWth pilot plant at their Schwartze Pumpe power station in
Germany has been in operation since mid-2009 [38]. A capture rate of over 90% has been
demonstrated and this experience offers Vattenfall sufficient confidence to advance the
technology to a 250 MWe oxy-coal demonstration plant. The Lacq project, owned by Total,
is at the same scale as the Schwartze Pumpe plant [39]. However, instead of selling the CO 2

2. Literature Review

32

by-product to industrial markets, the French project stores CO2 in a depleted gas field, 4,500
m below the surface. It is estimated that during the 2-year demonstration, about 120,000
tonnes of CO2 will be captured and stored. In the UK, Doosan Babcock has modified a 40
MWth burner for oxy-combustion at their Renfrew Plant in Scotland [40]. CIUDEN in Spain
is constructing an oxy-coal test facility incorporating a 20 MWth oxy-PC boiler and a 30
MWth oxy-CFB boiler [41]. In Australia, a retired 100 MWth pulverized coal-fired boiler at
the Callide A power station in Queensland has been retrofitted to oxy-combustion [42]. The
captured CO2 will be stored in Permian Denison Trough reservoir. The Futuregen project in
the US, originally proposed to be an IGCC new-built, has been redeveloped as an oxyfuel
boiler retrofitted to an existing coal plant [43]. Apart from these pilot-scale plants, there are
four sub-scale commercial demonstration plants in development according the Global CCS
Institute [37].
2.1.4 Other capture technologies
In addition to the three main categories of capture technologies, several emerging
technologies/processes are attracting significant interests and are developing rapidly, with
some of them entering into pilot scale stage. While far more capture technologies have been
covered in the literature [44], in this chapter we will only focus on three of them: two high
temperature processes (Ca looping and chemical looping) and adsorption with aminefunctionalized materials.
Ca looping applied to post-combustion is often regarded as a hot post-combustion capture
technology, analogous to conventional amine scrubbing. Similar to an amine-based
absorption process, it consists of two vessels: the first one, called the carbonator, is where
CO2 reacts with calcium oxide (CaO) to form calcium carbonate (CaCO3); in the second
reactor, known as the calciner, the sorbernt material (CaO) is regenerated. A pure stream of
CO2 suitable for subsequent compression is also released from the second vessel. Ca looping
is a high temperature process owing to the fact that both reactors are at elevated
temperatures, considerably higher than the conventional amine process: the carbonator is
normally at a temperature higher than 600 C, whereas the release of CO2 occurs in the
calciner at above 900 C. Because of the high temperature nature of the process, the highgrade waste heat from it can be integrated to drive the steam cycle which leads to a lower
efficiency penalty than the conventional amine process [45]. The process also integrates well
with cement production in that the deactivated CaO from the process can be used as a

2. Literature Review

33

feedstock for cement manufacturing. A simplified diagram for the process is shown in Fig.
2.4.
CaCO3
Flue gas
(N2 and CO2)

Carbonator
650-700C
(exothermic)
CaO+CO2CaCO3

CaO

Calciner
900-950C
(endothermic)
CaCO3CaO+CO2

CO2

O2

Fuel

Flue gas,
(mostly N2)

Energy

Figure 2.4 Schematic diagram of Ca looping process


There are several pilot-scale test rigs up to about 120 kWth for Ca looping in Spain, Canada,
China and Germany [46] which demonstrate the successful operation of a dual fluidised bed
Ca looping system under realistic conditions [46]. In addition to these, two larger pilot
projects are coming on stream including: a 1.7 MWth reactor in Spain under an EU
Framework 7 Programme, the Caoling project [45]; and a 1 MWth oxy-firing test facility in
Germany jointly funded by the government and industry. Besides the above projects, Cemex,
the worlds third largest cement producer, is operating a pilot plant in Monterrey, Mexico,
advancing the integration of Ca looping with cement manufacturing [20]. As with many other
capture technologies, the application of Ca looping is not just confined to post-combustion
capture. Although at a later stage of development, sorbent enhanced reforming using
carbonate looping, including the ZEC concept is also attracting significant R&D interests
[47].
Chemical looping combustion (CLC) is proposed as an alternative technology to oxy-fuel
combustion in that oxygen is separated from nitrogen by a reversible reaction of air with
suitable solids, notably reduced metal oxides; these oxidised solids then oxidise the fuels. The
most common configuration (Fig. 2.5) of the CLC process resembles that of Ca looping, i.e.
two interconnected fluidised bed reactors: an air reactor to produce oxidised solids and a fuel
reactor to oxidise the fuels.
The oxidation reaction in the air reactor is exothermic, whereas depending on the type of
metal oxide, the reaction in the fuel reactor can also be exothermic when, for example, copper
oxide, CuO [48] is used. Since the overall heat of reaction in both reactors equals that in the

2. Literature Review

34

case of the direct combustion of fuel with oxygen in air, CLC does not bring any enthalpy
gains. However, it does offer inherent separation of CO2 from N2. Thus, unlike other capture
technologies, where significant parasitic power consumption is required in order to separate
CO2 from flue gas, CLC has the potential to dramatically reduce the energy penalty of CO 2
capture [48]. The technology is still in its early stage of development with Alstom Power
currently developing a prototype facility based on a limestone-derived oxygen carrier [49].
According to the Global CCS Institute, the success from this work would advance CLC to
technology readiness level3 6 (TRL-6) [37]. To further scale up CLC, the availability of a
suitable oxygen carrier is critical. The ideal candidate should possess the properties of: (a)
high oxidation and reduction activity; (b) stability under repeated oxidation/reduction cycles,
including mechanical strength in fluidised beds and resilience to agglomeration; (c) low cost
and environmental impact of the oxygen carrier [48].
MyOx
Air

Fuel Reactor
(often endothermic)

Air Reactor
(exothermic)
MyOx-1 + 0.5O2 (air) MyOx+
(air: N2+unreacted O2)

MyOx-1

CO2 + H2O

(2n+m)MyOx + CnH2m
MyOx-1 + mH2O + nCO2

Fuel

N2, O2

Energy

Figure 2.5 Schematic diagram of CLC process


Apart from the above-mentioned processes, adsorption with solids has also been extensively
studied and wide ranges of materials have been considered, including: carbon materials,
alumina-silicas such as zeolites, alumino-phosphates (AIPOs), alumino-silico-phosphates
(SAPOs), and more recently metal organic frameworks (MOFs) [50]. Unlike those materials
which rely purely on physical interactions, chemical adsorbents (for example, amine
functionalized adsorbents) form chemical bonds with CO2 and selectively remove CO2 from
other gases. In general, amine functionalized adsorbents refer to those prepared through
incorporation of amine groups into solid supports such as mesoporous silica. Compared to
physical adsorbents, these hybrid materials possess several advantages including a higher

Technology Readiness Level (TRL) is used to indicate the development level of the technologies. 9 TRLs were
defined in the Global CCS Institutes report ranging from basic principles observed (TRL -1) to full-scale
commercial deployment (TRL-9). TRL-6 is defined as process development unit (0.1-5 per cent of full-scale).
Refer to the original report [12] for details.

2. Literature Review

35

adsorption selectivity of CO2 and larger adsorption capacity [51]. Meanwhile, they also offer
the potential to solve some of the drawbacks of the amine solvent absorption processes such
as high energy penalty and corrosiveness [50]. Three methods are often used to synthesize the
amine functionalized adsorbents, namely, impregnation [51-56], post-synthetic graft [57-59]
and co-condensation [60]. Sayari et al. [50] has summarized previous studies on the
adsorbents developed for the flue gas treatment of CO2, covering both physical and chemical
adsorbents.
Now the focus of this review will switch to amine-based solvent absorption, beginning from
the categories of amines commonly used in the absorption process.
2.2 Types of amines
The focus of this thesis is on amine scrubbing, particularly vapour-liquid equilibrium of the
system. Vast quantities of VLE data can be found in the literature covering single and mixed
amine systems. Among these amines, MEA (monoethanolamine), DEA (diethanolamine) and
MDEA (methyldiethanoamine) are among the most well-studied solvents. MEA, a primary
amine, has been widely employed in industrial gas treating processes for its notable
advantages, such as: fast reaction kinetics, relatively low cost, ease of reclamation, and low
absorption of hydrocarbons when used in natural gas processing [61]. However, it suffers
from the major drawbacks of a large heat of reaction which leads to a substantial process
energy penalty, as well as susceptibility to degradation and a corrosive nature. As a secondary
amine, DEA is less reactive with CO2 which results in slower reaction rates. On the positive
side, it requires less heat to reverse the reaction and therefore is less parasitic in the stripper
compared to MEA. As to the tertiary amine MDEA, since there is no -hydrogen atom
available for carbamate formation, it is much easier to regenerate CO2 in the stripper (much
less energy is required to reverse the carbonate and bicarbonate formation). Besides, its low
volatility, resistance to degradation and higher cyclic capacity combine to make MDEA an
attractive solvent [62]. However, the inherently slow absorption of CO2 requires a large
absorber to achieve the desired capture rate.
Chakravarty et al. [63] suggested that by blending a tertiary amine (e.g. MDEA) with a
primary or secondary amine, the resulting amine mixture may possess both of their
advantages such as high loading capacity and enhanced absorption rate. Since then, numerous
studies have been conducted on the thermodynamics and kinetics of various amine blends,
including

MEA/MDEA,

DEA/MDEA,

MEA/AMP,

DEA/AMP,

MDEA/AMP,

2. Literature Review

36

MDEA/DEA/AMP, PZ/MDEA, DIPA/MDEA, and DIPA/PZ [71, 142, 146, 168, 185, 203,
208, 224, 234, 236, 237, 239].
Sartori and Savage [64] proposed that a new class of amines, sterically-hindered amines,
should be used for CO2 capture. These are able to approach a capacity of 1 mole of CO2 per
mole of amine while retaining absorption rates comparable to those of secondary amines.
AMP is the most well known of these sterically-hindered amines 4 in the context of CO2
capture processes, owing to the fact that it is the simplest hindered form of MEA, and the
difference in properties can be attributed to the steric hindrance.
Apart from these conventional amines used in gas treating processes, many studies have
explored a much wider range of amines including BHEP [65], AMPD [66], AEEA [67],
DEEA/EEA [68] etc. Appendix I summarises the chemical structures of the common amines
in the context of carbon capture.
This work mainly involves two areas: one is the experimental measurements of the CO2
solubility in aqueous amine solutions; the other concerns the thermodynamic modelling of the
amine-H2O-CO2 systems. In the following sections, previous studies pertinent to this work
will be discussed in detail.
2.3 Experimental apparatus for VLE measurement
Earlier researchers favoured glass stills when measuring vapour-liquid equilibrium because of
its versatility for manufacturing into various shapes. Restricted by the availability of other
techniques such as mixing, circulation, heating and sampling, glass stills provided a
reasonably accurate way of establishing equilibrium and withstanding disturbance while
sampling both phases [69]. Recent studies prefer to use stainless steel (SS) construction,
partly out of safety reasons. For the systems of interest in this work, the VLE measurement
falls into the low to moderate-pressure range; therefore, both glass- and alloy- (stainless steel,
Hastelloy etc.) constructed apparatuses have been employed to measure the VLE for the
amine-H2O-CO2 systems [69], although the latter have been more common in recent years.
Experimental methods can be grouped into three categories: the static-synthetic method, the
static-analytical method and flow systems. The phase compositions can be either calculated
4

Sterically-hinered amine is defined as a chemical compound containing an amine functional group surrounded
by a crowded steric environment and generally refers to a primary amine in which the amino group is attached
to a tertiary carbon atom, or a secondary amine in which the amino group is attached to a secondary or a tertiary
carbon atom.

2. Literature Review

37

through mass balances, as in the synthetic method, or measured through chemical titration or
gas chromatography. For the static 5 type of design, a variety of means are available to
enhance mass-transfer between the phases, including an external rocking design, an internal
magnetic/mechanical stirrer, and one or two circulation pumps. In view of the interest in low
partial pressures of CO2, an inert gas (normally N2 or CH4) is often also introduced in order to
maintain the total pressure at a desirable level. Some of the typical apparatus designs
employed by previous studies are summarised in Table 2.1.
Table 2.1 Typical apparatus and analytical methods employed by previous studies
Authors

Ma'mun et al.
[67]

Silkenbumer
et al. [71]

Li and Chang
[72]

Jou et al. [73]

Types of
measurement

Types of equilibrium
vessel

Static-analytical

three 300 mL
stainless steel
cylinders with
recirculation of the
gas phase

Static-analytical

Static-analytical

Static-analytical

Liquid phase
analysis method

N2

Precipitationtitration (barium
chloride-HCl)
method

None

Gas
chromatography

Reported in total
pressure

Precipitationtitration method

Gas chromatography
to determine the
CO2:N2 ratio if
applicable

N2

N2

Precipitationtitration method
and gas
chromatography

Static-analytical

A magnetically
stirred equilibrium
cell made of
Hastelloy C-276

None

Gas
chromatography

Synthetic

Computer-operated
static apparatus

None

Calculated with
the knowledge of
global coposition

Synthetic

30 mL high-pressure
cell with sapphire
windows

None

Synthetic method

SidiBoumedine et
al. [74]

Kuranov et al.
[75]

950 mL stainless
steel cell with a
propeller stirrer and
gear pump to
circulate liquid phase
1 L stainless steel
vapour-recirculation
cell with another 0.3
L stainless steel
sample cylinder
attached to increase
the vapour phase
volume
200 mL vapourrecirculation cell with
an additional 50 mL
cylindrical reservoir
to increase the vapour
volume

Inert gas
used

Gas phase
composition analysis
Wilson equation to
account for the
solvent vapour
pressure, IR analyser
to determine CO2:N2
ratio

Gas chromatography
to determine the CO2:
N2 ratio
Gas chromatography
to determine
CO2:H2O ratio,
vapour pressure of
amine assumed to be
negligible
Reported in total
pressure, the initial
solvent pressure is
also measured
Reported in total
pressure

Here, static does not describe the situation of the components in the equilibrium vessel. Vigorous mixing
techniques are always used to enhance equilibration of these systems. Compared to the flow-type systems which
have a continuous gas flow, the static apparatuses are more like batch reactors.

2. Literature Review

38

Synthetic with
gas phase
sampling

A sapphire tube
between Hastelloy
flanges with an
internal volume of 34
mL

Bishnoi and
Rochelle [77]

Flow system

A wetted wall
column constructed
from a stainless steel
tube, liquid phase is
circulated

N2

Mondal [78]

Flow system

A bubble column
made of borosilicate
glass

N2

Dicko et al.
[76]

CH4

Calculated with
the knowledge of
global
composition
Acidulation with
phosphoric acid,
then swiped with
N2 carrier gas,
and then
determined with
IR analyzer
Acidulating with
0.6 mol% of HCl
and measuring the
volume of
evolved gas

Gas chromatography

Infrared spectroscopy
analyser

Microprocessorbased CO2 analyser

2.3.1 Static-analytical apparatus


Static-analytical apparatus generally consists of an equilibrium vessel, inlets for feeding in
the gas and liquid components, a mixing mechanism and a sampling system (GC, FTIR or
chemical titration).

The principle is rather generic: a fixed amount of gas and liquid

components are charged into the vessel, thoroughly mixed and thermostated until
equilibration. Afterwards, one or both phases are analysed to determine the equilibrium
concentrations. A variety of means have been utilised to enhance mass-transfer between
phases. In general, as mentioned above, at least three basic designs can be employed: an
external rocking system, an internal magnetic/mechanical stirrer, or one or two circulating
pumps.
a) Rocking system
The rocking design comprises a horizontally positioned cylindrical stainless steel pressure
vessel which is attached to a rocker mechanism. The vessel is generally enclosed in a
constant temperature bath. Some selected literature using a rocking mechanism will be
discussed below [70, 79-85].
Jones et al. [82] measured the solubility of CO2 and H2S in 15 wt% MEA at temperatures
between 40 C and 140 C and loadings from 0.02 to 0.73. A SS vessel of 8.2 L was
constructed with two separate tapped points for gas and liquid sampling. The VLE cell was
placed in a thermostated oil bath while a rocker device was employed to shake the assembly
and enhance phase mixing. During the experiment, 2 litres of preloaded MEA solution was
charged into the cell. It was observed that thermal and chemical equilibria were achieved

2. Literature Review

39

after about one hour. However, the shaking was continued for another hour to ensure the
achievement of equilibration before sampling both phases. The vapour phase composition
was determined using a hybrid method combining both mass spectrometer and calculated
values for the solvent vapour pressure. The liquid phase concentration was analysed with two
procedures: one is through precipitation as carbonate with barium chloride; another is using a
modified Knorr procedure6 [86].
Lawson and Garst [83] employed a similar design as Jones et al. [82] except some alterations
in the gas phase sampling and analysis methodology. To prevent loss of H2S owing to
adsorption and reaction on the SS surface, three types of containers for vapour phase
sampling were used respectively for high H2S concentration, low H2S concentration, and very
low H2S concentration (near or below mass spectrometer detection limit). At the beginning of
the experiment, 600 g of amine solution was added to the pre-evacuated cell followed by the
introduction of a known amount of H2S and/or CO2. In order to maintain the total pressure at
a desirable level, methane was charged into the vessel and the whole system was shaken or at
least 16 hours to reach equilibrium. The vapour phase was routinely analysed by mass
spectrometer and liquid-phase analysis was accomplished by a modified Knorr procedure
[86].
Mamun et al. [70] employed two different methods to measure the solubility of CO2 in MEA
and MDEA. For MEA, an open-type device was used while a shaking cell was utilised for
MDEA experiments. In the shaking cell design, two autoclaves of unequal sizes (1000 and
200 ml) were connected in series and enclosed in an oil bath. The pressure vessels could be
operated up to 2 MPa and 150 C. Initially, the autoclaves were purged several times using
CO2 to expel oxygen. Subsequently, 200 ml of unloaded 50 wt% MDEA was introduced into
the vessel followed by the addition of CO2 until the desired pressure was achieved. The
equilibrium was achieved in 4 to 30 hours. A titration method using barium chloride (BaCO3)
was employed for the liquid phase analysis whereas no direct measurement for vapour phase
was conducted. Instead, the partial pressure of CO2 was deduced from differences between
system total pressure and the initial vapour pressure of MDEA solutions at the experimental
conditions.b) Internally-Stirred system

Modified Knorr procedure involves taking an aliquot of the sample which is then acidified with sulphuric acid
and heated. The released carbon dioxide is swept from the reaction flask with nitrogen and passed in series
through a drying tube and a tared absorption tube containing a carbon dioxide absorbent, Caroxite. When the gas
stream contains hydrogen sulphide, acidified potassium permanganate is used to remove it.

2. Literature Review

40

The second method of mixing is normally achieved in a SS autoclave agitated with a


magnetic or mechanical stirrer while a bath or heater jacket is used for temperature control
[65, 66, 87-99]. To accommodate the stirring bar, the vessel is usually upright, in contrast
with the above-mentioned design, although a horizontal VLE setup for measuring the binary
dimethyl ether + propane/propene systems has been described by Horstmann et al. [100]. The
designs using a stirrer bar as the mixing mechanism are detailed as follows.
Chakma and Meisen [65] used a stirred type apparatus modified from previous research [91].
Solubility of CO2 in MDEA particularly at the high temperature end (100-200C) and in
BHEP (N, N-bis(hydroxyethyl)piperazine) between 40 to 180 C was measured. At the
beginning of the experiment, a 600 ml SS autoclave was charged with unloaded MDEA or
BHEP solution and heated to a desirable temperature before being attached to a SS gas bomb
filled with pure CO2. Equilibrium was usually achieved within half an hour but at least 4
hours was allowed for the system to equilibrate. The total amount of CO2 introduced into the
vessel was calculated from the weight variation of the gas bomb before and after the
connection with the equilibrium cell. For the vapour phase, the partial pressure of CO2 was
determined by subtracting the vapour pressure of the solvents at the experimental temperature
(before introducing CO2) from the total pressure after the equilibration. The amount of CO2
in the liquid phase was calculated from a mass balance in conjunction with applying the BW-R equation of state to the gas phase CO2.
Park et al. [95] designed equipment consisting of a magnetically-stirred equilibrium cell and
a loading cylinder for the introduction of CO2. Both of the two containers had an internal
volume of about 500 ml. An electrically-heated water bath was used to control the
temperature of the equilibrium cell at a constant value. During the experiment, about 100 ml
of the amine solution was added to the equilibrium cell followed by purging with N2 in order
to remove trace amounts of air. The addition of N2 also served to maintain the system
pressure above 1 atm. CO2 was introduced into the cell when switching on the gas inlet valve
between the loading cylinder and the cell. The equilibrium was assumed to have been reached
when the pressure of the equilibrium cell did not change for 5 h. By measuring the pressure
change of the loading cylinder, the total amount of CO2 added to the cell was calculated using
a virial equation of state. The vapour phase composition was determined by gas
chromatography (GC) while the liquid phase loading of CO2 was calculated from a mass
balance.

2. Literature Review

41

Kundu et al. [92] constructed a stirred-type reactor to measure the VLE of CO2 in AMP
solution over a temperature range of 30 to 50 C and loading from 0.4 to 1.0. In the glass cell,
a magnetic stirrer was used to enhance the liquid phase mixing while two impellers mounted
on a shaft over the top of the cell were used for gas phase homogenisation. After the cell
reaching a desired temperature, a mixture of CO2 and N2 saturated with water vapour at the
same temperature was purged through the system. Following this, 10 ml of amine solution
was transferred into the cell. Pure N2 was employed as the makeup gas to maintain the total
pressure near atmospheric. After establishment of equilibration, the gas phase was analysed
using GC and the CO2 loading was measured through titration with 6 M HCl solution.
Huttenhuis et al. [89] used a 1 litre Bchi reactor intensively stirred by an impeller and
measured the solubility of CO2 and H2S at relatively low temperatures (10 C and 25 C).
Initially, the reactor was half-filled with solvent and degassed in situ under vacuum. Then
CO2 and H2S were added from a gas bomb and CH4 was introduced to raise the system
pressure to a desired level. During the experiment, CH4 was also employed to sweep the gas
mixture in the reactor towards the outlet where an infrared analyzer was used for determining
CO2 and a GC was used for analysing H2S. Equilibrium was achieved when the
concentrations of these two gases stabilised (normally after 1 hour). Afterwards, liquid
samples were withdrawn and analysed through acid-gas titration.
c) Circulating system
The circulating system is one of the most common designs for equilibrium measurement [71,
101-105]. An advantage of this configuration is that the circulated phase can be passed
through the sampling apparatus (e.g. GC sampling loop) eliminating the dead volume in the
equilibrium system. The equilibrium vessel is usually made of SS, and placed inside a
constant temperature bath. Most apparatuses circulate only the vapour phase by a
magnetically driven pump; the vapour is bubbled through the liquid to facilitate mixing [72,
73, 98, 107-109]. However, there are designs which circulate the liquid phase or even both
phases [71, 110-112]. In some designs, a magnetic propeller is used together with the
circulation pump to further enhance mass transfer between phases [71, 113]. Some typical
circulating apparatus will be discussed here.
Jou et al. [73] modified a gas-circulation type of apparatus based on previous experiments.
The equilibrium cell was a 200 ml Jerguson liquid level gauge with a 50 ml vapour reservoir
on top of it. The vessel was placed in an air bath with an operating range of -25 to 160 C. To

2. Literature Review

42

enhance mixing, a magnetic pump developed from the design by Ruska et al. [114] was
utilised to circulate the gas and bubble it through the liquid. The experiment was started by
introducing approximately 100 ml of amine solution into the pre-evacuated cell. N2 or CO2
was purged through the system to expel O2 before more CO2 was introduced until a desired
pressure was reached. N2 was also used as makeup gas to maintain the total pressure above
atmospheric if necessary. After attainment of equilibrium, the liquid phase was analysed
through a precipitation-titration method, whereas gas samples were determined using GC.
The concentrations of liquid samples were also periodically checked with the GC.
Silkenbumer et al. [71] constructed a SS vessel of approximately 950 ml and measured the
VLE of aqueous AMP. A water bath was used to maintain the temperature at 40 to 80 C.
The cell was a combination of the stirred and circulation types, with a propeller stirrer
mounted on the top and circulation of the liquid phase using a gear pump. The circulation of
the liquid phase via the gear pump not only improved mixing, but also allowed sampling
through a liquid sampling valve and subsequent online analysis with GC. After introducing
the liquid and gas components, equilibrium was established after around 2 hours when
temperature and pressure readings both stabilised. The gas compositions were also
determined using the GC and were taken alternately with the liquid for analysis.
Rogers et al. [112] devised an innovative way of measuring both phases via Fourier
transform infrared (FTIR) spectrometer without the removal of any samples from the
equilibrium system (Fig. 2.6). The system was composed of a SS equilibrium cell, two
vapour cells and one liquid cell. Mixing was achieved by circulating liquid and vapour phases
around the apparatus via the liquid and gas FTIR cells and returned to the opposite phases.
Two pumps were deployed to circulate the phases: a metal bellows pump with SS bellows
and check valves for the vapour phase and a positive-displacement pump to circulate the
liquid sample. In order to prevent condensation of the vapour, the gas circulation line was
heated well above the equilibrium temperature whereas the liquid line was cooled to suppress
formation of bubbles. By coupling an FTIR to the apparatus using mirrors, the concentrations
of the components in both phases were determined in situ. This method was especially
advantageous when trace amounts of H2S was present, since this enclosed system resulted in
no loss of H2S owing to adsorption to the container surfaces etc.

2. Literature Review

43

Figure 2.6 FTIR coupled VLE apparatus [112]

2.3.2 Static-synthetic apparatus


Static-synthetic designs share some similarities with the static-analytical setups. They also
consist of an equilibrium vessel, temperature control system, and temperature and pressure
sensors. The most notable difference is that unlike the static-analytical counterpart, it does
not have a dedicated section for the direct analysis of phase compositions; instead phase
distribution is derived from the global compositions. There are fewer previous studies
carried out with synthetic apparatus probably owing to the difficulties in the equipment
design, including the stringent tolerance on dead volume. Previous studies using staticsynthetic design are discussed in detail as below [74, 75, 76, 116].
Sidi-Boumedine et al. [74] employed two synthetic apparatuses to measure the CO2 solubility
in MDEA, DEA and their mixtures at temperatures between 298 and 348 K. The first one was
computer-controlled and could be operated at temperature between 270 and 400 K, and
pressures up to 5 MPa. During the experiment the purified and degassed solvent were
expelled into a stirred equilibrium vessel using piston injectors. This gave a precise
knowledge of the injected liquid volume. Subsequently, the gas was introduced into the cell
through a gas bomb in a stepwise manner. The synthetic method does not measure the phase
composition directly, instead it requires knowledge of the global compositions, so that the

2. Literature Review

44

amounts of pure components added to the system need to be known precisely. For CO 2, the
information was derived by measuring the pressure difference of the gas bomb together with
an accountable equation of state. With regards to the solvent, its density was separately
determined as a function of temperature using a vibrating tube densimeter. In the second
apparatus, which could be manually operated at temperatures between 200 and 500 K and
pressures up to 15 MPa, the experimental procedures were similar to the first one, except that
CO2 was introduced as liquefied gas using the same piston injector pump as the addition of
the solvent. To derive the phase compositions of each component from the global amount
introduced into the system, the authors devised an iterative isothermal and isochoric
algorithm taking into account the effects of solvent vapour pressure, compressibility of the
solvent, the partial molar volume of the dissolved gas and the solvent activity coefficient.
Bougie and Iliuta [116] used an apparatus based on the static-synthetic method to produce
solubility data of CO2 in aqueous piperazine (PZ) solutions over a temperature range from
= (287 to 313) K and for amine concentrations from

= (0.1 to 2.0) mol.kg-1. The system

consisted of a titanium vessel agitated with a magnetic rod. Liquid was introduced into the
equilibrium cell with a variable volume press equipped with a linear encoder to allow the
precise knowledge of the longitudinal position of the piston. Gas was loaded inside the vessel
through a thermostated gas cylinder equipped with accurate pressure and temperature
measurements. During the experiment, liquid was introduced into the pre-evacuated vessel
followed by filling the gas. The equilibrium was assumed to be reached when the pressure
inside the cell varied by less than 0.5% for at least 30 min. The liquid phase loading was
derived from a mass balance where the difference between the introduced and residual gas in
the head space of the equilibrium cell was assumed to be the amount of gas dissolved into the
solution.
An approach based on bubble point determination was described by Kuranov et al. [75]. A 30
ml, high-pressure vessel with sapphire windows was employed to measure the solubility of
CO2 and H2S in aqueous MDEA solution between 313 and 413 K. The system also included a
high-pressure pump to introduce solvent. Initially, the equilibrium cell was filled with a
known amount of gas. Then aqueous solvent was added to the system until the gas was
completely dissolved. When the system reached equilibrium, very small amounts of the liquid
was withdrawn to reduce the system pressure until the first stable bubble appeared. The
solubility of CO2 was derived from the amounts of the gas and liquid introduced into the
system.

2. Literature Review

45

2.3.3 Flow systems


In a flow type system, two or more vessels are connected in series and immersed in a constant
temperature bath. The equipment is first charged with a stationary phase followed by purging
gas through the liquid. The establishment of equilibrium is usually indicated by comparing
the composition of the inlet and outlet gas streams. There are two ways of operating a flow
type apparatus: one is called the liquid saturation method and the other is the gas saturation
procedure. These two complementary methods simulate respectively the absorption and
desorption processes in the scrubbing process. In the liquid saturation method, unloaded
amine solution is first introduced into the system vessels. Subsequently, a continuous,
premixed gas stream, normally consisted of N2 and CO2, flows through the system and
redistributes itself between the phases. In contrast, the gas saturation method involves
charging a pre-loaded amine solution into the vessel and then stripping out the acid gas using
water-saturated inert gas, often N2. Numerous previous studies are based on the flow type
design [67, 78, 85, 117-124]. Some of these will be covered in detail in the following
paragraphs.
Issacs et al. [118] modified a flow type apparatus and used the gas saturation method to
measure the solubility of H2S and CO2 in 2.5 M MEA solution at 100 C. The apparatus
consisted of three SS cylinders, each with 500 ml capacity, connected in series. The whole
system was immersed in an oil bath controlled to within 0.5 C. Preloaded MEA solution
was charged to the three bottles followed by purging N2 through the liquid to strip the CO2 or
H2S out of the solution. When equilibrium was reached, gas samples were sent to analysis by
gas chromatography whilst the liquid phase was precipitated by BaCl2 before titration with
HCl.
Mamun et al. [67] designed two separate flow-type systems for measurements under
atmospheric or medium pressures. The low-pressure setup consisted of four glass flasks
heated in a water bath and could be operated up to 80 C. The first flask was used as a gas
stabilizer and the other three were charged with 150 ml of preloaded amine solution. Purging
gas was circulated around the system until the temperature and CO2 composition stabilized.
The medium-pressure apparatus (Fig. 2.7) included three SS cylinders and could be operated
at pressures up to 700 kPa and temperatures up to 130 C heated using an oil bath. The
procedure was generally similar to the low-pressure one. Once equilibrium was attained
(usually 2-3 hours), the liquid sample was discharged from the last flask into a 75-ml

2. Literature Review

46

evacuated sampling cylinder and analysed using the precipitation-titration method (BaCl2HCl). The CO2 concentration in the gas phase was determined with an online IR analyser.

Figure 2.7 Equilibrium vessels for medium-pressure measurements [67]


Mondal [78] used an unconventional saturation method to measure the solubility of CO2 in
DEA and piperazine blends under ambient pressure. During the experiment, a 500 ml, glass
bubble column was loaded with the amine mixture and immersed in a water bath. Then a
mixture of N2 and CO2 gas stream was slowly bubbled through the solution at a minimal gas
flow rate. The CO2 concentrations at the inlet and outlet of the apparatus were analysed by a
microprocessor-based analyser at 10 min intervals. Equilibrium was established when the
outlet reading equalled the inlet. Subsequently, the liquid sample was analysed with the
acidulation method with the addition of HCl to the solution and measuring the volume of gas
evolved.
2.3.4 Other methods
Although the majority of the experimental methods can be grouped into the three categories
discussed above, some exceptions including using a wetted wall column or the ebulliometric
method can be found in the literature. The wetted wall column is a simplified version of the
packed tower column where liquid and gas are brought into counter current flowing contact.
This apparatus can be used for measuring both the kinetic and thermodynamic properties of a
compound or mixture. The ebulliometric method basically utilises a type of specially built

2. Literature Review

47

glassware in which gas and liquid are brought to equilibrium in a dynamic manor. Once
equilibrium is attained, the liquid and gas samples can be drawn from the ebulliometer
through prebuilt ports and analysed in routine ways.
Bishnoi and Rochelle [77] constructed a wetted wall column and measured both the solubility
and kinetics of CO2 dissolution in aqueous piperazine solution. Amine solution was
circulated by a Cole-Parmer micropump. It entered the column via an inlet at the bottom
centre and climbed up in a single stream to the top of the column where it was separated into
two streams by the column walls. Then the two streams of solution travelled back to the
bottom along the wall where they were in contact with a stream of N 2/CO2 gas mixture.
Solubility information was derived by bracketing the equilibrium point in an absorption and
desorption cycle. The gas concentration was measured with a series of IR CO2 analysers
while the liquid loading was determined through titration with phosphoric acid and then
analysing the total amount of CO2 released using the IR analyser.
Kim et al. [125] conducted a series of VLE measurements for single and mixed aqueous
amines without acid gas in a modified Swietoslawski ebulliometer [69]. The glass
equilibrium still had an internal volume of 200 ml and could be operated at a maximum
temperature of 200 C and pressure of 1 bar. After purging the system with N2, approximately
80 ml of solution was charged into the glassware. The liquid was heated by an electric heater
and partially evaporated. Equilibrium was considered to be established when the temperature
and pressure stabilised for a minimum of 10 min. Next, samples of liquid and vapour were
withdrawn from the apparatus for further analysis; a standard titration procedure with 0.1 M
H2SO4 was used for liquid samples containing single amines while GC was employed for
mixed amine solutions. The vapour samples were analysed with both titration and GC
methods.
2.4 Thermodynamic models
Experimental data are important in terms of offering a direct and intuitive way of knowing
the solubility of the CO2 in amine systems. However, under extreme conditions such as very
lean acid gas loading, high system temperature, and very high amine concentrations, it is
sometimes difficult to obtain accurate data owing to analytical constraints. Thermodynamic
models complement the experimental data in that the former usually requires the latter to feed
in the parameter regression; reversely, the thermodynamic models can be used to predict the
system behaviour beyond the experimental measurements. Besides, the quantitative

2. Literature Review

48

understanding of the chemical and phase equilibria in the dissolution process is vital for the
design, modelling and operation of an amine scrubbing plant in real life.
2.4.1 Solution chemistry
The chemical absorption process of an acid gas (e.g. CO2) by the alkaline amine solution
involves a series of reactions shown below.
H2O

H+ + OH-

CO2 + H2O

HCO3-

RR'NH + H+

H+ + HCO3-

H+ + CO32-

RR'NH2+

IV
V
VI
VII

For primary and secondary amines, CO2 can directly react with the amines and form stable
carbamates:

RR'NCOO- + H2O

RR'NH + HCO3-

VIII

As pointed out by Austgen et al. [130], other side reactions may take place along with the
ones listed above. For instance, MEA and DEA are particularly vulnerable to degradation and
when exposed to stripping conditions (> 100 C) for a prolonged period, heterocyclic
compounds could form between amine and CO2. However, the author also noted that these
side reactions are relatively minor on a laboratory scale (especially with much less cycling
time compared to the industrial environment) and are generally neglected in studies of the
vapour-liquid equilibrium for the amine systems.
2.4.2 Thermodynamic framework
The thermodynamic framework underlying the vapour-liquid equilibrium for the amine-H2OCO2 system should cover both chemical and phase equilibria. Chemical equilibria are
described by the reactions I-V and can be summarised with the mass action law:
(2.1)

2. Literature Review
Here,

49

is the temperature-dependent equilibrium constant;

is the stoichiometric constant of species ;

is the activity of species ;

is the molality of species i;

is the activity

coefficient of species i.
Vapour phase equilibria are only relevant to the molecular species as ions are treated as nonvolatile. Under certain circumstances, high-boiling point amines (e.g. MDEA) are also
considered as non-volatile for simplicity. An extended Henrys law is usually employed to
represent the behaviour of solutes such as CO2:
(2.2)
Meanwhile for solvents such as H2O, an extended Raoults law is used, given by:
(2.3)
Here, the exponential terms in the above equations are the Poynting factors or corrections for
moderate pressure and are obtained from integration forms by assuming
the pressure range.

constant over

are the mole fractions of species and in the gas phase;

vapour phase fugacity coefficients for species and ;


mole fractions of species

and

coefficients for species and ;

in the liquid phase;


,

are

are the

are the liquid phase activity

are respectively the Henrys constant of

amine solution and saturation pressure of component ;


under saturation condition;

is the total pressure;

in the

is the fugacity coefficient of

are the partial molar volume of solute at infinite dilution

and the molar volume of the solvent ;

is the vapour pressure of solvent

at saturated

conditions.
Apart from the equilibria equations, the system is also subject to mass balance constraints
including:
Amine material balance:
(2.4)
Carbon material balance:
(2.5)

2. Literature Review
is the initial amine molality;
liquid phase;

50
is the total molality of CO2 or CO2 equivalent in the

is the liquid phase loading of CO2.

Charge balance:
(2.6)
The equilibrium constants and the Henrys constants are temperature dependent. Most of
these equilibrium constants can be obtained from the literature; however, carbamate stability
constants are scarcely reported and only limited types of amine and temperature conditions
are available from the literature. Jensen et al. [126] used the Bjerrum expression to calculate
the carbamate equilibrium constant but this was only viable at 18 C. Aroua et al. [127]
designed an experiment to measure the carbamate stability of MEA at temperatures of 25, 35,
45, and 55 C and ionic strength up to 1.7 M. A generalised form of the dependence of the
equilibrium constants on temperature was then derived. Although some modellers simply
extended Jensens result to the desired temperature, this was often believed to be a source of
error. More common practice was to simply regress this constant together with the interaction
parameters from measurements of the thermodynamic properties. This could be achieved by
either single-step or two-step procedures. Details of the regression methods will be discussed
in the following section.
Henrys constant of CO2 in water is a well-studied parameter, but much less known in the
case of aqueous amine solutions. The reactivity of CO2 in amine solutions has rendered direct
determination impossible. Therefore, Browning and Weiland [128] proposed the N2O
analogue method to measure the physical solubility of CO2 in amine solutions. N2O was
chosen to be the analogy of CO2 because: firstly, it does not chemically react with the amines;
secondly, it has a similar molecular weight as CO2, which is believed to result in similar
diffusivity in the liquid phase.
2.4.3 Important VLE model
Many researchers and industrial practitioners have attempted to develop predictive models,
built upon a large body of VLE data [117, 129-142]. The early attempt was simply a curvefitting method [122] and the gradually improved to rigorous thermodynamic models. The
major distinction between rigorous and non-rigorous models lies in the way of treating the
phase non-ideality. Generally speaking, the activity-fugacity method, or - approach, is the
most commonly used formulation of rigorous VLE models [139]. In this approach, different

2. Literature Review

51

methods can be selected for the liquid and gas phases to estimate activity and fugacity
coefficients. Some of the most important models are discussed in the following paragraphs.
a) Kent-Eisenberg Model
Kent and Eisenberg [137] built upon the work of Dankwerts and McNeil [132] and developed
a simple correlation method which included all the nonideality in the liquid phase into certain
selected equilibrium constants. Although not a predictive model, it is still frequently used
especially in industrial environments. Similar to the previous model, in the Kent-Eisenberg
model, vapour pressures of the molecular species were proportional to the free component
concentrations in the liquid phase governed by the Raoults law. Activity coefficients of all
the components were set to unity. Selected equilibrium constants, in this case, amine
protonation constant and carbamate formation constant were adjusted to accommodate the
phase nonideality. This greatly simplified the model as only two parameters were required for
each amine. Although only mixed CO2 and H2S gases in MEA or DEA solutions were
discussed in the original paper, this model can be easily extended to other amines and amine
mixtures.
The Kent-Eisenberg model is especially popular among industrial practitioners for its
simplicity and reasonable accuracy. Academic researchers also used this model to correlate
their experimental data with modified versions: Kritpiphat and Tontiwachwuthikul [143] also
fitted their solubility data of CO2 in AMP with this model. Park et al. [150] measured CO2
solubility in MEA, DEA, AMP, MEA + AMP and DEA + AMP and regressed their data to
obtain parameters of amine protonation and carbamate formation equilibrium constants in the
Kent-Eisenberg model. The correlation was then compared with experimental data obtained
by Tontiwachwuthikul et al. [144] and another prediction from Xu et al. [145] and proved to
be satisfactory.
The original Kent-Eisenberg model assumed that the equilibrium constants for reactions
involving the amines were only dependent on temperature. Haji-Sulaiman et al. [146]
modified the original Kent-Eisenberg model to incorporate the dependency on the free gas
concentration in solution and the amine concentration in the expressions of equilibrium
constants of amine protonation and carbamate formation. In addition, they extended the
model to applications where no cabamate was formed (i.e. MDEA) as well as mixed-amine
solutions (i.e. DEA-MDEA). Jou et al. [119] also assumed the equilibrium constants to be
dependent upon the CO2 loading and the amine concentration, in addition to the temperature.

2. Literature Review

52

Hu and Chakma [147] introduced a modified expression for the equilibrium constants for the
amine reactions as functions of temperature, acid gas partial pressure and amine
concentration. Li and Shen [148] also correlated CO2 solubility in MEA and MDEA mixture
using a modified version of the Kent-Eisenberg model assuming the amine related
equilibrium constants as functions of temperature, amine concentration and CO2 loading.
This method has been employed by Yang et al. [149] to correlate the CO2 solubility in
aqueous AMP and PZ solutions.
Despite its wide application in industry, there are some drawbacks of this correlation type of
model. Although a qualitative trend could still be useful [95], the simplified treatment of
phase non-ideality reduces its credibility in the regions beyond the experimental data.
b) Model of Edwards et al.
Edwards et al. [151] devised a model for aqueous solutions of ammonia, carbon dioxide,
hydrogen sulphide, sulphur dioxide, and hydrogen cyanide containing one or more
electrolytes. Ternary systems of ammonia-CO2-H2O and ammonia-H2S-H2O were
particularly addressed. Activity was used rather than concentration and related to molality
through activity coefficients. Vapour phase deviation was accounted through fugacity
coefficients using the method suggested by Nakamura et al. [152], while liquid phase
nonideality was taken into account using an expression derived from Pitzers theory. In
Pitzers equation, which included the long-range Debye-Hckel contribution and a shortrange term in the form of the Margules expression [153], activity coefficients were related to
molality and electronic charge of particular components, ionic stength of the solution, and
interaction parameters between the ions. Since like ions (i.e. the ions with the same sign of
charge) do not approach each other closely, their interaction parameters were set to be zero.
The molecular-molecular, molecular-ion, and unlike ion-ion interaction parameters were
temperature-dependent and were estimated from Bromley [154], Edwards et al. [155], and
Pitzer and Mayorga [156], respectively. The equilibrium constants and Henrys constants
were also assumed to be functions of temperature.
The model developed by Edwards et al. was the first thermodynamically rigorous approach
for correlating experimental data and has been employed by many authors. Arcis et al. [157]
used the equation for the activity coefficient from Edwards et al. to account for the liquid
phase nonideality. For the vapour phase, unlike in the work of Edwards et al. which used the
method of Nakamura et al. [152], the truncated virial equation was used. Many studies [130,

2. Literature Review

53

131, 158] based their equilibrium constants, including those of water ionisation, carbonate
and bicarbonate formation, on Edwards et al.s compilation. Some authors used the original
Pitzer [155] equation rather than the modified version from Edwards et al. For instance,
Kuranov et al. [75] calculated the activity coefficients based on the principal proposed by
Pitzer and regressed their experimental data on CO2 and H2S in aqueous MDEA. An
improvement of this work was that the influence of temperature on binary interaction
parameters was incorporated. In more recent work from Bougie and Iliuta [116], solubility
data of CO2 in aqueous PZ solutions were measured over a temperature range of 287 to 313
K in a static-synthetic apparatus. These experimental data together with others from the
literature for the ternary system PZ-H2O-CO2 were correlated using a model incorporating a
modified Pitzers thermodynamic model for activity coefficient with the virial equation of
state for the fugacity coefficient.
The major limitation of Edwards et al.s model, as pointed out in their paper, is the maximum
ionic strength-constrained by the expression of the activity coefficient. Another constraint is
the temperature limit owing to the validity range of the equilibrium and Henrys constants.
However, the upper limit of 170 C means that most industrial processes can be described
using this model.
c) Deshmukh-Mather method
Deshmukh and Mather [133] developed the first thermodynamically rigorous model
specifically for aqueous alkanolamine systems. Similar to its predecessors, the
thermodynamic framework included both chemical and phase equilibria as described by the
following equations:
(2.7)
(2.8)
For electrolyte species, activity coefficients approach unity in an infinitely dilute solution
whereas for the solvent, a pure compound leads to a value of unity for the activity coefficient.
To calculate the activity coefficient, an extended Debye-Hckel expression was selected
which took into account electrostatic forces and short-range Van der Waals force.
(2.9)

2. Literature Review
The binary interaction parameters,

54
, describe the influence from each pairs of species. The

values of these parameters were normally obtained from regression of relevant experimental
data. The distinctiveness of the Deshmukh-Mather model was its mathematical framework to
solve the large set of non-linear equations. As described in their work, Browns method [159]
based on a partial pivoting technique was employed to solve a system of non-linear algebraic
equations in a FORTRAN programme. For the vapour phase, the Peng-Robinson [160]
equation of state was used to calculate the fugacity coefficients.
This model has attracted a considerable amount of attention in the thermodynamic modelling
of CO2 in aqueous amine solutions. Chakravarty [161] extended the model to systems of
mixed amines. Weiland et al. [162] generalised the model and predicted CO2 and H2S
solubility in MEA, DEA, DGA and MDEA. Jou et al. [73] correlated their measurements of
CO2 in 30 wt% MEA with this model and compared with the data from Lee et al. [107]. HajiSulaiman and Aruoa [88] also utilised this model to simulate the VLE of CO2 in DEA and
AMP. Benamor and Aroua [131] modelled solubility of CO2 in DEA, MDEA and their
mixtures using the Deshmukh-Mather model. Unlike the original case, binary interaction
parameters as well as the equilibrium constant of carbamate formation were regressed
simultaneously from the experimental data. The regression was based on their previous
laboratory measurements [146], and in order to validate their model, the calculated results
were compared with the experimental data from other studies [119]. Mamun et al. [67] used
the Deshmukh-Mather model to correlate the experimental results obtained in the same work
and predict speciation in the loaded solution, which was compared to NMR analysis.
Pahlavanzadeh et al. [163] measured the solubility of CO2 in aqueous solutions of AMP at
temperatures from 293 K to 323 K, for concentrations of AMP from 1 M to 4 M and analysed
the experimental data with two different mathematical models including Deshmukh-Mather
and an artificial neural network.
The most notable advantage of the Deshmukh-Mather model is its computational efficiency
[67]. Compared to more complicated ones, such as the eNRTL model, this characteristic
makes it easier to be incorporated into a process model where the evaluation of equilibria
may be performed 105 to 106 times [67]; meanwhile, it is more rigorous than the KentEisenberg model and capable of generating credible speciation predictions.

2. Literature Review

55

d) Electrolyte-NRTL (e-NRTL) Model


The electrolyte non-random two-liquid (e-NRTL) model was first proposed by Chen et al.
[164] who derived it from the local composition concept of the non-random two-liquid
hypothesis [165]. Its original form was only applicable to single solvent and completely
dissociated electrolyte systems. Chen and Evans [166] generalised the model to aqueous
multicomponent electrolyte systems. In the e-NRTL model, the expression of excess Gibbs
energy was comprised of three parts: the unsymmetrical Pitzer-Debye-Hckel model (

was used to represent the long-range interactions and the symmetric non-random two-liquid
theory (
term (

) was employed to account for the short-range interactions. Finally, the Born
) is a correction of reference state for the ions.
(2.10)

A desirable feature of this model is that even in multicomponent systems, only binary
interaction parameters are required, i.e. higher-order parameters are unnecessary.
Theoretically, all the species pair interaction parameters can be obtained from binary solventsalt systems and used to predict the phase behaviour of multicomponent systems.
Austgen et al. [130] devised a model for single acid gases in MEA or DEA solutions using
the e-NRTL equation to calculate activity coefficients in the liquid phase. The interaction
parameters, assumed to be temperature-dependent, were regressed from various binary and
ternary VLE data collected from 20 different papers. Because of the lack of experimental
measurements of the carbamate stability constant, it was lumped into the fitting process with
the interaction parameters. The vapour phase fugacity coefficients were represented using the
Redlich-Kwong equation of state as modified by Soave (SRK model) [167].
Bishnoi [77] studied the kinetics and thermodynamics of CO2 in aqueous blends of MDEA
and piperazine (PZ). The e-NRTL model was used to represent the activity coefficients. Gas
phase non-ideality was taken into account using the SRK equation of state. The model has
been implemented in ASPEN PLUS software.
Liu et al. [140] further explored the e-NRTL model for the CO2-MEA-H2O system in an
attempt to improve Austgen et al.s work, which over-predicted the reboiler heat duty.
Several amendments were made to the original model. First of all, instead of using the
Henrys constant in water for the case of MEA, two indirect methods, namely the CO2-N2O
analogy method and the plant-data-optimisation method were employed to calculate the

2. Literature Review

56

physical solubility of CO2 in reactive amine solutions. Secondly, a two-stage procedure


proposed by Weiland et al. [162] was adopted to reduce the error from simultaneous fit of
carbamate stability constant and binary interaction parameters. Finally, some less-credible
experimental values were rejected based on a quality analysis rather than treating all the
gathered data equally, as in Austgens work. The improved model was then validated against
three industrial cases and proved to be more reliable than the original model.
Hilliard [168] measured and modelled the CO2 solubility in aqueous blends of MEA, PZ, and
potassium carbonate. In his thermodynamic model, liquid phase non-ideality was accounted
for using the e-NRTL model while the vapour phase was modelled using the Soave-RedlichKwong (SRK) equation of state.
Posey [169] constructed a model for the aqueous alkanolamine-acid gas systems based on the
e-NRTL model. Rather than just relying on VLE data, a wide range of sources covering
different thermodynamic properties including conductivity, pH, enthalpy of absorption, and
freezing point data were utilised to regress the binary interaction parameters. It was
recommended that for low loading predictions, data sources should come from other types of
measurements rather than VLE data since the latter were subject to severe errors caused by
pushing the concentration analysis to its detection limits.
Schmidt et al. [170] gathered a large amount of data for aqueous amine systems of MEA,
DEA and MDEA including VLE, activity coefficients, and excess enthalpy, and checked the
credibility of these data. They subsequently developed an algorithm to determine the binary
interaction parameters in the e-NRTL model by simultaneous regression of the three types of
data according to the objective function in eq. 2.11.

(2.11)
where , , and

are the total number of experimentally determined, activity coefficients,

VLE measurements and excess enthalpies. Comparison was made between their model and
previous ones. It was believed that the minor discrepancy was caused by the quality of data
fed into the regression procedures.

2. Literature Review

57

Bollas et al. [171] pointed out that the original e-NRTL model is inconsistent for systems
with multiple ions, and subsequently developed a refined version in which a simplifying
assumption, that ionic-charge fraction quantities remain constant, in the derivation of activity
coefficient expression was removed. Hessen et al. [172] applied the refined e-NRTL model
of Bollas et al. to alkanolamine-H2O-CO2 systems. Binary, ternary vapour-liquid equilibrium
data, freezing point depression data and excess enthalpy data were used to regress the binary
interaction parameters. Zong and Chen [173] developed a thermodynamic model for the
solubility of acid gases, CO2 and H2S, in aqueous DIPA solution, aqueous sulfolane-DIPA
solution, and aqueous sulfolane-MDEA solution. The model employed the 2009 version of
the e-NRTL model for the liquid phase activity coefficient and the PC-SAFT equation [174]
for vapour phase fugacity coefficient calculations. Zhang et al. [175] used the e-NRTL
activity coefficient model in Aspen Plus to formulate a rigorous and thermodynamically
consistent representation of the MEA-H2O-CO2 system. Interaction parameters were obtained
in two steps: for the MEA-H2O binary, VLE, heat capacity, and excess enthalpy data were
used to obtain the binary interaction parameters; while for the MEA-H2O-CO2 ternary system,
VLE, enthalpy of absorption, heat capacity and NMR spectroscopic data were employed to
regress the binary interaction parameters between MEA-CO2 and H2O-CO2 as well as those
standard state properties of amine ions, MEA protonate and carbamate.
e) SAFT-VR model
The statistical associating fluid theory for potentials of variable range (SAFT-VR) was
extended from the original SAFT model to include attractive potentials of variable range
[176]. The latter was originally derived from Wertheims perturbation theory [177-180] for
association and is a powerful tool for modelling properties of substances where association
phenomena and molecular shape are the dominant factors [181]. The excess Helmholtz
energy based model incorporates three terms, corresponding to interactions that contribute to
the total excess property.
(2.12)
where

is the ideal Helmholtz energy,

monomer segments,
and

is the residual energy caused by

is contribution attributed to the formation of chains of monomers,

describes the contribution from intermolecular association.

chain molecules in the mixture, k is the Boltzmann constant, and

is the number of

is the temperature.

2. Literature Review

58

Unlike the - models, SAFT-VR model gives equation of state formulations for both liquid
and vapour phases. This model has been successfully implemented to describe various simple,
non-associating systems such as alkanes and those comprising replacement refrigerants [182].
It has also been employed by Keskes [183] to predict the VLE for CO2-CH4-n-alkane.
Recently, the model has been applied to model the VLE of CO2-H2O-amine systems [15].
f) Other types of VLE model
In addition to those described, there are also various other VLE models that have been
applied to correlate the experimental data and predict the VLE behaviour of acid gas-aqueous
alkanolamine systems, such as the electrolyte-UNIQUAC model used by Kaewsichan et al.
[184], the electrolyte-LCVM model [185], a quasi-chemical hole model [186], and an excess
Helmholtz energy-based equation of state [89, 187]. Although effort has been made to limit
the number of adjustable interaction parameters, experimental data are still indispensible in
the regression of model parameters. Unfortunately, the reported literature values generally
suffer from discrepancies with each other and cause uncertainties in the models.

3. Static-analytical Apparatus

59

Chapter 3

Solubility Measurements for CO2


in Aqueous Amines with a Staticanalytical Apparatus
A static-analytical apparatus has been constructed and employed to measure the solubility of
CO2 in various aqueous amine solutions. The apparatus was designed for operation at
temperatures from (298 to 423) K and at pressures up to 2.5 MPa. A stainless steel
equilibrium vessel was used. To facilitate reaching equilibrium, besides using a magnetic
stirrer bar in the vessel, a gear pump was also used to circulate the liquid phase. Liquid phase
compositions were analysed with gas chromatography, whereas the vapour phase was not
measured directly. Instead, only the total pressure and the initial solvent partial pressure were
measured, and were both used to calculate the CO2 partial pressure.

3. Static-analytical Apparatus

60

3.1 Introduction
As discussed in Chapter 2, apparatuses consisted of both glass- and alloy- (stainless steel,
Hastelloy etc.) have been employed to measure the VLE for the amine-H2O-CO2 systems
[69], although the latter have been more common in recent years. Experimental methods are
generally grouped into three categories: the static-synthetic method, the static-analytical
method and the methods using a flow system. The phase compositions can be either
calculated through mass balances, as in the synthetic method, or measured through chemical
titration or gas chromatography. For the static type of design, a variety of means are available
to enhance mass-transfer between the phases, including an external rocking design, an
internal magnetic/mechanical stirrer, and one or two circulation pumps. The type of design
adopted in this work was a static-analytical apparatus which consisted of an equilibrium
vessel, where the components were allowed to equilibrate, a gear pump to facilitate the
equilibration and liquid phase sampling, and gas chromatography which was used to analyse
the solution concentrations.
The final version of the equipment has gone through some evolutions since the initial
commissioning, as a result of issues with the liquid phase sampling. In this chapter, the
original design and the issues surrounding the phase sampling will be discussed in the first
section; this will be followed by the introduction of the revised version of the apparatus
which has been used for the experimental measurements in this work; experimental
procedures and calibration details will be addressed in the remaining sections.
3.2 Original design of the apparatus
3.2.1 Layout of the static-analytical apparatus
Fig. 3.1 shows the initial design of the apparatus used to measure the vapour-liquid
equilibrium of amine-H2O-CO2 systems in this work. The apparatus was designed for
operation at temperatures from (298 to 423) K and at pressures up to 2.5 MPa. Amine
solution was injected into the vessel with a manually-operated piston pump (#87-6-5, High
Pressure Equipment Company). The vessel was made from type 316 stainless steel and had
an internal volume of around 100 ml. An aluminium heater shell was used as the thermostat
device, and was powered by four 100 W cartridge heaters which were controlled to confine
the temperature fluctuation in the equilibration vessel to within 0.02 K. Mixing of the
components was achieved by using a titanium-encased magnetic stirrer bar and a liquid
circulation pump (mzr-7255, Michael Smith Engineers Ltd.). It was found that, to ensure

3. Static-analytical Apparatus

61

smooth operation of the pump, the system pressure should be 10 kPa or above. The cell
temperature was measured with a 100 platinum resistance thermometer which was
calibrated by the manufacturer with an uncertainty of 0.15 K. The resistance of the PRT was
measured with an Agilent model data acquisition unit (#34970A). Two pressure transducers
(one for 0 MPa to 0.2 MPa and the other for up to 3.5 MPa) were employed to measure the
pressures at low and intermediate pressures respectively. The relative uncertainty of the
pressure was reported to be 0.03% of full range and both transducers were checked at
ambient pressure before each experiment by comparison with a Delta OHM barometer which
has an uncertainty of 0.5 hPa. The Agilent data acquisition unit was connected to a
computer which allowed real time monitoring and storage of temperature and pressure data.

Figure 3.1 Schematic diagram of the static-analytic apparatus. 3WV: 3-way valve; AJ:
aluminium jacket; DAU: data acquisition unit; DG: degasser; GC: gas chromatograph; GP:
gear pump; GSV: gas sampling valve; HPG: high pressure generator; LSV: liquid sampling
valve; MS: magnetic stirrer; PC: computer; VP: vacuum pump; VS: vessel
A Perkin-Elmer gas chromatograph (Clarus 500) was employed to analyse the liquid phase
composition. The GC system was equipped with two detectors: one thermal conductivity
detector (TCD) for detection of CO2 and H2O and one flame ionisation detector (FID) for
amines. Two columns connected in series were required in order to separate all the
components under study: a 0.53mm30m RTX-35 Amine column was used to retain the

3. Static-analytical Apparatus

62

amine components, while CO2 and H2O were separated with a 0.53mm30m Elite-PLOT Q
column. To protect the PLOT Q column from being poisoned by amines, a 6-way bypass
valve was installed, which was switched to allow only CO2 and H2O to pass through the
PLOT Q column. Considering the reactive nature of amines, all the surfaces inside the GC
system were treated to be inert to amines, by using a base-deactivated liner (#20833-211.5,
Restek), a base-deactivated column and sulfinert tubing.
3.2.2 Gas and liquid sampling systems
Initially, the apparatus was designed to allow automatic gas and liquid sampling through the
gas sampling valve (GSV) and liquid sampling valve (LSV), both integrated with the gas
chromatography system. In Fig. 3.1, to better illustrate the circulation process, the GSV and
LSV are intentionally left outside of the GC system. In fact, the gas sampling valve was
located in a separate heated chamber integrated with the GC system, and the liquid sampling
valve was placed on top of the GC. To facilitate the liquid sampling, a gear pump was
employed to circulate the liquid phase. The liquid phase, passing through the LSV, was
finally returned to the gas phase of the equilibrium vessel. Gas phase was not circulated with
a pump. Instead, a vacuum pump was connected to the GSV, which allowed the sampling
loop to be evacuated and the loop to be filled with gas sample once the valve between the
GSV and the vessel was open.
To prevent condensation of the water and amine vapour samples when filling the sampling
valve, apart from placing the GSV in the heated chamber, the lines leading from the
equilibrium vessel to the valve were also wrapped with heating tape and maintained at 20 K
above the temperature in the cell. In contrast, the liquid circulation path including the gear
pump, the LSV and the tubing, were unheated in order to ensure that all the components
stayed in the liquid phase.
The GSV was rarely used in the actual experiment for two reasons: firstly, most of the
experiments were conducted at low to medium pressures. Filling the substantial volume of
the sampling loop of the GSV and the lines between the GSV and the equilibrium cell would
lead to the fluctuation of the system pressure and affect the equilibrium; secondly, no inert
gas, such as N2 or CH4, was used in this work. Therefore, it was not necessary to analyse the
gas phase concentration in order to derive the partial pressure of CO2.
During the sampling process, the liquid sample trapped in the LSV was pushed by the
carrier gas to the hot injection port. In the initial design, the tubing between the LSV and the

3. Static-analytical Apparatus

63

injection port was not heated. This was found to be problematic even in the test with wateracetone solutions. Fig. 3.2 shows the GC signal obtained with this initial design.
H2O
18

16

14

Acetone
12

10

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

2.6

2.8

Figure 3.2 GC signal for H2O-acetone solution with the tubing unheated
As can be seen from Fig. 3.2, the water peak has a long tail and asymmetric shape. This was
owing to the lag in the water movement along the unheated tubing. Fig. 3.2 shows clearly that
acetone has a much better peak shape than water which proves that the carrier gas was not
very effective in pushing the high boiling point liquid phase component. To solve this
problem, this section of tubing was also wrapped with heating tape and heated to as high as
473 K (Fig. 3.3). The improved sampling system led to a more symmetric water peak as
shown in Fig. 3.4.

Figure 3.3 Tubing connecting LSV and vaporisation chamber being wrapped with heating
tape

3. Static-analytical Apparatus

64

MEA

H2O

MEA

Figure 3.4 GC signal for MEA-H2O with the tubing heated to 473 K
The top and bottom diagrams represent signals from FID and TCD respectively. Obvious
improvement in the peak shape of water can be seen after the installation of the heating tape.
MEA, which has even higher boiling point than water, also showed a satisfactory peak shape.
However, when starting to measure the MEA-H2O-CO2 system, even with the tubing heated
up to 523 K, the obtained GC signal was very unsatisfactory (Fig. 3.5)

3. Static-analytical Apparatus

65

Figure 3.5 GC signal for CO2-H2O-MEA with the tubing heated to 523 K
Not only was the MEA peak shape irregular, the peak areas for MEA and CO2, which were
measures of the amounts, were much smaller than they should be. To identify the problem,
liquid samples were withdrawn with a manual syringe, and directly injected into the
vaporisation chamber, i.e. bypassed the LSV and heated tubing. Fig. 3.6 illustrates the GC
output signal obtained from the same liquid samples, through direct sampling with the
manual syringe.

3. Static-analytical Apparatus

66

Figure 3.6 GC signal for CO2-H2O-MEA through direct sampling with the manual syringe

Switching to the manual syringe sampling has greatly improved the peak shape for MEA.
Besides, the peak areas for MEA and CO2 were also much larger than the ones in Fig. 3.5.
This suggested that the problem lay in the sampling system, i.e. LSV and/or the tubing. It was
therefore suspected that the liquid sample only partially reached the vaporisation chamber
when using the LSV. If the tubing was left unheated, the obtained GC plot was shown in Fig
3.7. The peaks for MEA and CO2 were almost invisible. Certainly, heating the tubing was
essential to enable the liquid sample to reach the vaporisation chamber; however, it may not
be enough to guarantee all the samples entered into the GC system. As the tubing temperature
was high enough (> 473 K) to vaporise the liquid in-situ, it was unlikely that the sample,
which was already in the gaseous form, could not be pushed by the carrier gas to the
vaporisation chamber. More likely, part of the liquid sample was not pushed successfully by
the carrier gas out of the LSV and stayed there and then returned to the equilibrium system
when the valve was switched back.

3. Static-analytical Apparatus

67

H2O

Figure 3.7 GC signal for MEA-H2O-CO2 with the tubing unheated

Owing to the difficulty in verifying this supposition experimentally, a literature survey was
conducted instead to explore the issue. Hampe and Rudkevich [188] investigated the
reversible reactions between CO2 and primary amines using UV-vis and NMR. From the
experiment, they found that there was partial precipitation of carbamate salt when bubbling
N2 through the solution. This offered a possible explanation for the situation in the current
experiment: when the carrier gas, helium, serving the same function of nitrogen, could lead to
the partial precipitation of the carbamate salt. The precipitate could be even more difficult
than the liquid to be pushed through the LSV by the carrier gas.
Carbamate salt is only formed with CO2 in the case of primary /secondary amines. For
tertiary amines, there is no carbamate salt, and they should not encounter the same problem
as the MEA-H2O-CO2 system. However, this was not verified at this stage. Although the
problem was identified, it was not possible to solve it by heating the LSV because first, the
maximum operating temperature for LSV was only 348 K, which was much lower than the
boiling temperature of the amine; second, the liquid phase was supposed to be left at a lower
temperature than the vessel in order to prevent bubble formation during the interval of the
phase sampling. Therefore, it was not possible to fulfil these requirements with the continued
deployment of the LSV. When sampling with a manual syringe, the GC signals obtained

3. Static-analytical Apparatus

68

were much better than using the LSV (Fig. 3.6), therefore the revised design adopted later in
this work was based on the principle of using manual sampling without interfering the
equilibrium.
3.3 Revised design of apparatus
To allow sampling of the liquid phase with a manual syringe while not interrupting the
equilibrium, two 3-way valves were installed in the liquid circulating loop as shown in Fig.
3.8. As the system was sometimes under elevated pressure, as was the liquid sample, a
special type of manual syringe, Pressure-Flo (VICI, Thames Restek, UK), was deployed to
withdraw the liquid, which was capable of sealing samples up to 3.4 MPa. The pressure-Flo
syringe was also fitted with a cylinder adapter (MSHP), with which liquid samples could be
withdrawn at higher pressures. For lower pressure conditions (< 450 kPa), a septum was used
as the seal, which also allowed direct sampling (MSLP).

Figure 3.8 Schematic diagram of the static-analytic apparatus. 3WV: 3-way valve; AJ:
aluminium jacket; DAU: data acquisition unit; DG: degasser; GC: gas chromatograph; GP:
gear pump; GSV: gas sampling valve; HPG: high pressure generator; MS: magnetic stirrer;
MSHP: manual syringe for high pressure; MSLP: manual syringe for low pressure; PC:
computer; VP: vacuum pump; VS: vessel

3. Static-analytical Apparatus

69

3.4 Experimental procedures


Analytical grade monoethanolamine (MEA, 99.5%) and 2-amino-2-methyl-1-propanol (AMP,
99%) were supplied by Fisher Scientific. CP grade (purity: 99.995%) CO2 was used.
Deionised water obtained from a Millipore Direct-Q Ultrapure Water System was degassed
before diluting the pure amine. Degassing was achieved by placing a given amount of water
into a 500 ml glass bottle which was immersed in a water-filled ultrasonic bath (FB15049,
Fisher Scientific). Vacuum was applied to help remove dissolved gas in the liquid. The
degassing process normally lasted for 30 min to 1 h until no bubble evolution was observed.
The amine solution was then prepared by weighing desired quantities of the degassed water
and the amine into a 100 ml volumetric flask on a Kern analytical balance (AES 220-4,
Bucks). The repeatability and resolution of the balance were both 0.1 mg. A similar
degassing procedure was then executed on the prepared amine solution, except that only
partial vacuum was applied this time. Since the ultrasonic bath was kept at ambient
temperature and vacuum was applied only intermittently, the concentration of the prepared
solution was assumed to remain unchanged after the degassing process (Periodic checking
during the experiment with the GC analysis was also conducted to verify the solution
concentration stayed the same throughout the measurement).
Approximately 90 ml of the solution was loaded into the pre-evacuated vessel (VS) for each
experiment (the total volume of the system was estimated to be around 120 ml including 100
ml for the vessel volume and 20 ml for the dead volumes and the circulating line). The initial
pressure was below 0.1 kPa, so the remaining gas in the system was negligible. To avoid the
interference of the measurements caused by degradation of the amines, the solution was
refreshed once a week. The solution was brought to the desired temperature and the
corresponding vapour pressure was recorded when it reached thermal equilibrium. Then CO2
was introduced to the vessel from the cylinder to a desired pressure by adjusting the regulator
on the gas cylinder. It can be observed from experiments that MEA-H2O-CO2 systems reach
equilibrium faster than AMP-H2O-CO2, but an extended time of (10 to 12) h was allowed for
the equilibration for both amine systems. The attainment of equilibrium was verified by
checking stabilisation of the total pressure. We noted that owing to the low partial pressures
of CO2 in the experimental range of interest, many previous studies introduced an inert gas
(N2 or CH4) in order to increase the total pressure of the system [72, 73]. This work
eliminated the introduction of any inert gas so the CO2 partial pressures were simply

3. Static-analytical Apparatus

70

calculated by deducting the corresponding initially-measured vapour pressure of the solvent


from the total pressure.
In the present configuration, liquid samples were withdrawn through a septum using a
Pressure-Lok precision analytical syringe (0.5l, VICI). This syringe can seal samples in its
calibrated needle at pressures up to 3.5 MPa. Since each liquid sample is less than 1/105 of
the total liquid volume in the vessel, the disturbance of the equilibrium from the sampling
process was negligible. At least five samples were analysed for each equilibrium point. The
GC injection port was operated at

= 553 K, which was high enough to decompose the

reaction products and almost instantly vaporize liquid samples. Since the time for the samples
to stay in the injection chamber was very short, no degradation product should have formed
during the injection process and this was verified with the GC analyses where only peaks for
CO2, H2O and amines were seen. The TCD was maintained at
at

= 473 K while the FID was

= 553 K. A stream of 15 ml/min helium makeup gas was used to push the samples

quickly through the TCD detector to minimize the adsorption of amines onto the active sites
within the detector. The initial oven temperature of the GC was set to 313 K for 2 min and
then it was heated to

= 423 K at 10 K/min and held for 5 min. Helium, used as the carrier

gas, flowed through the GC at 15 ml/min.


3.5 Validation of the apparatus
The validation of the apparatus included: validation of the temperature and pressure
measurements; and the sampling, GC analysis system. The sensors were tested with
measurements of water vapour pressure at several temperatures; while the procedures for
analysis of liquid phase concentrations were validated by measuring the CO2 solubility in
aqueous MEA solutions. The former will be discussed in this section, and the latter will be
addressed in the next chapter.
3.5.1 Validation of the pressure and temperature sensors
The pressure and temperature sensors were validated through measuring the vapour pressure
of water. To fully validate the conditions for this work, the vapour pressure was measured
from (303 to 403) K and was compared with the literature data from NIST Chemistry
Webbook [189] as shown in Fig. 3.9. The AAD between the experimental data and the
literature values is 1.1%.

3. Static-analytical Apparatus

71

300

250

P / kPa

200

150

100

50

0
300

320

340

360

380

400

T/K

Figure 3.9 Comparison of water vapour pressure measurements from this work: ( ), and data
from NIST Chemistry Webbook [189] (

3.5.2 Calibration of the GC system


Before analysing the concentration of a component, it was essential to know its relation with
the response of the GC system. This required the calibration of the GC system with samples
of known concentrations/amounts. With the designs of apparatus (initial and revised) in this
work, two calibration methods have been examined and will be discussed below.
3.5.2.1 Calibration of the GC system with LSV and GSV
The initial plan was to calibrate the GC system with the LSV for the liquid components, i.e.
H2O and amines, and GSV for the gaseous component, i.e. CO2. This was based on the
method described by Silkenbumer et al. [71]. In their work, the liquid phase was circulated
through a liquid sample injection valve with a gear pump, and a Valco six-way valve was
used for gas sampling. The response factor, which is the ratio of the molar amount to the area
of a gas chromatographic peak, was determined on a relative basis: the response factor of
CO2 was related to that of water. By filling the equilibrium cell with the pure substances and
taking water with the LSV and CO2 with 6-WV, together with the knowledge of the volumes
of the liquid and gaseous samples, the response factor was obtained. This work was based on
a similar principle to Silkenbumer et al.s except that instead of relating the response factor

3. Static-analytical Apparatus

72

of CO2 to that of water to derive a relative response factor, this work determined the absolute
response factors for the two components separately. For CO2, the equilibrium vessel was
filled with pure substance at different pressures. The temperature of the GSV was measured
with a surface type Pt 100 RTD. Since the pressure of the GSV was the same as the vessel,
and the volume of the GSV was known from its specification as 0.5 ml, the amount of gas
filling the sampling loop was then calculated from the equation of state derived by Carroll et
al. [17]. Ten different pressures of CO2 were measured and the calibration curve is shown in
Fig. 3.10.
7
6

y = 0.2906x
R = 0.999

A 10-6

5
4
3
2
1
0
0

10
nCO2

106

15

20

/ mol

Figure 3.10 Relation between GC peak area and molar amount for CO2: Experimental data:
( ); Linear correlation: (

For water, different solutions of water and acetone were prepared and loaded into the vessel.
Subsequently, 0.2 L of liquid samples were withdrawn with the LSV. From the density of
the sample, which was determined separately with an Anton Paar density meter, its volume
and the concentration, the amount of water in the liquid sample was derived. The peak areas
were plotted against the molar amounts of water as in Fig. 3.11.

3. Static-analytical Apparatus

73

1.2
y = 0.087x
R = 0.991

A 10-6

0.8
0.6
0.4
0.2
0
0

6
nCO2

106

10

12

/ mol

Figure 3.11 Relation between GC peak area and molar amount for H2O: Experimental data:
( ); Linear correlation: (

To validate the calibration results, Henrys constant for CO2 in H2O at 313 K was measured
and calculated from the above calibration. However, the experimental results showed as
much as AAD = 37% discrepancy from literature values. This may have been caused by an
inaccurate knowledge of the volumes of the sampling valves, particularly the LSV, as the
overall volume was only 0.2 L. Since it was not convenient to measure 0.2 L of volume
accurately, the only way to bypass this information was to use the same valve to perform a
calibration for both CO2 and H2O, i.e. either liquefying CO2 so that both components could
be analysed with the LSV or vaporising H2O so that GSV could be used for both of them.
The first method was not achievable with the current setup owing to the constraints of the
maximum operating pressure of the gear pump; the second method was also tried, but the
repeatability of H2O concentration in the vapour phase was not as desired. It meant that the
absolute calibration method was not possible with the current equipment and as a result, a
novel calibration method based on the prior knowledge of the Henrys constant was
employed in this work.
3.5.2.2 Calibration of the GC system with LSV and Henrys constant
From the above experience, it was known that the successful calibration required both CO 2
and H2O to be sampled through the same valve. Considering that the Henrys constant for

3. Static-analytical Apparatus

74

CO2 in H2O has been thoroughly studied [17, 18], it was used to derive the relative response
factor of the two components as described below. The calibration method used in this work
was the relative response factor approach, i.e. the response factor of CO2 was determined
relative to that of H2O. The response factor ( ) is defined as the proportionality coefficient
between the peak area ( ) and the amount of the corresponding compound ( ):
(3.1)
Thus, the relative response factor of CO2 with respect to H2O is defined as:
(3.2)
In this work, the relative response factor was determined with an in-situ calibration by using
the Henrys constants of CO2 in H2O from literature data. The reference Henrys constants
were calculated from the model of Carroll et al. [17] which was developed from
approximately 100 experimental investigations below a total pressure of 1 MPa. Gas phase
nonideality in the experiment was accounted for using the virial equation of state truncated
after the second virial coefficient [129]. The CO2 to H2O concentration ratio in the liquid
phase was calculated from the corresponding Henrys constant and CO2 partial pressure. This
concentration ratio was then correlated to the ratio of the CO2 to the H2O peak area to obtain
the relative response factor as specified in eq. (3.2). A typical GC diagram is shown in Fig.
3.12.
H2O

CO2

Figure 3.12 GC diagram for CO2-H2O at


heated to 423 K

= 313 K, total

= 1884 kPa, and with tubing

3. Static-analytical Apparatus

75

The calibration was carried out at 313 K and six different total pressures. For each pressure,
at least five measurements were conducted. Table 3.1 summarises all the results including the
repeatability of the measurements at the same pressure (RD%). The

for the calibration

of the relative response factor was found to be 0.85%.


Table 3.1 Calibration results of relative response factor
P/kPa

at 313 K

RD%

422.3

0.003134

0.001787

0.5702

0.29%

669.4

0.005022

0.002836

0.5647

1.24%

871.6

0.006407

0.003687

0.5755

-0.66%

1163.6

0.008588

0.004894

0.5699

0.33%

1438.9

0.01056

0.006057

0.5736

-0.32%

1884.2

0.01352

0.007825

0.5788

-1.20%

Average

0.5721
0.85%

When measuring the samples from amine systems,

was larger in amine solution

than that in H2O alone. Therefore, the linearity of the TCD needed to be verified before
extending the calibration results to the calculation of the loadings of samples. The linear
dynamic range for the TCD was claimed to be of the order of 104, but was affected by the
design of specific detectors [190]. In this work, the linearity of the response factors for CO2
and H2O for our GC system was checked with methods as described in section 3.5.2.1. When
using linear correlations to represent the experimental data,

values for CO2 and H2O data

were 0.999 and 0.991 respectively. Based on these results, it was reasonable to extend the
TCD response relation derived from the calibration to the experimental measurements with
amine systems.

4. Experimental Results from the Static-analytical Apparatus

76

Chapter 4

Experimental Results from the


Static-analytical Apparatus for
CO2 solubility in aqueous amines
The static-analytical apparatus described in chapter 3 was first validated by measuring CO2
solubility in 30 mass% aqueous monoethanolamine (MEA) solutions at (313 and 393) K.
Subsequently, CO2 solubility in 30 mass% aqueous 2-amino-2-methyl-1-propanol (AMP)
solutions was measured from (313 to 393) K and the loading capacity was compared with
that of 30 mass% MEA under typical absorber and stripper conditions, i.e. (313 and 393) K.
The influence on loading capacity when replacing mass equivalent of AMP with piperazine
(PZ) was investigated with two different PZ concentrations, namely, 5 mass% and 10 mass%.
Further experimental measurements were conducted on CO2 solubility in 30 mass% aqueous
dimethylethanolamine (DMMEA) solutions and its mixture with 5 mass% PZ. Comparisons
of all the amines investigated in this work were presented and the implications on the solvent
selection were discussed.

4. Experimental Results from the Static-analytical Apparatus

77

4.1 Introduction
The static-analytical apparatus has been employed to measure the CO2 solubility in various
amine solutions, in order to study the influence of chemical structure on the loading capacity,
i.e. by comparing sterically-hindered amine (AMP) and tertiary amine (DMMEA) with
primary amine (MEA), and the impact on the cyclic loading capacity with the addition of an
activator, namely PZ to the AMP and DMMEA solutions. All the measurements were based
on the same overall amine concentration: 30 mass%. Table 4.1 summarises all the
experiments conducted in this work.
Table 4.1 Experimental conditions conducted in this work
Amine

Chemical Structure

Concentration

Temperature

MEA

30 mass%

(313, 393) K

AMP

30 mass%

(313, 333, 353, 373, 393) K

AMP+PZ

25 mass% AMP + 5

(313, 333, 373, 393) K

mass% PZ; 20 mass%


AMP + 10 mass% PZ
DMMEA

30 mass%

(313, 333, 353, 373, 393) K

DMMEA+PZ

25 mass% DMMEA + 5

(313, 333, 373, 393) K

mass% PZ

This chapter begins with the validation of the apparatus in measurements on the MEA-H2OCO2 system. After the validation, the sterically-hindered amine, in this case, AMP and its
activated mixtures with PZ is discussed. CO2 solubility results for AMP and AMP + PZ
obtained using the apparatus are also presented. This is followed by discussion on tertiary
amine, particularly DMMEA together with its PZ blend. Finally, comparisons of the amine
systems studied in this work are covered which lead to the discussion on the selection of
amines in the context of carbon capture.

4. Experimental Results from the Static-analytical Apparatus

78

4.2 Experimental uncertainty


The standard uncertainty in the loading measurements was calculated based on equation (4.1),
taking into account the contribution from calibration of the chromatograph, pressure and
temperature measurements:
(4.1)
where

is the combined standard uncertainty in the loading,

kPa) or 1 kPa (> 200 kPa),

= 0.05 K, and

= 0.06 kPa (< 200

is the standard uncertainty related to the

calibration of chromatograph for component which was estimated to be below 1.5%. This
included both the standard relative deviation of the measurements and the model uncertainty
estimated from Carroll et al. [17]. The derivatives in Eq. (4.1) were estimated using the
thermodynamic model (- approach) which will be discussed in Chapter 5, and the
combined uncertainty in loading measurement was found to be approximately 0.006.
A second contribution of uncertainty is from the repeatability of the concentration
measurements caused by the random sources of errors. This was quantified in terms of
standard deviations:

(4.2)
where

is the average loading for each equilibrium point from

The relative standard deviation in loading

times of measurements.

was found to be below 1.5%.

4.3 Validation of the static-analytical apparatus by measuring CO2 solubility in aqueous 30


mass% MEA solutions
4.3.1 MEA in the context of carbon capture
MEA is undoubtedly the most commonly used amine in industry. Before being utilised for
carbon capture, MEA has been widely used in natural gas processing, ammonia production
and other gas sweetening processes [191].
(1) Solution chemistry
The most important reaction for primary amines such as MEA is the carbamate formation:

4. Experimental Results from the Static-analytical Apparatus

79

(4.3)
Although other reactions can also take place (CO2 reacts with water or hydroxyl ion to form
bicarbonate), they are generally considered negligible compared to (4.3):
(4.4)
(4.5)
(2) Kinetics
One of the advantages of MEA is its fast CO2 absorption rate [192] which is attributed to the
readiness to form carbamate. The kinetics of the system has been widely studied and was
summarised in the work of Mahajani and Joshi [193], Versteeg et al. [194] and Aboudheir et
al. [195]. The zwitterions mechanism reintroduced by Danckwerts [196] based on the earlier
work of Caplow [197] was generally considered to be adequate in describing the reaction
kinetics. It involves a formation of an intermediate, followed by the deprotonation of the
compound by a base (B) such as H2O, MEA, or OH-:
Mechanism MEA_I:

(4.6)

(4.7)
Aboudheir et al. [195] conducted kinetic measurements over the temperature range from (293
to 333) K, MEA concentration range from 3 to 9 M, and CO2 loading from 0.1 to 0.49. They
concluded that only the single-step termolecular mechanism of Crooks and Donnellan [198]
could be used to explain all observed kinetic phenomena. The termolecular mechanism is
illustrated as:

4. Experimental Results from the Static-analytical Apparatus

80

Mechanism MEA_II:

(4.8)
(3) Previous work on solubility measurement of MEA-H2O-CO2 system
The solubility of CO2 in different concentrations of aqueous MEA has been measured in
numerous previous studies as listed in Table 4.2.

4. Experimental Results from the Static-analytical Apparatus


Table 4.2 Summary of literature for CO2 solubility in aqueous MEA solution
*Converted from molarity assuming solution density = 1000 g/L

81

4. Experimental Results from the Static-analytical Apparatus

82

4.3.2 Validation results


The setup was validated by measuring the solubility of CO2 in a 30 mass% aqueous solution
of MEA at (313 and 393) K and pressure range of (3 to 295) kPa. No inert gas, such as N2,
was used to dilute the CO2; therefore it was possible to derive the CO2 partial pressure (from
total pressure and initial solvent vapour pressure) without directly measuring the gas phase
compositions. Liquid phase compositions were determined with GC at least five times for
each equilibrium point.
1000

100000
10000

100
10

100

PCO2 /kPa

PCO2 /kPa

1000

10
1

1
0.1

0.1
0.01

0.01

0.001

0.001
0

0.2

(a)

0.4
0.6
mole_CO2

0.8

0.1

(b)

0.2

0.3

0.4

0.5

mole_CO2

Figure 4.1 Comparisons of solubility data of CO2 in 30 mass% aqueous MEA solutions at
313 K (a) and 393 K (b): (
(

): This work; (

): Lee et al., [32]; ( ): Shen and Li, [33]; (

):Jou et al., GC and titration methods [9];


): Mamun et al., [12]; ( ) Goldman and

Leibush [34]
Fig. 4.1 shows the results for MEA-H2O-CO2 in comparison with the four data sets from the
literature that pertain to the temperature and concentration of this study. Our data at

= 313

K agree well with Jou et al. [9], while the loading data of Lee et al. [32], and Shen and Li [33]
deviate from those of Jou et al. by about 0.04. Jou et al. claimed that the consistent offset
between their data and those of Lee et al. was caused by neglecting the remaining CO2 in the
acidic solution in Lees analysis of the liquid phase loading. For

= 393 K, less scattering

exists in the previously reported data, and our results showed satisfactory agreement with the
published values.
The validation results were also quantified with a thermodynamic model which will be
discussed in chapter 5. Table 4.3 gives

between the measured values (i.e. Jou et al. and

this work) and the model-correlated results. As can be seen, Jous data show

of 3.6%,

4. Experimental Results from the Static-analytical Apparatus


while

83

for data from this work is 2.1%. This demonstrates that the apparatus used in this

work is capable of providing reliable and repeatable results. Owing to the calibration methods
used in this work, the validation of the high loading range is enough to justify the accuracy of
the entire loading range and is irrespective of temperature conditions and amine types.
Table 4.3 Comparison of CO2 (1) solubility data in a solution of water (2) (70 mass%) and
MEA (3) (30 mass%) at

= (313 and 393) K from Jou et al. [73] and this work

Jou et al.
= 313 K

This work
= 313 K
_model

0.00147
0.00896
0.0677
0.604
2.57
8.09
36.1
103
293
593
993
2992
5986
9969
14945
19914
= 393 K

0.0888
0.203
0.365
0.461
0.513
0.557
0.609
0.646
0.709
0.794
0.844
0.965
1.049
1.097
1.132
1.18

0.090
0.201
0.366
0.466
0.504
0.535
0.589
0.645
0.722
0.786
0.837
0.962
1.054
1.116
1.146
1.148
_model

0.00202
0.0221
0.0984
2.29
46.8
122
222
422
822
2804
5809
9770
14741
17723

0.00333
0.0112
0.0247
0.119
0.349
0.403
0.444
0.473
0.536
0.644
0.719
0.78
0.829
0.863

0.00295
0.00928
0.0199
0.0965
0.327
0.383
0.431
0.480
0.533
0.644
0.720
0.778
0.832
0.862

RD%
1.9%
-0.8%
0.2%
1.0%
-1.7%
-4.0%
-3.3%
-0.2%
1.9%
-1.1%
-0.9%
-0.3%
0.4%
1.7%
1.2%
-2.7%
RD%
-11.4%
-17.1%
-19.4%
-18.9%
-6.3%
-5.0%
-2.9%
1.5%
-0.7%
-0.1%
0.1%
-0.2%
0.3%
-0.1%
3.6%

3.95
19.12
71.50
101.00
159.40
161.52
211.92
297.02
408.17
= 393 K
18.00
44.45
76.63
90.63
91.92
121.62
171.2195
186.53
188.82
191.04
236.82
295.20

= 7.24 kPa
0.530
0.514
0.585
0.562
0.632
0.623
0.639
0.644
0.668
0.675
0.670
0.676
0.687
0.696
0.706
0.724
0.748
0.751
= 164.46 kPa
0.211
0.292
0.338
0.351
0.356
0.382
0.406
0.415
0.42
0.434
0.443
0.461

0.224
0.298
0.342
0.352
0.358
0.360
0.393
0.410
0.417
0.418
0.436
0.453

RD%
-3.0%
-4.0%
-1.5%
0.7%
1.0%
0.8%
1.3%
2.5%
0.4%
RD%
6.1%
2.1%
1.3%
0.2%
0.7%
-5.7%
-3.2%
-1.1%
-0.7%
-3.6%
-1.6%
-1.7%
2.1%

4. Experimental Results from the Static-analytical Apparatus

84

4.4 Solubility of CO2 in aqueous AMP and AMP+PZ solutions


After successfully validating the apparatus, it was used to study the loading capacity of
different types of amines. Sterically-hindered amines and tertiary amines were of particular
interest in this work for the reason that both of these types are believed to have much larger
potential capacity than primary or secondary amines [192]. In order to make fairly easy
comparison, the amines chosen for study in this work were the simplest stercially-hindered
and tertiary equivalents of MEA, namely, AMP and DMMEA. Since both these amines are
subject to the drawback of slow reaction kinetics with CO2, an activator is often added to the
system to enhance the reaction rate. Thus, in this work, PZ activated AMP and DMMEA
solutions were also studied. We will first look at the results for AMP and AMP+PZ systems,
and in the next section, the DMMEA and DMMEA+PZ systems will be discussed.
4.4.1 Sterically-hindered amines in the context of carbon capture
(1) Solution chemistry
Sartori and Savage [64] proposed that a new class of amines, sterically-hindered amines,
should be used for CO2 capture. These are able to approach a capacity of 1 mole of CO2 per
mole of amine while retaining absorption rates comparable to those secondary amines. AMP
is the most well known of these sterically-hindered amines in the context of CO2 capture
processes, owing to the fact that it is the simplest hindered form of MEA, and the difference
in properties can be attributed to the steric hindrance [20]. As a primary amine, it is still
generally agreed that AMP reacts with CO2 directly to form carbamate. This was verified by
Xu et al. [207] who studied the absorption of CO2 in a nonaqueous solvent, e.g. AMP+1propanol. Not only was the reaction kinetics well represented with the zwitterions mechanism,
but also white carbamate precipitate was observed during their experiment. Two proposed
mechanisms similar to those for MEA are presented below:
Mechanism AMP_I is the zwitterion mechanism originally proposed by Caplow
[197]. He assumed that a hydrogen bond is formed between the amine and a water
(or AMP/OH-) molecule before it reacts with CO2 molecule:

4. Experimental Results from the Static-analytical Apparatus

85

(4.9)
The proton transfer within a hydrogen-bonded complex is exceedingly rapid in the
equilibrium-favoured direction. The first step is rate-limiting, and involves nucleophilic
attack of the nitrogen atom on the carbon atom leading to the zwitterion formation. Later
studies [64, 208] divided the zwitterion mechanism into two steps: instead of assuming that
there is hydrogen bonding between amine and water molecules in the first place, the CO 2 was
assumed to react with amine first to form a zwitterion, and then a base molecule deprotonates
the zwitterion:

(4.10)

(4.11)
Here, B is a base molecule, in this case, either H2O, AMP or OH-. The rate of deprotonation
is strongly dependent upon the basicity of the molecule B and hence by the pH of the solution.
The steric hindrance of the zwitterion and molecule B also has effect.
Mechanism AMP_II is a single-step, termolecular reaction proposed by Crooks and
Donnellan [198]:

4. Experimental Results from the Static-analytical Apparatus

86

(4.12)
The main difference between the two mechanisms is whether the hydrogen bonding is formed
before the amine reacts with CO2 or both reactions take place simultaneously. Previous
studies claimed that the rate expression of the zwitterion mechanism fits better with the
experimental data [209]. However, some authors argued that both mechanisms fit equally
well [210].
Sharma [211, 212] found that the steric hindrance effects reduce the stability of the
carbamates. Chakraborty et al. [213] reported 13C NMR data which show that the carbamate
of 2-amino-2-methyl-1-propanol is formed to a much lesser extent than the carbamte of the
corresponding unsubstituted amine, i.e. monoethanolamine. Xu et al. [145] also measured the
concentration of the carbamate and found it to be only of the order of 10-4 of the amine
concentration. This piece of evidence suggests that the carbamates of sterically-hindered
amines may readily undergo hydrolysis, leading to the formation of bicarbonates and free
amine molecules:

(4.13)
Vaidya et al. [68] suggested that zwitterion may directly undergo hydrolysis bypassing the
formation of carbamate:
(4.14)
Sartori and Savage [64] attributed the instability of the carbamates to steric hindrance caused
by the substitution on the -carbon adjacent to the amino group. Chakraborty and co-workers
[215] investigated effects of substituents on the -carbon using molecular orbital approach.

4. Experimental Results from the Static-analytical Apparatus

87

They concluded that the interaction of the lone-pair orbital with the unfilled methyl group
orbitals should lead to a lower charge at the donor site. Applying hard and soft acid-base
theory, the effects of methyl substitution at the -carbon atom make the amine a softer base.
Since CO2 falls into the hard acid category, therefore, the softer the base, the weaker the
NC bond in the zwitterion/carbamate. Moreover, OH- as a hard base reacts more favourably

with CO2 to form bicarbonate.


Apart from directly reacting with CO2, AMP may contribute the CO2 dissolution process via
increasing in solution pH and base catalytic effect, shown in Mechanism III and IV.
Mechanism AMP_III describes the process of AMP protonation which raises the solution pH
and promotes the formation of H2CO3/HCO3-:

(4.15)
(4.16)
(4.17)
Mechanism AMP_IV illustrates the base catalytic effect of AMP and is similar to (4.12)
except that the positions of H2O and AMP exchange. This is believed to be less favourable
than (4.12), since AMP has a stronger nucleophilicity than H2O:

(4.18)
(2) Kinetics
The zwitterion mechanism is generally accepted to be adequate for describing the reaction
kinetics for CO2 dissolution in aqueous AMP [207]. Xu et al. [207] measured the apparent
reaction rate constant of AMP in a stirred-cell reactor. They found that the zwitterion
formation is the rate-limiting step, followed by easy deprotonation and rapid hydrolysis of the
carbamate. In terms of the reaction rate constants at 298 K, Xu and co-workers [207]

4. Experimental Results from the Static-analytical Apparatus

88

compared three sterically-hindered amines in a concentration range of (0.25 to 3.5) kmol/m3


and found that:

(810.4 m3/kmols) <

Although much higher than that of MDEA (

(2375 m3/kmols) <

(2585 m3/kmols).

at 298 K = 18.2 m3/kmols as measured by

Mimura et al. [216]), the reaction rate of AMP is much slower than MEA (

at 298 K =

3630 m3/kmols as measured by Mimura et al. [216]).


To enhance the reaction rate of AMP, an activator is usually added, normally primary (e.g.
MEA) or secondary amines (e.g. DEA) with fast kinetics. Piperazine is an effective promoter
identified by many previous investigators [217-220]. It is a diamine and can react with CO2 to
form both single and dicarbamate products:

(4.19)

(4.20)
Besides, it can protonate to form single and diprotonated PZ and protonated carbamate:

(4.21)

(4.22)

4. Experimental Results from the Static-analytical Apparatus

89

(4.23)
The reactions of CO2 in the AMP+PZ mixture are generally summarised as mutuallyenhanced reactions shown in (4.24) and (4.25), according to Samanta et al. [223]:
(4.24)
(4.25)
The above reactions are considered to be significantly faster than the reactions if the PZ was
absent, which explains the rate-enhancing effect of PZ to other amines, as shown in (4.26)
and (4.27):
(4.26)
(4.27)
Bishnoi and Rochelle [221] measured the rate constant of CO2 absorption into aqueous
piperazine and concluded that PZ is an effective promoter because of its large rate constant
(i.e. an order of magnitude higher than primary amines such as MEA or DGA) and
comparable first carbamate stability constant. In other studies, it has been observed that at
comparable conditions of temperature, PZ concentration and CO2 partial pressure, the PZAMP blends exhibit larger absorption flux than PZ in isolation [222, 223]. Puxty and
Rowland [224] studied the CO2 mass transfer in PZ-AMP mixture with a model taking into
account chemical reactions and diffusion in a falling thin film. They claimed that by
comparing the concentration profiles in the film at the same conditions with and without
AMP, which shows a greater PZ concentration near the gas-liquid interface in the former
system than latter, the reason for the rate enhancement effect of AMP on PZ becomes clear:
although the two amines have similar

values (i.e. 9.46 for PZ and 9.29 for AMP at 313

K), the higher concentration of AMP in the solution means that it takes the priority in
accepting protons, which results in more free PZ in the solution enhancing the mass transfer.
This has two implications: first, the rate enhancement effect is mutual, i.e. the amine blends
have fast kinetics than either individual amine under similar conditions; second, to generate
an effective mixture, the bulk amine (e.g. AMP) should have a higher

value than the

4. Experimental Results from the Static-analytical Apparatus

90

promoter (e.g. PZ) to allow the former the priority in accepting protons while retaining the
concentration of the promoter in the solution. This can be achieved either with a large
difference in
concentration ratio

values between the bulk amine and the promoter or by keeping the
at a relatively high level.

Samanta et al. [223] measured the reaction rate of absorption of CO2 into PZ activated
aqueous AMP solutions using a wetting wall contactor. They found that by replacing 2 mass%
AMP with 2 mass% PZ, the reaction rate increases to 3.3 times of the original 30 mass%
AMP; a further 3 mass% PZ replacement of AMP leads to the kinetics to 4.6 times of the
original level; replacing 8 mass% AMP with 8 mass% PZ results in the kinetics rising to 5.6
times of the 30 mass% AMP reference value. Apparently, there is a diminished enhancing
effect when increasing the amount of AMP replaced with PZ (the enhancement effect of PZ
to other amines tends to level off as its concentration increases). If combining the rate
enhancement results with reaction rates for AMP and MEA published by Xu et al. [207] and
Littel et al. [208] respectively, it can be deduced that 25 mass% AMP + 5 mass% PZ has a
comparable reaction rate as 30 mass% MEA solutions (see section 7.2.1 for more detail).
In this work, apart from determining CO2 solubility in 30 mass% aqueous AMP solutions, we
also measured in two different AMP+PZ mixtures including: 25 mass% AMP+5 mass% PZ
and 20 mass% AMP + 10 mass% PZ. The aim was to investigate the influence on solvent
loading capacity at typical absorption conditions when replacing equal mass% of AMP with
PZ.
(3) Previous work on solubility measurement of AMP-(PZ)-H2O-CO2 system
Limited solubility data have been reported for CO2 in aqueous AMP, particularly at stripper
conditions (373 to 393) K as shown in Table 4.4.

4. Experimental Results from the Static-analytical Apparatus

91

Table 4.4 Summary of literature for CO2 solubility in aqueous AMP and AMP+PZ blends
Author
AMP
Kundu et al. [92]
Park et al. [150]
Roberts and Mather [225]
Seo and Hong [109]
Silkenbumer [71]
Teng and Mather [105]
Tontiwachwuthikul et al. [144]

Concentration (mass %)

/ kPa

/K

18,25,30*
30
18,28.5*
30
18, 35.9**
18*
18, 26.7*

303, 313, 323


313, 333, 353
313, 373
313, 333, 353
313, 333, 353
313,343
313, 333, 353

3.25-94
0.041-206.8
1.25-866
3.94-336
7.3-2743
0.3-500
1.4-82.7

Dash et al. [218]

38 wt% AMP + 2 wt% PZ,


35 wt% AMP + 5 wt% PZ,
32 wt% AMP + 8 wt% PZ

313

0.127-140.4

Yang et al. [149]

22.9 wt% AMP + 5.5 wt% PZ*,


22.9 wt% AMP + 11.1 wt% PZ*,
22.9 wt% AMP + 16.6 wt% PZ*,
34.4 wt% AMP + 5.5 wt% PZ*,
34.4 wt% AMP + 11.1 wt% PZ*,
34.4 wt% AMP + 16.6 wt% PZ*

313, 333, 353

1.06-132.4

AMP+PZ

*Estimated from molarity based on the assumption that the density of amine solution = 1 gml-1
**Converted from molality

4.4.2 Experimental results on CO2 solubility in aqueous AMP solutions


For AMP-H2O-CO2 system, experimental results obtained at
summarised in Table 4.5.

= (313 to 393) K are

4. Experimental Results from the Static-analytical Apparatus

92

Table 4.5 CO2 (1) solubility data in a solution of water (2) (70 mass%) and AMP (3) (30
mass%) at

= (313, 333, 353, 373 and 393) K

The data are also plotted together with published literature data in Fig. 4.2. Most of the
literature data are reported at the lower end of the temperature range ( 353 K) for 30 mass%
AMP solutions. Only one set of measurements conducted by Li and Chang [72] is at
K. As can be seen from Fig. 3, at

= 373

= 353 K, there is good agreement among the data from

this work, Li and Chang, and Seo and Hong [109]. At

= 313 K, our results agree well with

the data from Kundu et al. [92] and most of the points from Seo and Hong. The data at

333 K and CO2 loadings above 0.6 show the largest discrepancy among all the literature
sources with our data bisecting the data from Li and Chang, Seo and Hong, and
Tontiwachwuthikul [144]. For the high temperature range, Li and Changs data at

= 373 K

and loadings below 0.4 show a richer liquid phase than our results at the same CO2 partial
pressures. This may be caused by withdrawing the gas phase from the equilibrium system in
Li and Changs work, as at high temperatures ( = 373 K), there is significant amount of the
total solvent in the vapour phase. So depending on the sample size, it may interrupt the

4. Experimental Results from the Static-analytical Apparatus

93

equilibration and change the solvent composition. This is, however, uncertain without
knowing the details of their experimental procedures.

1000

1000

393 K

353 K

10

373 K

100
PCO2 /kPa

PCO2/ kPa

100

10

333 K

313 K
1

1
0

0.2

0.4
0.6
mole_CO2

0.8

(a)

0.2

0.4
0.6
mole_CO2

0.8

(b)

Figure 4.2 Solubility of CO2 in 30 mass% aqueous AMP solutions at


and 393 K and (b): 333K and 373 K. (

), (

= (a): 313 K, 353 K,

), ( ), ( ), ( ): this work; ( ): Kundu et al.,

[92]; ( ): Tontiwachwuthikul et al. (26.7 mass% converted from 3M AMP), [144]; ( ), (


), (

): Li and Chang, [72]; (

), ( ): Seo and Hong, [109]

4.4.3 Experimental results on CO2 solubility in aqueous mixtures of AMP and PZ


To evaluate the influence of replacing AMP with equal mass concentration of PZ, two
different concentrations of the amine blends, i.e. 25 mass% AMP + 5 mass% PZ, and 20
mass% AMP + 10 mass% PZ, were measured at four temperatures as summarised in Table
4.6 and 4.7.

4. Experimental Results from the Static-analytical Apparatus

94

Table 4.6 CO2 (1) solubility data in a solution of water (2) (70 mass%), AMP (3) (25 mass%)
and PZ (4) (5 mass%) at

= (313, 333, 373 and 393) K

Table 4.7 CO2 (1) solubility data in a solution of water (2) (70 mass%), AMP (3) (20 mass%)
and PZ (4) (10 mass%) at

= (313, 333, 373 and 393) K

4. Experimental Results from the Static-analytical Apparatus

95

There were no literature data conducted at the same amine concentrations as this work. Fig.
4.3 illustrates all the data from this work.
1000

393 K

373 K

PCO2/ kPa

100

373 K
333

10

313 K

1
0.0

0.2

0.4

0.6

0.8

1.0

mole_CO2

Figure 4.3 CO2 solubility in aqueous AMP and PZ blends at

= (313, 333, 373, and 393) K:

( ,

): 20 mass% AMP + 10 mass%

, ): 25 mass% AMP + 5 mass% PZ; ( ,

PZ
As can be seen from Fig 4.3, the solubility of CO2 in the two AMP and PZ blends of different
concentrations is quite close at the same temperature conditions, except at 393 K, where more
PZ in the solution leads to increase in the solution loading. More detailed discussions on the
influence of the loading capacity from the types of amines will be presented in section 4.6.
4.5 Solubility of CO2 in DMMEA and DMMEA solutions
DMMEA is a tertiary amine with exactly the same types and quantities of functional group as
AMP, except that the -CH3 branch is not on the -carbon. Instead, it is directly connected to
the nitrogen atom, which leaves no hydrogen atom on the amine functional group. This
eliminates the possibility of carbamate formation between the DMMEA and CO2, which have
two direct effects on the amine properties: first, compared with primary and secondary
monoamines, DMMEA has larger absorption capacity; second, the energy required to reverse
the amine-CO2 reaction is lower for DMMEA than for amines which can form carbamte with

4. Experimental Results from the Static-analytical Apparatus

96

CO2. In this work, the solubility of CO2 in 30 mass% aqueous DMMEA solution and 25 mass%
DMMEA + 5 mass% PZ blended system was measured at temperatures from (313 to 393) K.
4.5.1 Tertiary amines in the context of carbon capture
(1) Solution chemistry
The most commonly used tertiary amine in industry is methyldiethanolamine (MDEA),
which can be found in selective removal of the acid gas (H2S and CO2) from natural gas and
refinery gas streams [191]. Compared with a typical primary amine (i.e. MEA), it offers the
benefit of high acid gas loading, low enthalpy of reaction, low solvent loss (due to low
vapour pressure and slow degradation rates), low corrosion even at high solution loadings
which also means a high solution concentration is possible (up to 50 to 55 mass%). The
major disadvantage of MDEA is its slow reaction rate with CO2. Since tertiary amines do not
have a hydrogen atom on the amine group, unlike primary and secondary amines, it is
generally considered to not react directly with CO2 to form carbamate [226-228]. Instead it
contributes to the reaction with its contribution to solution basicity and catalytic effect on
CO2-H2O reaction as described by mechanisms I and II [228].
Mechanism MDEA_I: MDEA protonates and increases the basicity of the solution, and
enhances the reaction of CO2 with H2O/OH-.

(4.28)
(4.29)
(4.30)
Mechanism MDEA_II: MDEA forms hydrogen bonding with H2O which weakens the
original H-O bond and increases the nucleophilic strength of oxygen in H2O. This facilitates
its attack of the carbon atom in CO2 as illustrated below:

4. Experimental Results from the Static-analytical Apparatus

97

(4.31)
In the above scheme, due to the involvement of MDEA, the reaction barrier was reduced
significantly compared to CO2-H2O reaction, according to the quantum mechanical
calculation conducted by Silva and Svendsen [210]. The authors also suggested that this
effect could take place for a number of Brnstad bases, including primary and secondary
amines. However, for these amines, the reactions with CO2 are dominated by carbamate
formation and base catalysis plays a minor role. Barth et al. [228] investigated the relative
importance of these two mechanisms and concluded that mechanism I was much more
important than mechanism II for MDEA system, with less than 6% contribution in the
reaction rate expression coming from the latter. However, contradictory results were reported
by Littel et al. [208], who claimed that the base-catalysed mechanism is the predominant
reaction for tertiary amines. Apart from the above two reaction pathways, Barth et al. [228]
proposed another possible mechanism with the formation of zwitterionic intermediary,
analogous to primary and secondary amines:
Mechanism MDEA_III:

(4.32)

(4.33)

(4.34)

4. Experimental Results from the Static-analytical Apparatus

98

Barth et al. [228] pointed out that it was not possible to differentiate between mechanism II
and III on kinetic grounds. However, the former was more favoured than the latter. This was
because, for primary and secondary amines, the loss of one proton from the nitrogen atom
allowed the formation of a partial

double bond and made the zwitterions formation

more likely than tertiary amines.


Despite the many postulated reaction routes, the reaction mechanisms are far from clear for
the amine-CO2 system. Some researchers hinted at other possible reactions such as the
formation of an ion pair between the solvated tertiary amine and dissolved CO2 by
conducting experiments in non-aqueous solutions [194, 229-231]. For the CO2-MDEAethanol system, recent studies carried out by Sada et al., Kierzkowska-Pawlak and Zarzyzki
and Park et al. [229-231] suggested that the dissolution of CO2 in MDEA-ethanol was not
just a physical process owing to the following experimental observations: first, N2O solubility
in ethanol solutions of MDEA decreased with the rise of amine concentration whereas the
CO2 solubility strongly increased; second, the reaction of CO2 with MDEA in ethanol
followed pseudo-first-order reaction kinetics. This conclusion was contradictory to earlier
work from Versteeg, which stated that the absorption rate of CO2 in to MDEA-ethanol
solution could be described completely with the non-stationary mass balance for physical
absorption and was almost identical to the absorption rate of N2O, provided the differences in
physical constants were corrected.
In this work, a less studied tertiary amine, DMMEA was investigated for its CO2 absorption
capacity. As described above, although the reaction mechanisms for tertiary amines are not
fully understood, it was generally agreed that using base catalysis of the CO 2 hydration
(termolecular) model proposed by Donaldson and Nguyen [232] can satisfactorily interpret
the experimental data. Two important reaction mechanisms for CO2 reaction with aqueous
DMMEA are as below:
Mechanism DMMEA_I:

(4.35)
(4.36)

4. Experimental Results from the Static-analytical Apparatus

99
(4.37)

Mechanism DMMEA_II:

(4.38)
(2) Kinetics
Despite the catalytic effect of tertiary amines, their aqueous solution reaction rates with CO2
are still too slow to make them economically justifiable for industrial scale applications.
Henni et al. [233] investigated the reaction kinetics of CO2 in seven primary, secondary and
tertiary amines including MEA, DEA, MDEA, DMMEA, DEMEA, 1-AP and 3-AP using the
stopped-flow technique. For the three tertiary amines studied, they concluded that the
reaction rate for CO2 in DMMEA was higher than that in MDEA, but not in DEMEA. Based
on the interpretation of the kinetic data obtained, they also confirmed that the mechanism of
the base catalysis of the CO2 hydration proposed by Donaldson and Nguyen [232] was
appropriate for describing the CO2 absorption into aqueous DMMEA solution.
As with sterically-hindered amines, a common way to enhance the reaction rates of tertiary
amines is by mixing them with promoters. The most well studied blended systems are
MDEA-based, in which primary amines (such as MEA, DGA), secondary amines (such as
DEA, DIPA), or other amines with fast kinetics (such as PZ) have been added as rateenhancer [220, 234-240]. The generally agreed mechanism for the CO2 absorption into
promoted MDEA solutions involves zwitterion formation, followed by deprotonation by a
base molecule in the system.
(4.39)
(4.40)
If MDEA has a higher pH than the promoter, then it will preferably react with the zwitterions.
As a result, larger amount of promoter will remain in the solution to enhance the overall

4. Experimental Results from the Static-analytical Apparatus

100

reaction rate. Despite many previous studies on the amine mixtures, the reaction mechanisms
and expressions for the rate constants are far from reaching a general consensus. Liao and Li
[241] investigated the rate enhancement effect of MEA on MDEA in CO2 absorption and
summarised the reaction rate with a model comprising a zwitterions mechanism for MEA and
a first-order reaction mechanism for MDEA. On the contrary, Ramachandran et al. [242]
pointed out that neither zwitterions nor termolecular mechanism is sufficient to describe CO2MDEA-MEA system without modification and consequently proposed a revised
termolecular-based model including the contribution from hydroxide. For DEA-promoted
aqueous MDEA solution, Littel et al. [208] suggested that the deprotonation of zwitterions is
the rate-limiting step, and by changing the concentration and type of the tertiary amine, the
rate of this deprotonation could be affected significantly. However, Rinker et al. [239]
claimed the opposite by concluding that the contribution to the zwitterion deprotonation was
mainly from DEA rather than MDEA. Zhang et al. [236] suggested that the overall process
can be regarded as a CO2 transferring mechanism where DEA first reacts with CO2, and then
regenerates itself through transferring CO2 to MDEA. Apart MEA and DEA, PZ is also
commonly used as an activator and has been commercially used by BASF to promote MDEA
[243].
As far as we know, there is only one publicly available kinetic study on DMMEA-PZ system
[244]. They found that, among all the measured tertiary and sterically-hindered amines,
DMMEA and AMP have the largest cyclic capacity; while considering the activators MAPA
and PZ are both better than MEA. Another DMMEA based system is studied by Littel et al.
[245], who investigated the reaction mechanisms of secondary amines (DEA and DIPA)
blended with tertiary amines (MDEA, DMMEA, DEMEA, and TEA).
4.5.2 Experimental results on CO2 solubility in aqueous DMMEA and DMMEA+PZ
solutions
A complete set of CO2 solubility data in 30 mass% aqueous DMMEA solutions has been
measured at temperatures from (313 to 393) K. The results are summarised in Tables 4.8 and
4.9.

4. Experimental Results from the Static-analytical Apparatus

101

Table 4.8 CO2 (1) solubility data in a solution of water (2) (70 mass%), and DMMEA (3) (30
mass%) at = (313, 333, 353, 373 and 393) K

Table 4.9 CO2 (1) solubility data in a solution of water (2) (70 mass%), DMMEA (3) (25
mass%), and PZ (4) (5 mass% ) at

= (313, 333, 373 and 393) K

To compare the solubility of CO2 in DMMEA and DMMEA+PZ systems, the data obtained
are plotted in Fig. 4.4.

4. Experimental Results from the Static-analytical Apparatus

102

1000

393 K

373 K

PCO2/KPa

100

353 K
10

333 K

313 K

1
0

0.2

0.4

0.6

0.8

mole_CO2

Figure 4.4 CO2 solubility in aqueous DMMEA and DMMEA + PZ at


and 393) K: (
,

): 30 mass% DMMEA; at

= (313, 333, 353, 373,

= (313, 333, 373, and 393) K: (

): 25 mass% DMMEA + 5 mass% PZ

The effect of replacing equal mass% of DMMEA with PZ is similar to that for the AMP
system, i.e. the lean loadings at typical stripper conditions increase significantly, while the
rich loadings at absorber conditions are comparable. The overall effect is a reduced cyclic
capacity which suggests that a trade-off has to be made between the reaction rate and the
theoretical cyclic capacity while keeping the total amine mass concentration the same. A
more detailed discussion will be presented in the following section 4.6.
4.6 Comparisons of the CO2 solubility in amines solutions
To compare the theoretical CO2 loading capacity of different amine systems, we define the
cyclic capacity as the difference in rich loading ( = 313 K,
loading ( = 393 K,

= 3 kPa or 10 kPa) and lean

= 100 kPa) in two different ways, i.e. as a mole-ratio and as a mass-

ratio:

(4.41)

4. Experimental Results from the Static-analytical Apparatus

103

4.6.1 30 mass% MEA vs 30 mass% AMP


As mentioned before, the difference in the loading capacity of CO2 in aqueous MEA and
AMP solutions is attributed to the different reaction mechanisms. To compare these two
systems, the experimental data from our work are plotted in Figs. 4.5 and 4.6 in two different
loading units, i.e. mole-ratio loading (mole CO2 per mole amine) and mass-ratio loading
(gram CO2 per gram amine) respectively.
1000

AMP

313 K
MEA

393 K

PCO2/kPa

100

MEA
AMP
10

1
0

0.2

0.4

0.6

0.8

mole_CO2

Figure 4.5 CO2 solubility in aqueous MEA and AMP at


work, 30 mass% MEA; ( ,

= (313 and 393) K: (

): this

): this work, 30 mass% AMP; ( ):Kundu et al. [92], 30.3 mass%

AMP converted from 3.4 M AMP assuming the density of amine solution = 1
As can be seen in Fig. 4.5, the theoretical mole-ratio cyclic capacity of MEA is about 0.16 (if
rich loading is at
is 0.32 (for

= 3 kPa) or 0.2 (if rich loading is at


= 3 kPa) or 0.48 (for

= 10 kPa); while that of AMP

= 10 kPa). This shows that aqueous AMP solution

has 2-2.4 times the cyclic capacity of MEA of the same mass concentration, which means, for
the same moles of amine, AMP can carry twice the moles of CO2 as MEA. Since AMP has a
larger molecular weight than MEA, i.e. 89 for AMP compared to 61 for MEA, and
considering both amine solutions are of the same mass concentration, the cyclic capacity is
also compared according to the mass-ratio loading, i.e. g/g as in Fig. 4.6.

4. Experimental Results from the Static-analytical Apparatus

104

1000

AMP
AMP

393 K

PCO2/kPa

100

MEA
MEA
313 K
10

1
0

0.1

0.2

0.3

0.4

0.5

0.6

mass_CO2

Figure 4.6 CO2 solubility in aqueous MEA and AMP at


work, 30 mass% MEA; ( ,

= (313 and 393) K: (

): this

): this work, 30 mass% AMP; ( ):Kundu et al. [92], 30.3 mass%

AMP converted from 3.4 M AMP assuming the density of amine solution = 1
From Fig. 4.6, we know that the mass-ratio cyclic capacity for MEA is about 0.12 (if rich
loading is at

= 3 kPa) or 0.15 (if rich loading is at

= 10 kPa). Mass-ratio loading

capacity for AMP is respectively 0.16 and 0.24 for the low and high CO2 partial pressures at
rich loading condition. Although not as significant a difference as the loading capacity in
moles, the loading per unit mass capacity in AMP is still larger than that for MEA, especially
at

= 10 kPa for rich loading condition, where AMP shows a 60% larger cyclic capacity

than does MEA. The larger cyclic capacity means that less solvent is needed to absorb same
amount of CO2, which will potentially reduce the operating cost.
4.6.2 30 mass% AMP vs 25 mass% AMP + 5 mass% PZ vs 20 mass% AMP + 10 mass% PZ
In section 4.3.1, the reaction kinetics of AMP with CO2 was discussed and it was known that
they are only ~ 1/5 of those for MEA solution. This suggests that if keeping other conditions
the same, to achieve equal reaction extent, AMP requires a larger absorber than MEA which
will add to the capital cost. To make the process economic, an activator is indispensible. In
this work, PZ was added to the solution at two different concentrations, i.e. 5 mass% and 10

4. Experimental Results from the Static-analytical Apparatus

105

mass%. The total amine concentrations were chosen to be the same in this study, i.e. 30
mass%.
1000

393 K

PCO2/kPa

100

313 K
10

1
0

0.2

0.4

0.6

0.8

mole_CO2

Figure 4.7 CO2 solubility in aqueous AMP and AMP + PZ blends at

= (313 and 393) K: ( ,

): this work, 30 mass% AMP; ( ): Kundu et al. [92], 30.3 mass% AMP converted from 3.4
M AMP assuming the density of amine solution = 1; (
mass% PZ; (

): this work, 25 mass% AMP + 5

): this work, 20 mass% AMP + 10 mass% PZ

Values for cyclic capacity expressed in mole-ratio loading for the three AMP-based solutions
are listed in Table 4.10.
Table 4.10 Comparisons of mole-ratio loadings for three aqueous AMP-based solutions
30 mass% AMP
rich loading (3 kPa)
0.48
rich loading (10 kPa)
0.64
lean loading (100 kPa)
0.16
cyclic capacity (3 kPa)
0.32
cyclic capacity (10 kPa) 0.48
* Estimated from Fig. 4.7 with high uncertainty.

25 mass% AMP
+ 5 mass% PZ
0.53*
0.68
0.24
0.29*
0.44

20 mass% AMP
+ 10 mass% PZ
0.58*
0.7
0.28
0.3*
0.42

Taking 10 kPa for example, replacing 5 mass% of AMP with 5 mass% of PZ leads to the
reduction in the cyclic capacity by about 8%, compared to that of 30 mass% AMP. A further

4. Experimental Results from the Static-analytical Apparatus

106

replacement of 5 mass% reduces the capacity by another 4.5%, i.e. 12.5% smaller compared
to that of 30 mass% AMP. However, according to Samanta et al. [223], the reaction kinetics
of 20 mass% + 10 mass% PZ should be more than 5.6 times of 30 mass% AMP and is
equivalent or higher than 30 mass% MEA. Despite the sacrifice in the loading capacity
advantage of AMP, i.e. AMP + PZ blends have smaller cyclic capacity than AMP solution at
the same mass concentration, the AMP + PZ blends still benefit from larger loading capacity
than MEA solutions at the same total amine mass percentage, although the advantage is more
apparent when expressed in mole-ratio loading than mass-ratio loading.
Another way of comparing the theoretical cyclic capacity is using the mass-ratio loading. As
plotted in Fig. 4.8, the solubility CO2 in the three systems is in mass-ratio loading and the
values of cyclic capacity are summarized in Table 4.11.
1000

PCO2/kPa

100

10

1
0

0.1

0.2

0.3
mass_CO2

0.4

0.5

Figure 4.8 CO2 solubility in aqueous AMP and AMP + PZ blends at

0.6

= (313 and 393) K: ( ,

): this work, 30 mass% AMP; ( ): Kundu et al. [92], 30.3 mass% AMP converted from 3.4
M AMP assuming the density of amine solution = 1; (
mass% PZ; (

): this work, 25 mass% AMP + 5

): this work, 20 mass% AMP + 10 mass% PZ

4. Experimental Results from the Static-analytical Apparatus

107

Table 4.11 Comparisons of mass-ratio loadings for three aqueous AMP-based solutions

rich loading (3 kPa)


rich loading (10 kPa)
lean loading (100 kPa)
cyclic capacity (3 kPa)
cyclic capacity (10 kPa)

30 mass% AMP
0.24
0.32
0.08
0.16
0.24

25 mass% AMP
+ 5 mass% PZ
0.26*
0.33
0.12
0.14*
0.21

20 mass% AMP
+ 10 mass% PZ
0.28*
0.35
0.14
0.14*
0.21

* Estimated from Fig. 4.8 with high uncertainty.


The numbers in Table 4.11 show us that if the mass-ratio loading was used instead of the
mole-ratio loading to compare the cyclic capacity of these three amine systems, the PZ
activated AMP system has a CO2 absorption capacity of 12.5% smaller than the aqueous
AMP solution. There is no distinguishable difference between the two AMP+PZ blends.
4.6.3 30 mass% MEA vs 30 mass% AMP vs 30 mass% DMMEA
Apart from sterically-hindered amines, tertiary amines are also known to offer large cyclic
capacity as discussed in section 4.4.1. The tertiary amine studied in this work is DMMEA and
the data for this amine system are plotted together with AMP and MEA for comparison in Fig.
4.9.

4. Experimental Results from the Static-analytical Apparatus

108

1000

313 K

393 K
PCO2/kPa

100

10

1
0

0.2

0.4

0.6

0.8

mole_CO2

Figure 4.9 CO2 solubility in aqueous MEA, AMP and DMMEA at


): this work, 30 mass% MEA; ( ,

= (313 and 393) K: (

): this work, 30 mass% AMP; ( ): Kundu et al. [92],

30.3 mass% AMP converted from 3.4 M AMP assuming the density of amine solution = 1;
( ,

): this work, 30 mass% DMMEA

From the above figure, the mole-ratio cyclic capacity of DMMEA is approximately 0.3 (if
rich loading is at

= 3 kPa) or 0.44 (if rich loading is at

= 10 kPa). This is much

larger than that of MEA (0.16 or 0.2), i.e. the capacity of DMMEA is 1.9 or 2.2 times that of
MEA, and slightly smaller than AMP (0.32 mol/mol or 0.48 mol/mol). As DMMEA has the
same molecular weight as AMP (MW = 89), but larger than MEA (MW = 61), the cyclic
capacities of the three amines expressed in mass-ratio loading are also worth comparing and
are shown in Fig. 4.10.

4. Experimental Results from the Static-analytical Apparatus

109

1000

393 K

PCO2/kPa

100

313 K

10

1
0

0.1

0.2

0.3

0.4

0.5

0.6

mass_CO2

Figure 4.10 CO2 solubility in aqueous MEA, AMP and DMMEA at


): this work, 30 mass% MEA; ( ,

= (313 and 393) K: (

): this work, 30 mass% AMP; ( ): Kundu et al. [92],

30.3 mass% AMP converted from 3.4 M AMP assuming the density of amine solution = 1;
( ,

): this work, 30 mass% DMMEA

Values for cyclic capacity expressed in mass-ratio loading for the MEA, AMP and DMMEA
solutions are listed in Table 4.12.
Table 4.12 Comparisons of mass-ratio loadings for aqueous MEA, AMP and DMMEA
solutions

rich loading (3 kPa)


rich loading (10 kPa)
lean loading (100 kPa)
cyclic capacity (3 kPa)
cyclic capacity (10 kPa)

30 mass% MEA
0.38
0.41
0.26
0.12
0.15

30 mass% AMP
0.23
0.32
0.08
0.15
0.24

30 mass% DMMEA
0.22
0.29
0.07
0.15
0.22

As shown in Table 4.10, the cyclic capacity in terms of mass-ratio loading still follows the
sequence of AMP > DMMEA > MEA. Although slightly smaller than AMP, the values for
the cyclic capacity of DMMEA are respectively 25% and 47% larger than those of MEA at 3
kPa and 10 kPa for rich loadings.

4. Experimental Results from the Static-analytical Apparatus

110

4.6.4 30 mass% DMMEA vs 25 mass% DMMEA + 5 mass% PZ


Only one concentration of PZ activated DMMEA solution was studied in this work, i.e. 25
mass% DMMEA + 5 mass% PZ blend. The experimental data for this final amine system
studied in this work is plotted together with those of DMMEA in Fig. 4.11.
1000

393 K

PCO2/kPa

100

313 K
10

1
0.0

0.2

0.4

0.6

0.8

1.0

mole_CO2

Figure 4.11 CO2 solubility in aqueous DMMEA and DMMEA + PZ blend at


393) K: ( ,

): this work, 30 mass% DMMEA; ( ,

= (313 and

): this work, 25 mass% DMMEA + 5

mass% PZ
As can be seen from Fig. 4.11, the mole-ratio cyclic capacity of DMMEA + PZ blend is the
same as that of DMMEA at the same total amine mass percentage, i.e. 0.3 (if rich loading is
at

= 3 kPa) or 0.44 (if rich loading is at

= 10 kPa). This is a great advantage since

the replacement of DMMEA with PZ greatly enhances the kinetics of the system while the
cyclic capacity is unaffected. A further comparison of the DMMEA + PZ system with 25
mass% AMP + 5 mass% PZ suggests that the cyclic capacity of PZ activated tertiary amine is
very close to that of the promoted sterically-hinered amine. Meanwhile the mass-ratio loading
of the DMMEA based systems are also plotted in Fig. 4.12 and the comparison of the results
to the AMP based systems are presented in Table 4.13.

4. Experimental Results from the Static-analytical Apparatus

111

1000

393 K

313 K

PCO2/kPa

100

10

1
0

0.1

0.2

0.3

0.4

0.5

0.6

mass_CO2

Figure 4.12 CO2 solubility in aqueous DMMEA and AMP based systems at
K: ( ,

): this work, 25 mass% DMMEA + 5 mass% PZ; (

= (313 and 393)

): this work, 30 mass%

AMP; ( ): Kundu et al. [92], 30.3 mass% AMP converted from 3.4 M AMP assuming the
density of amine solution = 1; ( ,

): this work, 25 mass% AMP + 5 mass% PZ

Table 4.13 Comparisons of mass-ratio loadings for AMP, AMP+PZ and DMMEA+PZ
systems

rich loading (3 kPa)


rich loading (10 kPa)
lean loading (100 kPa)
cyclic capacity (3 kPa)
cyclic capacity (10 kPa)

30 mass% AMP
0.23
0.32
0.08
0.15
0.24

* Estimated from Fig. 4.12 with high uncertainty.

25 mass% AMP
+ 5 mass% PZ
0.26*
0.33
0.12
0.14*
0.21

25 mass% DMMEA
+ 5 mass% PZ
0.25
0.32
0.10
0.15
0.22

5. Thermodynamic Models

112

Chapter 5

Thermodynamic Models for the


Solubility of CO2 in Single and
Blended Amine Systems
Thermodynamic models were developed to interpret the experimental CO2 solubility data for
the amine systems studied in this work. A semi-empirical, quasi-chemical model based on the
- approach was developed for CO2 in aqueous MEA and AMP solutions. Phase equilibria
were represented using extended forms of Henrys law and Raoults law respectively for the
solute and solvent. Chemical reactions in the solution were taken into account using the mass
action law and equilibrium constants. The virial equation of state truncated after the second
term was used to account for the vapour-phase non-ideality while the Debye-Hckel theory
as modified by Guggenheim was used in the liquid phase. For the PZ-activated amine
systems, the model proposed by Kent and Eisenberg was employed owing to its simplicity.
Apart from the solubility, the quasi-chemical models developed in this work can provide
information on solution speciation. This additional information was compared with available
literature data to further validate the thermodynamics models. In the final section of this
chapter, both models were also applied to the MDEA systems based on the experimental data
from synthetic apparatus.

5. Thermodynamic Models

113

5.1 Introduction
The measuring of vapour-liquid equilibrium data is a rather laborious and time-consuming
process owing to long equilibration times and sometimes difficult sample analysis. For the
experiments in this work, it usually took one day to collect one or two equilibrium state
points. Thermodynamic models based on phase and chemical equilibria relations of all the
molecular and ionic species in the system form a good complement for the experimental
measurements. Chapter 2 summarised the different types of thermodynamic models proposed
or developed by previous researchers. In this work, we employed two different types of
quasi-chemical models to interpret experimental data. The Kent-Eisenberg model is a simple
and easy to use model in which all the liquid phase non-idealities are lumped into the userselected equilibrium constants. A more rigorous, but more complicated, model also used in
this work is based on the well-known - approach where phase non-idealities are considered
using activity and fugacity coefficients respectively for liquid and vapour phases.
This chapter begins with discussion of the - model developed for CO2 in aqueous single
amine system, i.e. MEA or AMP. Liquid phase non-ideality was accounted for using the
Debye-Hckel theory as modified by Guggenheim, while vapour phase non-ideality was
considered using the virial equation of state truncated after the second term. The experimental
data obtained from this work and a number of selected literature sources were used to regress
the binary interaction parameters in the expression for the activity coefficients. The models
were then used to predict phase speciation information at the experimental conditions and
both phase and chemical equilibria at other temperatures not studied experimentally. For the
blended amine systems including AMP + PZ and DMMEA + PZ, the Kent-Eisenberg model
was employed. Although less rigorous, the Kent-Eisenberg model has been widely used in
industry because of its simplicity. All the experimental data from single amine and blended
amine systems, e.g. AMP-H2O-CO2 and AMP-PZ-H2O-CO2, were used to derive the
adjustable parameters simultaneously. Speciation information derived from the KentEisenberg model was also included in this section. Finally, both types of models were used to
interpret the experimental data obtained from the synthetic method which will be discussed in
Chapter 6. The models applied to the amine systems in this work are summarised in Table 5.1.

5. Thermodynamic Models

114

Table 5.1 Models applied to the amine systems in this work ( = was applied in this work;
= was not applied)
MEA

AMP

AMP+PZ

DMMEA

DMMEA+PZ

MDEA

- model

Kent-Eisenberg model

5.2 - model for CO2 solubility in single amine solutions


The absorption of CO2 into an amine solution is a complex process including both phase and
chemical equilibria. Fig. 5.1 illustrates the dissolution process in the system.

Figure 5.1 Schematic diagram for the CO2 dissolution into aqueous amine system
The - model is composed of phase and chemical equilibria and equations which are used to
describe the phase non-idealities. All these facets will be discussed in the following sections.
5.2.1 Physical and chemical equilibria
As mentioned before, the absorption of CO2 into an amine solution includes both phase and
chemical equilibria. The gas phase CO2 first dissolves into the aqueous phase:
(I)
The dissolved CO2 undergoes a series of chemical reactions and forms various ionic species:
(II)

5. Thermodynamic Models

115
(III)
(IV)
(V)
(VI)

The last reaction only takes place in amines where there is a free -H to be stripped off:
normally primary and secondary amines. The equilibrium constants (

) for the above

chemical reactions are defined as:


(5.1)

(5.2)

(5.3)

(5.4)
(5.5)
Here,

represents amine,

and

are respectively molality and activity coefficients for

species in the solution, and m = 1 molkg-1. The vapour-liquid equilibria for water (solvent)
and CO2 can be described by an extended Raoults law and an extended Henrys law
respectively:

(5.6)

(5.7)

where

are, respectively, fugacity coefficient, partial pressure, and molar volume of

component , i.e. water and CO2. Superscript represents saturation.

is the gas constant.

5. Thermodynamic Models

116

is Henrys constant of CO2 in water on a molality scale and is an approximation for the
Henrys constant of CO2 in amine solution. The equilibrium constants
Henrys constant

( = 1-5) and

are temperature dependent and represented in this work by the

following empirical expression:

where

(5.8)

are constants. Values of these constants taken from the literature, are given in

Table 5.2.
Table 5.2 Equilibrium and Henrys constants used in this work for reactions I-VI (all

and

values are in mol/kg basis)


Range of
Parameter
ai
bi
ci
di
validity (K)
Source
-13445.9 -22.4773
0
140.932
273-498
Edwards et al. [134]
-12092.1 -36.7816
0
235.482
273-498
Edwards et al. [134]
-12431.7 -35.4819
0
220.067
273-498
Edwards et al. [134]
*
-17.3
0
0.05764 -38.846
293-353
Hamborg and Versteeg [260]
*
-16.651
0
0.0621 -40.718
293-353
Hamborg and Versteeg [260]
**
-1545.3
0
0
2.151
293-323
Aroua et al. [127]
-2546.2
0
0
11.555
298-313
Silkenbumer et al. [71]
-9624.4
-28.749 0.01441 192.876
273-473
Rumpf and Maurer [261]
* correlated based on eq. (5.8) from Hamborg and Versteegs experimental data
** converted from the original correlation to the form of eq. (5.8)
Vapour pressure

and partial molar volume

of water are calculated from the equations

given by Saul and Wagner [246]:


(5.9)

(5.10)
(5.11)
where
kgm-3. Parameters

= 0.01802 kgmol-1,

,
and

= 22.064 MPa,

are listed in Table 5.3

= 647.14 K, and

= 322

5. Thermodynamic Models
Table 5.3 Parameters

117

and

and partial molar volume


1
-7.8582
1.99206

used in eqs. (5.9) and (5.10) to calculate vapour pressure


of water from Saul and Wagner [246]

2
1.83991
1.10123

3
-11.781
-0.512506

4
22.6705
-16.75263

5
-15.939
-45.4485

6
1.77516
-6.75615105

.
The correlation of Brelvi and OConnell [247], equation (5.12), is used to calculate the partial
molar volume

of CO2 in water at infinite dilution (also this is an approximation for the

partial molar volume of CO2 in amine solution):

(5.12)

5.2.2 Phase non-ideality


Vapour phase non-ideality for the molecular species is represented with a truncated virial
equation of state. The fugacity coefficient is calculated according to equation (5.13):
(5.13)
where

are a pure component coefficients,

are a cross coefficients.

for CO2 and H2O

are calculated using equation (5.14) suggested by Bieling et al. [248]. Values for parameters
-

are listed in Table 5.4

(5.14)

Table 5.4 Parameters of eq. (5.14) for pure components coefficient

CO2
H2O

The cross coefficient

65.703
-53.53

-184.854
-39.29

304.16
647.3

1.4
4.3

for CO2-H2O is calculated using a correlation equation proposed by

Plyasunov and Shock [249]. Values of

are given in Table 5.5.

(5.15)

Table 5.5 Parameters for cross coefficient

used in eq. (5.15)

0
1
-211.31 -729.48

2
-1064.54

3
-656.13

5. Thermodynamic Models

118

The mixture second virial coefficient

is then calculated using the equation (5.16):

where

and

(5.16)

are the mole fractions of components and in the vapour phase.

A modified extended Debye-Hckel expression proposed by Guggenheim and Stokes [250]


for electrolyte solutions is used to calculate activity coefficients for both molecular and ionic
species in the solution phase. The first term represents the electrostatic forces and the second
term takes into account short-range Van der Waals forces:
(5.17)

where

and

are related to the dielectric constant of the solvent and system temperature and

pressure. The value of

is taken as a temperature dependent function (equation (5.18)) from

Lewis et al. [251]. Values for parameters

are summarized in Table 5.6.

is considered to

be a constant (=1.2). is the ionic strength of the solution, defined in equation (5.19).
is the ionic charge of component (for molecular species,

=0),

or

is the molality of

component .
(5.18)
(5.19)
Table 5.6 Parameters used in eq. (5.18)
0
1.313
The binary interaction parameter (

1
1.33510-3

2
1.16410-5

) of components and is assumed to be temperature

dependent:
(5.20)
and

are regressed from experimental data by optimizing the objective function

equation (5.21) [252]. The non-linear equations are solved using the numerical solvers
available in the gPROMS software package [253].

5. Thermodynamic Models

119

(5.21)

where

is the total number of measurements taken during all the experiments.

model-predicted value of solution loading.

is

is experimentally measured solution loading.

is variance of the th measurements of the solution loading. One thing worth mentioning is
that for both the - and Kent-Eisenberg models, the objective functions for the regression
can be based on either partial pressure of CO2 or solution loading. Ideally, both of these types
of objective functions should be tested for both models in order to compare which one leads
to a better result if there is any difference; however, this study only completed part of this
owing to the restrictiveness of time. In the case of the - model, the objective function was
based on solution loading, while with regards to the Kent-Eisenberg model, differences in the
experimental and modelled partial pressure of CO2 was used.
5.2.3 Balance equations
Apart from the above mentioned chemical and phase equilibria as well as phase non-idealities
equations, three balance equations also need to be included, which respectively are:
Amine balance:
(5.22)
Carbon balance:

(5.23)

Charge balance:
(5.24)
where

is the initial molality of amine,

is CO2 loading.

5.2.4 - model for MEA-H2O-CO2


There are 15 possible temperature-dependant binary interaction parameters for MEA-H2OCO2 system. According to equation (5.20), this leads to 30 adjustable parameters and would
cause over-parameterisation of the model if all were employed. After a sensitivity analysis
(step-wise regression), the number of adjustable parameters was reduced to three
temperature-dependent and four temperature-independent (Table 5.7), i.e. 10 adjustable
parameters.

5. Thermodynamic Models

120

Table 5.7 Binary interaction parameters for MEA-H2O-CO2 system


Regressed values for eq. (5.20)
(kg/mol)
(kg /(Kmol))
-0.171
2.08610-4
-1.001
3.20910-3
0.489
/
-0.202
/
-0.192
4.14010-4
-0.328
/
-0.154
/

Selected ions/molecules
interaction (kg/mol)
CO2MEA
CO2MEAH+
CO2CO32MEACO32MEAH+HCO3MEAH+CO32MEACOO-HCO3-

The correlation results are shown in Fig. 5.2. As can be seen, the model satisfactorily
represented the experimental data with

of 2.7% between the model calculation and

experimental loadings. Fig. 5.3 shows the parity plot between the experimental and model
data.

100000
10000
1000

393 K

PCO2 / kPa

100
10

313 K

1
0.1
0.01
0.001
0

0.2

0.4

CO2

0.6

0.8

Figure 5.2 Correlation of experimental results from this work and Jou et al. [73] at 313 K ( );
this work, Jou et al., Mamun et al. [70], Goldman and Leibush [201] at 393 K (
smoothed lines: model calculation

);

5. Thermodynamic Models

121

0.8

1.2

0.7

1
Model loading

Model loading

0.6
0.5
0.4
0.3

0.8
0.6
0.4

0.2
0.2

0.1
0

0
0

0.2

0.4

0.6

0.8

Experimental loading

(a)

0.2

0.4

0.6

0.8

1.2

Experimental loading

(b)

Figure 5.3 Parity plot between experimental and model loadings: (a) Experimental data from
this work vs. model results; (b) Experimental data from literature vs. model results. ( ): 313
K; ( ): 393 K
0.1

MEA/MEAH+

0.01

MEACOO-

0.001

HCO3-/CO32-

0.0001

CO2

1E-05
0

0.2

0.4

0.6
CO2

(a)

0.8

5. Thermodynamic Models

122

0.1

MEA

MEAH+

0.01

0.001

HCO3CO32-

0.0001

1E-05
0

0.2

0.4

0.6

0.8

CO2

(b)
Figure 5.4 Liquid phase speciation in 30 mass% MEA at 313 K. (a): Poplsteinova et al.,
[254]: MEACOO-: ( ); MEA/MEAH+: ( ); HCO3-/CO32-: ( ); CO2: ( ). Bttinger et al.,
[255]: MEACOO-: (
MEACOO-: (

); MEA/MEAH+: (

); HCO3-/CO32-: ( ). Hilliard, [256]:

); MEA/MEAH+: ( ); HCO3-/CO32-: ( ). Lines: model predictions. (b):

Poplsteinova et al.: MEA: ( ); MEAH+: ( ); CO32-: ( ); HCO3-: ( ). Lines: model


predictions
The speciation results from the model at 313 K for 30 mass% MEA are compared with the
literature NMR data from Poplsteinova et al. [254], Bttinger et al. [255] and Hilliard [256]
in Fig. 5.4 (a) for MEACOO-, CO2, MEA/MEAH+ and HCO3-/CO32-; and (b) for MEA,
MEAH+, HCO3- and CO32-. The model predictions agree well with the experimental results in
Fig. 5.4 (a), generally within the experimental uncertainties. Owing to the difficulties in
distinguishing between MEA and MEAH+, HCO3- and CO32-, only Poplsteinova reported
individual ion mole fractions for MEA/MEAH+ and HCO3-/CO32- as shown in Fig. 5.4 (b).
The most noticeable differences are predictions for high loadings of CO32- and MEA. As
explained in their studies, if the species fraction is < 10%, the concentration evaluation of
species with common peaks is not reliable. At loadings above 0.6 the concentrations of

5. Thermodynamic Models

123

carbonate and MEA are too low to allow accurate measurements; therefore, the model
predictions may be more reliable.
To further check the quality of the model, the interaction parameters obtained from the above
correlation were applied to predict CO2 solubility in 30 mass% aqueous MEA at

= (298 and

333) K. The comparison of the predictions with Jous experimental results is shown in Fig.
5.5.

values are 5.9% and 6.7% for 298 K and 333 K respectively.

1000

10000

100

1000
100

PCO2 / kPa

PCO2 / kPa

10

298 K

0.1

10

333 K

1
0.1

0.01

0.01
0.001

0.001
0

0.2

0.4

0.6

0.8

0.2

CO2

(a)

0.4

0.6

0.8

CO2

(b)

Figure 5.5 Predictions from this work compared with Jou et al.s experimental results at (a):
298 K (

); and (b): 333 K ( ); smoothed lines: model calculation

5.2.5 - model for AMP-H2O-CO2


The same model framework has also been applied to the AMP-H2O-CO2 system. The binary
interaction parameters (Table 5.8) were regressed against the experimental results obtained in
this work from

= (313 to 393) K. The correlation results are shown in Fig. 5.6 with

of

1.9%. Besides, a parity plot as shown in Fig. 5.7 shows that the data are well distributed by
the model.

5. Thermodynamic Models

124

Table 5.8 Binary interaction parameters for the AMP-H2O-CO2 system


Regressed values for eq. (20)
(kg/mol)
(kg /(Kmol))
-13.961
-4.87810-2
-0.833
1.20610-3
-2.407
7.18710-3
0.488
-7.18710-4
-7.214
0.022
0.940
-2.58910-3
2.341
-6.21310-3

Selected ions/molecules
interaction (kg/mol)
CO2AMPH+
HCO3-CO32HCO3-AMP
HCO3-AMPH+
HCO3-AMPCOOCO32-AMPH+
AMPAMPH+

1000

393 K

PCO2 / kPa

100

353 K

333 K

10

313 K

373 K
1
0

0.2

0.4

0.6

0.8

CO2

Figure 5.6 Correlation of experimental results for AMP-H2O-CO2 system from this work at
313 K (

), 333 K ( ), 353 K ( ), 373 K ( ), and 393 K ( ); smoothed lines: model

calculation

5. Thermodynamic Models

125

1
0.9
0.8

Model loading

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.2

0.4

0.6

0.8

Experimental loading

Figure 5.7 Parity plot between experimental and model loadings. ( ): 313 K; ( ): 333 K; (
): 353 K; ( ): 373 K; (

): 393 K

0.1

AMPH+

AMP
0.01

HCO3-

0.001

CO32-

AMPCOO-

0.0001

1E-05
0

0.2

0.4

0.6

0.8

CO2

Figure 5.8 Liquid phase speciation in 30 mass% AMP at 373 K from model predictions.
AMPCOO-: (

); AMPH+: (

); AMP: (

); HCO3-: (

); CO32-: (

5. Thermodynamic Models

126

Speciation for 30 mass% AMP at 373 K from model predictions is presented in Fig. 5.8. As
loading increases, the concentration of AMPH+ also increases owing to the reduction in the
solution pH. Simultaneously, the concentration of AMP decreases with increase in loading.
Most of the dissolved CO2 exists in the form of HCO3-, with only a small fraction of CO32and AMPCOO-.
5.3 Kent-Eisenberg model for AMP-based and DMMEA-based blended amine systems
It is much more complicated to model CO2 in aqueous AMP + PZ and DMMEA + PZ blends
than in single amine solutions (AMP or DMMEA), since PZ, as a diamine, can react with
CO2 to form various products, including first and second order carbamates, protonated PZ
and carbamate. This means that if the above-described - model were used and all these
chemical reactions were taken into account alongside the ones laid out in section 5.2.1
(reactions I - VI), the model would be numerically too difficult to be solved. For a suitable
VLE model to be incorporated into a process model, it should be as accurate as necessary
(compared to the experimental data) but also with reasonable computational time and
difficulty. Considering this, Kent-Eisenberg model is a good compromise owing to its
simplicity and reasonable accuracy in correlating experimental data. Since its introduction by
Kent and Eisenberg in 1974, it has been widely used in industry and academic research,
although the actual form has been through several modifications [149]. The Kent-Eisenberg
model developed in this work will be discussed in the following sections.
5.3.1 Physical and chemical equilibria
The physical dissolution and most of the chemical equilibration reactions are the same as
described in section 5.2.1, except those involving PZ. As a diamine, PZ undergoes reactions
with H2O and CO2 as shown below:
Deprotonation of PZ:
(VII)
First order carbamate deformation:
(VIII)
Second order carbamate deformation:

5. Thermodynamic Models

127
(IX)

Deprotoation of protonated first order carbamate:


(X)
For the above reactions, the equilibrium constants in the Kent-Eisenberg model are functions
of apparent molality instead of activity as in the - model.
(5.25)
(5.26)
(5.27)

(5.28)
With regard to AMP and DMMEA, the equilibrium constants of reactions II V in section
5.2.1 are also related to the apparent concentrations of the species in the solution:
(5.29)
(5.30)

(5.31)

(5.32)
From Chapter 4, we know that chemical reaction VI (carbamate formation) is present in the
solution of AMP; however, the equilibrium concentration of AMP carbamate was extremely
small as shown in Fig. 5.8. In the mixture with PZ, competing reactions would further reduce
the concentration of AMP carbamate. For simplification, it was neglected for the KentEisenberg model in this work. For DMMEA, since it is a tertiary amine with no free hydrogen available for deprotonation, it does not react with CO2 directly to form carbamate.
The equilibrium constants for equations. (5.29) (5.32) were taken from various literature

5. Thermodynamic Models

128

sources and were summarised in Table 5.9 (the ones kept the same as in Table 5.2 were not
included here).
Table 5.9 Equilibrium constants used in the Kent-Eisenberg model for reaction VII-X, and V
for DMMEA (all

Parameter

values are in mol/kg basis)

ai
3814.4

bi
0

ci
-1.5096

di
14.119

Range of
validity (K)
273-323

3616

-8.635

283-333

1322.1

-3.654

283-333

-6066.9

-2.29

0.0036

6.822

273-343

-2546.2

11.555

298-313

0.04388

-34.306

293-353

Source
Kamps et al. [257]
Ermatchkov et al.
[258]
Ermatchkov et al.
[258]
Cullinane and
Rochelle [259]
Silkenbumer et al.
[71]
Hamborg and
Versteeg [260]

5.3.2 Phase non-ideality


The treatment of phase non-idealities in the Kent-Eisenberg model was significantly
simplified as compared to the - approach, i.e. the vapour phase non-ideality was neglected
while all the non-ideality in the liquid phase was lumped into a number of selected
equilibrium constants.
In this work, we chose

as adjustable equilibrium constants based on the criteria of

relative importance. Considering that PZ readily reacts with CO2 to form carbamate, whereas
it has a smaller pKa than AMP in the blended system, reactions VIII (formation of first order
carbamate) and IX (formation of second order carbamate) were selected for PZ. For AMP
and DMMEA, as the carbamate formation was neglected, the protonation of these two amines
was used to represent the non-ideality.
The original form of the Kent-Eisenberg model assumed that the adjustable equilibrium
constants were merely functions of temperature. Jou et al. [119] revised the model by
introducing dependency on loading and amine concentration; however, the exact form was
not described in their paper. In the work of Hu and Chakma [147], the equilibrium constants
were considered to be dependent on temperature, acid gas partial pressure and amine
concentration. Li and Shen [148] modified the Kent-Eisenberg model to be dependent on

5. Thermodynamic Models

129

temperature, amine concentration and CO2 loading. The final form of the equilibrium
constants was as follows:
(5.33)
,

are adjustable parameters regressed from experimental data. In this work, we assumed

the equilibrium constants are a function of temperature, CO2 partial pressure and amine
concentration. The adjustable equilibrium constants,

, are related to the literature values

( , calculated from eq. (5.8) and parameters in Table 5.8) with deviation parameters (

):

(5.34)

(5.35)
,

and

are adjustable parameters obtained from regression of experimental data for both

ternary (AMP-H2O-CO2 or DMMEA-H2O-CO2) and quaternary systems (AMP-PZ-H2O-CO2


or DMMEA-PZ-H2O-CO2) simultaneously. The objective function selected for the KentEisenberg model is different from that for the - model: the former was minimising the
difference in partial pressure of CO2 while the latter was based on the solution loading.

(5.36)

where

is the total number of measurements taken during all the experiments.

predicted value of CO2 partial pressure.

is model-

is experimentally measured solution loading.

is variance of the th measurement of the system pressure. This gives different weighting
to the measurements according to the quality of the data points.
5.3.3 Balance equations
The mass and charge balance equations also need to be observed and are expressed as follows:
Amine balance:
(AMP)

(5.37)

(DMMEA)

(5.38)

5. Thermodynamic Models

130

Carbon balance:

(AMP) (5.39)
(DMMEA) (5.40)

Charge balance:
(AMP)
(5.41)
(DMMEA) (5.42)
5.3.4 Kent-Eisenberg model for AMP-PZ-H2O-CO2
As mentioned above,

are selected to represent the phase nonideality and

according to equation (5.35), the total adjustable parameters amounted to as many as 21.
Some of the parameters are superfluous; as a result, a sensitivity analysis was conducted to
reduce the quantity of parameters. The final numbers of adjustable parameters has been
reduced to 15 as listed in Table 5.10. This is acceptable, especially considering the
complicated solution chemistry when PZ is introduced in the mixture.
Table 5.10 Adjustable parameters in eq. (5.35) for the AMP-PZ-H2O-CO2 system
Regressed adjustable parameters
0.3293
/
0.02139
/
-0.005099
/
-0.8711

20.30
/
-0.07474
/
0.0007053
/
-0.01358

-0.3386
504.5
-0.002334
0.05995
-0.001382
-0.3419
-0.1997

The correlation results are illustrated in Figs. 5.9 and 5.10 for all the single and blended
amine systems. The combined

between the model prediction of CO2 partial pressure and

experimental data is 12.6%. For AMP-PZ-H2O-CO2, since there was no literature data
conducted at the same conditions as in this work, only our data were included. The

for

the 25 mass% AMP+5 mass% PZ and 20 mass% AMP+10 mass% PZ are 11.1% and 8.4%
respectively; for AMP-H2O-CO2 system, literature data from Kundu et al. [92] at 313 K, and
Li and Chang [72] at 333 K and 353 K were also incorporated in the data regression,
alongside the data from this work. The
work and 20.5% for all the literature sources.

for AMP-H2O-CO2 system is 15.5% for this

5. Thermodynamic Models

131

1000

1000

373 K

393 K

373 K

393 K

PCO2 /kPa

100

PCO2 /kPa

100

10

10

333 K

313 K

333 K

313 K

1
0

0.5
CO2

(a)

0.5
CO2

(b)

Figure 5.9 Correlation of experimental results for AMP-PZ-H2O-CO2 system from this work
at 313 K (

), 333 K (

), 373 K ( ), and 393 K ( ); smoothed lines: model calculation. (a):

25 mass% AMP+5 mass% PZ; (b): 20 mass% AMP+10 mass% PZ


1000

373 K

393 K

353 K

PCO2 / kPa

100

10

313 K
333 K
1
0

0.2

0.4

0.6

0.8

CO2

Figure 5.10 Correlation of experimental results for AMP-H2O-CO2 system from this work at
313 K (
(

), 333 K ( ), 353 K ( ), 373 K ( ), and 393 K ( ); Kundu et al. [92] at 313 K

); Li and Chang [72] at 333 K (

) and 353 K (

); smoothed lines: model calculation

5. Thermodynamic Models

132

The parity plot between the experimental data from this work and the model data is shown in
Fig. 5.11.
1000
30% AMP_313 K
30% AMP_333 K
30% AMP_353 K
30% AMP_373 K
30% AMP_393 K
Model CO2 partial pressure

25% AMP+5% PZ_313 K


100

25% AMP+5% PZ_333 K


25% AMP+5% PZ_373 K
25% AMP+5% PZ_393 K
20% AMP+10% PZ_313 K
20% AMP+10% PZ_333 K
20% AMP+10% PZ_373 K
20% AMP+10% PZ_393 K

10

1
1

10

100

1000

Experimental CO2 partial pressure

Figure 5.11 Parity plot between the experimental data from this work and the model data for
30 mass% AMP, 25 mass% AMP + 5 mass% PZ, and 20 mass% AMP + 10 mass% PZ
systems
Figs. 5.12 and 5.13 illustrate the speciation predictions of AMP-PZ-H2O-CO2 systems at 313
K from the Kent-Eisenberg model.

5. Thermodynamic Models

133

0.07
0.06

AMPH+

AMP
0.05

HCO3x

0.04
0.03

PZ

0.02

PZCOOCO32-

PZCOO-H+

PZ(COO-)2

0.01

CO2
0
0

0.2

0.4

0.6

0.8

CO2

Figure 5.12 Liquid phase speciation in aqueous 25 mass% AMP + 5 mass% PZ solution at
313 K from model predictions
0.06

0.05

AMPH+
AMP

0.04

HCO3PZ

0.03

PZCOO-H+
0.02

PZ(COO-)2
0.01

CO2

CO32-

PZCOO
0
0

0.2

0.4

0.6

0.8

CO2

Figure 5.13 Liquid phase speciation in aqueous 20 mass% AMP + 10 mass% PZ solution at
313 K from model predictions

5. Thermodynamic Models

134

From the above graphs, it is apparent that the concentrations of molecular amines, i.e. AMP
and PZ, reduce as the solution loading increases. In the mean time, the concentrations of
protonated AMP, protonated piperazine carbamate and bicarbonate rise with the increase in
the CO2 loading in the range of our model prediction. The concentrations of the second order
piperazine carbamate and carbonate peak at just after

= 0.6 and at

< 0.1 respectively.

This is reasonable, as the increase in the CO2 loading leads to rise in the solution pH which
results in the prevalence of protonated forms of all the species. The concentration of the first
order piperazine carbamate is minimal compared to other species in the solution. To
summarize, the key conclusion is that PZ exists mainly in the form of PZCOO -H+ and
PZ(COO-)2, whereas PZCOO- is of less importance. Besides, since the formation of AMP
carbamate was neglected in our Kent-Eisenberg model, AMPH+ is the only viable specie in
the CO2 loaded AMP solution. Finally, the most significant species derived from CO2 are
HCO3-, PZ(COO-)2, PZCOO-H+.
Fig. 5.14 shows the liquid phase speciation of 30 mass% AMP at 313 K from the KentEisenberg model. To compare with the speciation in AMP-PZ-H2O-CO2 (Figs. 5.12 and 5.13)
and - model predictions (Fig. 5.8), the results are plotted with a linear (left) and a
logarithmic scale (right).
0.08

0.1

0.07

AMPH+

AMP

0.01

0.06

CO320.001

HCO3-

0.04

CO2

0.02

CO3

0.01

CO2

1E-05

2-

1E-06

0
0

(a)

HCO3-

0.0001

0.03

AMP

AMPH+

0.05

0.2

0.4

CO2

0.6

0.8

0.2

0.4

0.6

0.8

CO2

(b)

Figure 5.14 Liquid phase speciation in aqueous 30 mass% AMP solution at 313 K from
model predictions. (a) normal scale for y-axis; (b) logarithmic scale for y-axis

5. Thermodynamic Models

135

By comparing Fig. 5.14 a) with Figs. 5.12 and 5.13, it can be observed that the general trends
of the major species are similar to those of mixed amine systems. One noticeable difference is
that the concentration profile of HCO3- in the mixed amines has a smaller slope at low CO2
loadings followed by a steeper slope when the loading exceeds 0.6. Comparatively, the
HCO3- concentration in AMP-only solution has a quasi- linear profile. This can be explained
by the presence of PZ, which preferably reacts with CO2 at lean loadings and leads to the
retaining of AMP in the solution. Fig 5.14 b) is not directly comparable with Fig. 5.8 owing
to the difference in the model assumptions (whether or not there is a formation of carbamate)
and temperature conditions presented here (373 K vs 313 K). However, it is still apparent that
the common species in these two graphs share similar concentration profiles.
5.3.5 Kent-Eisenberg model for DMMEA-PZ-H2O-CO2
As a tertiary amine, the formation of DMMEA carbamate is not feasible. Therefore, the
reactions in DMMEA-PZ-H2O-CO2 are similar to those described above for AMP-PZ-H2OCO2. Similarly,

are chosen to incorporate phase nonideality. The total number of

adjustable parameters (Table 5.11) has also been reduced from 21 down to 15 after a
sensitivity analysis.
Table 5.11 Adjustable parameters in eq. (5.35) for the DMMEA-PZ-H2O-CO2 system
Regressed adjustable parameters
-12.20
/
0.1220
/
-0.002360
/
-22.25

46.61
/
-0.1645
/
0.002513
/
-0.001098

-4.224
2958
-0.03982
0.07063
-0.002274
-4.002
-4.133

The correlation results are illustrated in Figs. 5.15 and 5.16. The combined

for all the

experimental points is 11.7%, including both DMMEA-only and DMMEA+PZ systems,


while the respective

for these solutions is 13.6% and 7.8%. The blended amines are

better correlated than the single amine case because more adjustable parameters are relevant
to the former than the latter.

5. Thermodynamic Models

136

1000

373 K

393 K

PCO2 / kPa

100

10

333 K

313 K
1
0.0

0.2

0.4

0.6

0.8

1.0

CO2

Figure 5.15 Correlation of experimental results for DMMEA-PZ-H2O-CO2 system from this
work at 313 K (

), 333 K (

), 373 K ( ), and 393 K ( ); smoothed lines: model

calculation
1000

393 K

373 K

PCO2 / kPa

100

353 K
10

333 K
313 K
1
0.0

0.2

0.4

0.6

0.8

1.0

CO2

Figure 5.16 Correlation of experimental results for DMMEA-H2O-CO2 system from this
work at 313 K (
model calculation

), 333 K (

), 353 K ( ), 373 K ( ), and 393 K ( ); smoothed lines:

5. Thermodynamic Models

137

1000
25% DMMEA+5% PZ_313 K
25% DMMEA+5% PZ_333 K

Model CO2 partial pressure

25% DMMEA+5% PZ_373 K


25% DMMEA+5% PZ_393 K
100
30% DMMEA_313 K
30% DMMEA_333 K
30% DMMEA_353 K
30% DMMEA_373 K
10

30% DMMEA_393 K

1
1

10

100

1000

Experimental CO2 partial pressure

Figure 5.17 Parity plot between the experimental data from this work and the model data for
30 mass% DMMEA, and 25 mass% DMMEA + 5 mass% PZ systems
Liquid speciation results are shown in Figs. 5.18 and 5.19, respectively, for the DMMEA-PZH2O-CO2 and DMMEA-H2O-CO2. Comparing Fig. 5.18 and Fig. 5.12, it can be noticed that
the general trends of the species in the solutions are largely similar. The only apparent
difference appears at near loading of 0.8, where there are larger concentration of PZ(COO -)2
and smaller concentration of PZCOO-H+ in DMMEA+PZ than AMP+PZ. This may largely
be influenced by the solution pH and it can be inferred that at the same solution loading, the
DMMEA+PZ mixture has a greater solution pH than the AMP+PZ.
Regarding the single amines, i.e. AMP-H2O-CO2 (Fig. 5.14 a)) and DMMEA-H2O-CO2 (Fig.
5.19), the concentration profiles have no discernable difference. This is also reflected by the
similarity in the solubility profile of these two amines as shown in Chapter 4 Fig. 4.9.

5. Thermodynamic Models

138

0.07
0.06

DMMEAH+

DMMEA

0.05

HCO3-

0.04
0.03

PZ

0.02

PZCOO-H+
CO32-

CO2

PZ(COO-)2

0.01

PZCOO0
0

0.2

0.4

0.6

0.8

CO2

Figure 5.18 Liquid phase speciation in aqueous 30 mass% DMMEA + PZ solution at 313 K
from model predictions
0.08

DMMEA

0.07

DMMEAH+

0.06

HCO3-

0.05
0.04
0.03
0.02

CO32-

0.01

CO2

0.00
0

0.2

0.4

0.6

0.8

CO2

Figure 5.19 Liquid phase speciation in aqueous 30 mass% DMMEA solution at 313 K from
model predictions

5. Thermodynamic Models

139

5.4 Thermodynamic models for CO2 solubility in aqueous MDEA solutions


MDEA is a typical tertiary amine with no direct reaction with CO2 to form carbamate. The
reactions in the solution are similar to those for DMMEA presented in section 5.2.1, II-V.
Table 5.12 summarizes the constants used in equation (5.8) for the MDEA protonation
reaction (V).
Table 5.12 MDEA protonation constant described in eq. (5.8)

Parameter

ai

bi

ci

di

Range of
validity (K)

0.041447

32.2592

Source
Kuranov et al.
[139]

Both the Kent-Eisenberg model and - approach have been applied to the MDEA-H2O-CO2
system based on the data from this work (Chapter 6) and Sidi-Boumedine et al. [74]. The
procedures are the same as described in the previous sections and will not be described here
again.
5.4.1 Kent-Eisenberg model for MDEA-H2O-CO2
As DMMEA, the liquid phase nonideality is lumped into the equilibrium constant of MDEA
protonation, shown in equation (5.43). The regressed values of the parameters are shown in
Table 5.13.
(5.43)
Table 5.13 Regressed parameters in eq. (5.43)
Regressed adjustable parameters
0.08511
424.1
-0.006372
0.005794
-0.3228
-0.001275
0.01676
0.04441

Figs. 5.20 and 5.21 show the correlation of the experimental results from Sidi-Boumedine et
al. and this work respectively. It can be seen that the correlation is generally satisfactory. The

5. Thermodynamic Models

140

most noticeable discrepancy between the correlation and experimental data from this work as
shown in Fig. 5.21 is at the range of 0.8-1.0 solution loading. This may be attributed to the
fact that the vapour phase non-ideality was neglected in the Kent-Eisenberg model, while at
high solution loading, it become an important influencing factor. The

for the

experimental and correlation CO2 partial pressure data is 18.4% (or 6.8% on the loading
basis).
10000

1000

348 K
PCO2/kPa

100

313 K
10

298 K
1

0.1
0

0.2

0.4

0.6

0.8

1.2

CO2

Figure 5.20 Correlation of experimental results for MDEA-H2O-CO2 system from SidiBoumedine et al. [74] at 298 K ( ), 313 K ( ), 348 K (
calculation

); smoothed lines: model

5. Thermodynamic Models

141

1000

PCO2/kPa

100

316 K
10

1
0.0

0.2

0.4

0.6

0.8

1.0

1.2

CO2

Figure 5.21 Correlation of experimental results for MDEA-H2O-CO2 system from this work
at 316 K
The speciation in the solution is also predicted from the Kent-Eisenberg model as illustrated
in Fig 5.22. The predominant species are MDEA, MDEAH+ and HCO3-, whereas CO32- and
CO2 exist in much smaller concentrations. As expected, the concentrations of MDEAH+ and
HCO3- increase with the rise in solution loading. On the contrary, molecular MDEA is
gradually consumed by the CO2 introduced in the system. When comparing Fig. 5.22 with
Fig. 5.19, which shows the speciation in DMMEA solution, it can be noticed that the
concentration of CO32- in MDEA is generally smaller than that in DMMEA. This can be
easily explained by the difference in solution pH - 30 mass% DMMEA has a larger pH than
25 mass% MDEA and preferred the formation of bicarbonate than carbonate at the same
solution loading.

5. Thermodynamic Models

142

0.07

MDEA

0.06

MDEAH+

0.05
0.04

HCO3-

0.03
0.02

CO2
0.01

CO320
0

0.2

0.4

0.6

0.8

CO2

Figure 5.22 Liquid phase speciation in aqueous 25 mass% MDEA solution at 313 K from
model predictions
5.4.2 - model for MDEA-H2O-CO2
A more complicated method is using the rigorous - approach to account for the liquid and
vapour phase non-ideality respectively. The detailed procedures are already discussed in
section 5.2 and will not be reiterated here. There were 10 possible temperature-dependent
adjustable parameters for MDEA-H2O-CO2 system which have been reduced to 7 after a
sensitivity analysis. Table 5.14 included the final regression results for these interaction
parameters.
Table 5.14 Binary interaction parameters for MDEA-H2O-CO2 system in the - model
Selected ions/molecules
interaction (kg/mol)
CO2HCO3CO2MDEAH+
HCO3--CO32HCO3--MDEAH+
CO32-MDEA
CO32-MDEAH+
MDEAMDEAH+

Regressed values for eq. (24)


(kg/mol)
(kg /(Kmol))
-2.244
-1.22510-2
0.7682
1.71910-2
-5.523
1.87310-2
-2
-8.78310
4.77310-4
-8.690
4.29810-2
-1.827
6.12110-3
0.2765
-3.08610-4

5. Thermodynamic Models

143

The correlations results from the - model are plotted in Figs. 5.23 and 5.24. Comparing
Figs. From 5.20 to 5.24, it can be seen that the - model generally correlates better than the
Kent-Eisenberg model. Part of the explanation can be attributed to the fact that - model is
thermodynamically more rigorous. However, it also contains considerably more adjustable
parameters, i.e. 14 compared to 8 in the Kent-Eisenberg model, it is therefore clear that it
would be expected to correlate the experimental data better. The

for the experimental

and correlation solution loading data is 5.8%. One thing to note is that although the
value for the Kent-Eisenberg model (18.4%) is substantially larger than that for the - model
(5.8%), it does not fully reflect the quality of the correlation, as the optimisation function of
the former was based on CO2 partial pressure compared to the solution loading for the latter.
This also shows that a larger value of
equivalent to a smaller value of

based on the optimisation of system pressure is

dependent on the solution loading optimisation.

10000

1000

100

PCO2/kPa

348 K
313 K

10

298 K

0.1
0

0.2

0.4

0.6

0.8

1.2

CO2

Figure 5.23 Correlation of experimental results for MDEA-H2O-CO2 system from SidiBoumedine et al. at 298 K ( ), 313 K ( ), 348 K (

); smoothed lines: model calculation

5. Thermodynamic Models

144

1000

PCO2/kPa

100

316 K
10

1
0.0

0.2

0.4

0.6

0.8

1.0

1.2

CO2

Figure 5.24 Correlation of experimental results for MDEA-H2O-CO2 system from this work
at 316 K
1
0.9
0.8

Model loading

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.0

0.2

0.4

0.6

0.8

1.0

Experimental loading

Figure 5.25 Parity plot between the experimental data from this work and the model data for
30 mass% MDEA systems (solid points: this work; hollow points: Sidi-Boumedine et al.)

6. Solubility Measurements with the Static-synthetic Apparatus

145

Chapter 6

Solubility Measurements for H2OCO2 and MDEA-H2O-CO2


Systems with Synthetic Apparatus
In chapter 3, we discussed the solubility measurements for three single-amine and two
blended-amine systems with a static-analytical apparatus. In this chapter, another type of
experimental apparatus is presented. This equipment, based on the principle commonly
known as the synthetic method, is a good complement to the static-analytic one, especially
for the high pressure ranges and certain types of amines which cannot be easily quantified
with gas chromatography owing to their high boiling point, such as MDEA.

6. Solubility Measurements with the Static-synthetic Apparatus

146

6.1 Introduction
As mentioned in Chapter 2, one of the major alternatives to the static-analytical method is the
synthetic method. To compare the two methods, experiments were conducted with a highpressure synthetic apparatus. At the time the equipment was built, it was novel both in its
design and experimental procedures. Therefore, there was a necessity to validate the
apparatus by measuring a well-studied system, like e.g. H2O-CO2, before applying it to more
complicated amine-H2O-CO2 systems. Since we started our measurements with H2O-CO2
system, it is worthwhile to draw attention to the previous data obtained for this system and
empirical models discussed in the literature before going into experimental details.
Solubility measurements for CO2 in water have been extensively studied owing to its wide
applications in physical, chemical, and biological science [17]. The earliest investigation
dates back to 1803, when Henry postulated his limiting law of solubility. However, it was not
until the middle of the 19th century that the accurate solubility data started to emerge, with the
pioneering work of Bunsen [262].
A number of literature reviews and data compilations on the mutual solubility of the CO2H2O system were published from the 1940s. Among them, the most noteworthy ones were
overviews carried out by Carroll et al. [17] and Crovetto and Mather [18]. The former
covered the pressure range below 1 MPa and temperature range from 273 K to 433 K, while
the latter included elevated pressures and temperature range between 273 and the critical
temperature of the solvent at 647 K. Carroll et al. derived Henrys constant from the raw
experimental data, i.e. total pressures and liquid phase mole fractions at given temperatures,
based on the equality of fugacities for both components in both phases. For water, regarded
as the solvent under the conditions within their study scope, the extended version of Raoults
law was given by:
(6.1)
The extended version of Henrys law was used to describe the solute CO2:
(6.2)
In eqs. (6.1) and (6.2),

stands for the mole fraction of CO2 in the liquid phase and

accordingly for the mole fraction in the vapour phase.

and

represent the vapour pressure

above the saturated pure water and the total pressure in the system, respectively. In eq. (6.1),

6. Solubility Measurements with the Static-synthetic Apparatus

147

stands for the fugacity coefficient of water in the vapour phase, while
represents the fugacity coefficient of the pure solvent under saturation
conditions.

represents the activity coefficient of water in the liquid phase,

normalized according to the convention of Raoults law and referring to the mole fraction as
the composition variable.

stands for the molar volume of pure liquid water.

and

are respectively the fugacity

coefficient of CO2 in the vapour phase, the activity coefficient of CO2 in the liquid phase (its
composition dependence being expressed in terms of the mole fraction and normalized
according to Henrys law) and the partial molar volume at infinite dilution of CO 2 in the
liquid phase.

is the Henrys law constant (referring to the mole fraction as concentration

unit) in the limit of infinite dilution in which the liquid phase stands under the vapour
pressure of the pure solvent.
In the work of Carroll et al. [17], good approximation valid for the low pressure region
(i.e.

) was obtained by imposing simplifying assumptions on eqs. (6.1) and (6.2),

including: neglecting ionization in the liquid phase [17]; assuming that for both components,
the Poynting correction, accounted for with the terms

and

respectively, was negligible and the activity coefficients were unity. The

following expressions for the solvent component were derived [17]:


(6.3)
(6.4)
To solve the above equations, the vapour pressure of water was taken from Haar et al. [263],
while the fugacity coefficients for the components in the gaseous phase were calculated by
means of the Redlich-Kwong equation of state [264]. For each experimental point ( , ,
the vapour phase composition

as well as the Henry constant

were calculated from the

set of eqs. (6.3) and (6.4) in conjunction with Redlich-Kwong equation of state through an
iterative procedure.
Henrys constants were taken from the work of Carroll et al. [17] and were correlated with
temperature over the interval 273 K <

< 433 K:
(6.5)

6. Solubility Measurements with the Static-synthetic Apparatus

148

A step-wise procedure was adopted by Carroll et al. in performing the least square analysis,
in which any values from a given author deviating by more than 5% from the regression were
rejected (except for the data of Novk et al. [265] where four data points were found to fall
out of the tolerable range). Subsequently another regression was executed until all the data
fitted the criterion. The values for the optimal set of coefficients were found to be:
,

and

Compared with the study of Carroll et al., broader ,


work of Crovetto which covered the range from
calculate

conditions were considered in the

= 273 K to the critical point of water. To

from experimental ( , , ) information, she employed a similar framework

based on eqs. (6.1) and (6.2). Likewise, the activity coefficients were set equal to unity
and acid base reactions leading to the formation of new species such as H2CO3
(aq), HCO3- (aq) and CO32- (aq) were not accounted for since the sum of mole fractions of
these species was small compared to the mole fraction of CO2 (aq) and therefore could be
ignored in the temperature range of interest [266].To account for the non-ideality of the
gaseous phase,

and

were estimated by means of the virial equation of state (truncated

after the second virial coefficients) over the temperature interval 273 K< /K < 353 K, while
the Peng-Robinson equation of state was used to determine the fugacity coefficients in the
range of 353 < /K <

. In contrast to the work of Carroll et al., in which the Poynting

correction was ignored, Crovetto accounted for the pressure dependence of the Henrys
constant through experimental data for
of water such as the density

[18]. Wherever necessary, pure component data

, the reference fugacity

and

were taken from Haar et al.

[263].
Crovetto divided the literature data into three groups according to their pressure and
temperature range. Group A included solubility measurements from (273 to 353) K,

< 0.2

MPa; Group B covered the same temperature range as group A, but at the higher end of
pressure, i.e.

> 0.2 MPa; whereas group C contained all the remaining data, from 353 K to

the critical temperature of the solvent at any pressure. Only results from sources compiled in
the first and the third groups were used for the calculations of
(6.1) and (6.2) for

and

. By solving the set of eqs.

, and then fitting the resulting pairs of data ( ,

empirical equation for describing

) to an

over the lower temperature and pressure range,

273 < /K < 353 and 0 < /MPa < 0.2 of the form:

6. Solubility Measurements with the Static-synthetic Apparatus

149
(6.6)

the following set of coefficients was obtained by applying least squares analysis:
and

The corresponding standard deviation of the fitting

equation was found to be 1.1%

. For the data set in the range 353 < /K <

and

at any pressure, the fitting was completed and anchored with five low-temperature points
calculated from eq. (6.6) at
range 353 < /K <

= (273.15, 293.15, 303.15, and 323.15) K. The data set in the


and at any pressure

was first correlated according to eq. (6.6).

Then an additional asymptotic term given by the expression

was

adjusted based on the results obtained from fitting the data to the form of eq. (6.6). This led to
the following expression of

:
(6.7)

with the following optimal values of parameters


and

. This representation led to a relative standard deviation

(with respect to the fitting function of

) of 5.1%.

In this chapter, the synthetic apparatus and the experimental procedures are discussed in
section 6.2, followed by the experimental results on the H2O-CO2 system in section 6.3. We
derived Henrys constant at a given temperature for each experimental data point (

, ) based on a model developed specifically for this setup, as discussed in section 6.4.By
comparing values for

from this work and calculations from the empirical models

developed by Carroll et al. [17] and Crovetto [18], we proved that the synthetic apparatus
was capable of generating reliable results. After validating the apparatus, it was then used to
measure the solubility of CO2 in 25 mass% aqueous MDEA as described in section 6.5.
6.2 Apparatus and experimental procedures
6.2.1 Apparatus
The main components in the synthetic apparatus include: a Hastelloy C-276 equilibrium cell,
a high pressure generator for liquid injection, pressure transducers and temperature sensors
which are connected to and monitored in real-time by a data acquisition unit. The schematic
diagram of the synthetic apparatus is illustrated in Fig. 6.1.

6. Solubility Measurements with the Static-synthetic Apparatus

(1)

High pressure generator for liquids

(2), (2)

(3)

High pressure cell (Hastelloy C-276)

(4)

(5)

Electrical cartridge heaters

(6)

Magnetic stirrer

(7)

PTFE coated magnetic follower

(8)

Temperature control unit

(9)

Digitalmultimeter (temperature readout unit)

(10), (10)

150

Gas cylinders containing CO 2 and He, resp.

Heater shell (Al)

Pressure readout unit

(11) Vacuum pump

(12)

Flask containing solvent component

(13) Atmosphere

(14)

Ultrasonic bath

(15) Personal computer


S

Safety rupture disk (Pmax = 35 MPa)

T1, T2

Platinum resistance thermometer

T3

Thermoprobe

P1, P2 Pressure transducer (0<p/MPa<20)

Needle valves (V1-V8)

High pressure stainless steel tube

PTFE tubing

Electrical connections/signal transmission cables

Figure 6.1 Schematic diagram for the synthetic apparatus

6. Solubility Measurements with the Static-synthetic Apparatus

151

6.2.1.1 Equilibrium vessel


The equilibrium cell as shown in Fig. 6.2, was
made of Hastelloy C-276 with an internal
volume of approximately 68 ml. The cylindrical
vessel had a pressure rating of up to 35 MPa.
Owing to the solubility of CO2 in elastomeric o-

Heater shell
body

ring materials, which could potentially lead to


leakage in the system, explosive decompression
damage and loss of CO2 into the elastomer, a
specially made gold-plated stainless steel o-ring
(Wills, WYKO Industrial Services Ltd., West

Lid
Seal retaining
plate
Cell body

Midlands, UK) was placed between the lid and


cell body. Elasticity of the o-ring was created by
filling the hollow, torus-shaped ring with nitrogen

Heater shell
base

at 20 MPa. To hold the parts together, eight M-10


bolts were screwed into the lid and cell body. The
system was also equipped with a heating

Figure 6.2 Assembly view for the


equilibrium vessel and heater shell

mechanism, i.e. an aluminium heater shell encased


the cell. This heater shell was powered by four 100 W cartridge heaters (Watlow Ltd.,
Nottingham, UK), which were controlled to within 0.02 K (Eurotherm, West Sussex, UK).
Thermal insulation of the cell during operation at elevated temperatures was achieved by
wrapping four layers of silicon rubber sponge around the heater shell with a total thickness of
approximately 12 mm. A glass-encased stirrer bar, rotated by the magnetic force from a
magnetic stirrer plate (MR Hei Standard, Heidolph, Essex, UK), was placed inside the cell to
enhance the equilibration .
6.2.1.2 High pressure generator
Liquid components, in this work either water or 25 mass% MDEA solutions, were injected
into the equilibrium cell in a stepwise manner by a high pressure generator (High Pressure
Equipment Company, Erie, USA). The procedure of the calibration for the high pressure
generator is detailed in the Appendix III. For the calculation of Henrys constant, precise
knowledge of the liquid volume injected in each step was essential. This information was
obtained through calibration with water, a liquid with well-understood physical properties, so
that the linear relation existing between the number of turns corresponding to a particular

6. Solubility Measurements with the Static-synthetic Apparatus

152

piston stroke and the volume displaced by the piston was established. Fig. 6.3 illustrates the
high pressure generator with a total volume of approximately 60 ml and pressure rating of 34
MPa.

Figure 6.3 Schematic diagram of a high pressure generator


6.2.1.3 Temperature and pressure measurements
The temperature inside the cell was measured by a Platinum Resistance Thermometer (Pt 100)
inserted into a thermowell drilled vertically in the cell wall. The temperature of the liquid in
the high pressure generator was monitored by a surface-type Pt sensor. Both signals were
captured with an Agilent 6 Digital Multimeter LXI (Model 34410A) where the resistance of
the probes was measured. Subsequently, the temperatures can be calculated from resistance
according to the Callandar van Ducen Equation [267]:
(6.8)
where
and

is the resistance at temperature , and

is the resistance at

= 273.15 K, and

are coefficients determined in the calibration of the sensor.

Two pressure transducers were used to measure the pressures in the gas inlet line ( ) and
liquid injection line ( ). The pressure transducer

(Druck, Model DPI260, Leicester, UK )

was only used to determine the pressure of the initial dry gas introduced into the equilibrium
cell while the vessel pressure during the experiment was only measured with

(Digiquartz,

Model 43K-101, Paroscientific Inc., Richmond, USA) located in the liquid injection line.
Real-time pressure data were indicated with a read-out unit (Model 710 Digiquartz) which
was interfaced with a computer through a data logging software (Digiquartz Interactive 2.0).
The accuracy of P1 was 0.01% of the full scale of 20.7 MPa as indicated by the manufacturer.

6. Solubility Measurements with the Static-synthetic Apparatus

153

6.2.2 Experimental procedures


6.2.2.1 Experimental methodology
Before discussing the experimental details, it is important to understand the methodology of
the experiment. The synthetic apparatus was designed to study the solubility of gases, like e.g.
CO2 or SO2, in liquid solvents, such as water. The word synthetic means that the
compositions of different components in different phases are derived from the knowledge of
the amount of each component in the cell. Therefore, only temperature, the total pressure, the
initial amount of CO2 introduced into the vessel, and the amount of water injected in each
step were measured during the experiment. No direct concentration analysis was carried out
on the liquid or gas phase. Another characteristic of the design in this work was that the
liquid phase itself served as the moving boundary to compress the gas phase into the liquid
instead of using a moving piston. In other words, there were two contributing factors to the
dissolution of CO2: one was the increased system pressure; another was the extra amount of
liquid introduced into the equilibrium cell.
6.2.2.2 Calibration
As stated above, the synthetic method requires precise knowledge of the global amount for
each component at any given equilibrium point. Consequently calibration was necessary for
both the high pressure generator (see section 6.1.1.2) and the equilibrium cell (i.e. accurate
volumes of both of them). The detailed calibration procedures are included in Appendix III.
6.2.2.3 Experimental procedures
Before starting experiments, the whole system including the equilibrium vessel, stainless
steel tubing, and valves etc. was completely cleaned, dried and evacuated with a vacuum
pump through V3. The experimental steps were divided into three parts:
(1) Filling the high pressure generator with solvent
Deionised water was loaded into a 500 ml glass bottle and placed in an ultrasonic bath to
degas for at least 30 min. Vacuum was applied to the water bottle to facilitate the removal of
the gas. The degas process was considered as completed when no visible bubbles emerged
from the water. The 25 mass% aqueous MDEA solution was prepared by weighing the
samples on a Kern analytical balance (AES 220-4, Bucks) which has both repeatability and
resolution at 0.1 mg. Then the solvent was filled into the high pressure generator from the
liquid reservoir via V5. During the filling process, V6 was always open to keep the pressure in

6. Solubility Measurements with the Static-synthetic Apparatus

154

the liquid reservoir at atmospheric pressure. After filling the high pressure generator with
solvent, V5 was closed.
(2) Filling the cell with gas
A minimum amount of water vapour and air should be left in the system before introducing
CO2. To minimize the interference from any remaining water vapour and air, the vessel was
repeatedly filled with CO2 and then evacuated, 2 to 3 times. Gas was introduced into the
thermostatic equilibrium cell via V2 and V4 to the desired pressure. After equilibration, which
usually took 3 to 5 min, V2 and V4 were closed. The pressure inside the cell was measured by
P2. By knowing the temperature and pressure of the gas, the amount of gas introduced was
calculated from the equation of state proposed by Span and Wagner [268].
(3) Executing solubility measurements
Next, V1 was switched on to allow solvent to fill the cross between V1, V2, V3, and V4. In
order to collapse any bubbles that might have been present in the liquid, the liquid was
pressurised to around 5 MPa for approximately 10 min. Prior to the injection of the solvent
into the cell, the pressure of the solvent was adjusted to just above the pressure of the gas
(normally 0.01 MPa higher) to prevent back-flushing of the gas. The liquid was brought into
contact with gas by turning on V4 for one turn. The strict turn of V4 was to ensure constant
dead volume of the valve.
Solvent was injected by manually operating the high pressure generator in a stepwise manner.
It usually took 45 to 60 min to reach equilibrium and during equilibration, the pressure
variation in the cell followed the trend shown in Fig. 6.4. The rapid increase in pressure
immediately after injection of liquid was due to the compression of the gas. As water is
almost incompressible, its introduction reduced the volume taken up by the gas and led to the
sharp increase in the total pressure. Immediately after the introduction of water, gas started to
dissolve, causing the decrease in pressure. The equilibrium was reached when the pressure
inside the cell stabilised. Since CO2 was only slightly soluble in water, the equilibrium
pressure was always higher than the previous state.
For each equilibrium point, the temperature, equilibrium pressure, and the volume of the
injected solvent were recorded. To convert the volume to the amount of water injected in
each step, its density was derived from the equation of state from Wagner and Pruss [269],
based on the knowledge of its temperature and pressure. The pressure was calculated by

6. Solubility Measurements with the Static-synthetic Apparatus

155

taking the mean value of two equilibrium states before and after the water was injected. For
aqueous MDEA solution, the conversion of the volume to molar quantity was different from
that for water: only the influence of temperature on the density was taken into account;
pressure effects were neglected due to the imited pressure ranges during the experiment.
5.2
5

P/MPa

4.8
4.6
4.4
4.2
4
500

1000

1500

2000

2500

3000

t/s

Figure 6.4 Change of total pressure plotted against time


The injection procedure was repeated until all the components in the cell became liquid phase.
Afterwards, a few more injections, usually 3 to 5 points, were conducted in the single phase
region. The whole process is illustrated in Fig. 6.5.

6
5

P/MPa

4
Two phase region
3

Bubble Point
2

One phase region

1
0
0

2
n/mol

Figure 6.5 Total pressure against cumulative moles of water injected at

= 313 K

6. Solubility Measurements with the Static-synthetic Apparatus

156

The initial stage with mild increase in total pressure as more water was introduced represents
the two-phase region, whereas the sharp increase in pressure with increasing amount of water
injected suggested that all the components thereafter existed in the liquid phase. The
intersection of these two branches is called the bubble point, i.e. the point at which all the
initial gas just dissolves in the amount of solvent injected at this particular pressure and
temperature.
6.2.2.4 Experimental overview
Experiments were conducted at five temperatures, i.e. 313 K, 333 K, 353 K, 373 K, and 393
K, in order to validate the apparatus. To evaluate the repeatability of measurements, for (333
and 393) K, more than one experiment was conducted with different initial filling pressures
of the CO2. The detailed experimental conditions are summarised in Table 6.1.
Table 6.1 Matrix for experimental conditions
/K
Initial Filling /MPa
Total no. of Experiments

313
0.4
/
/
1

333
0.1
0.2
/
2

353
0.1
/
/
1

373
0.1
/
/
1

393
0.1
0.4
0.9
3

6.3 Results and discussions


6.3.1 Calculation of the Henrys constant with graphical interpretation
For each experiment, Henrys constant was calculated using two methods: one was from
graphical interpretation of experimental results (using polynomial and linear correlations for
the two-phase and one-phase regions respecively) and another was via theoretical modelling
for the two-phase region. The first method is presented in this section, whereas the modelling
method will be discussed in Section 6.3.3.
As discussed before, when plotting the equilibrium pressure against cumulative moles of
water injected, a diagram with clearly distinguishable two branches can be obtained. In this
work, a polynomial equation was used for the two-phase region, while a linear correlation
was employed to describe the all-liquid region. These two branches intersect at the bubble
point of CO2 in water at the given temperature and pressure. At the bubble point, all the
components were in the liquid phase. The cumulative number of moles for H2O injected at
the bubble point was determined by solving the two equations representing the two branches,

6. Solubility Measurements with the Static-synthetic Apparatus

157

while the moles of CO2 in the liquid phase, equal to the initial filling amount, can be
calculated as per eq. (6.9):
(6.9)
where

, the cell volume, was carefully measured in the calibration process.

molecular weight of CO2.

is the

was calculated from the equation of state described by Span

and Wagner [268] at the cell temperature measured by T2 (as shown in Fig. 6.1) and the
initial filling pressure measured using P2.
The mole fraction of CO2 was then calculated based on the total amount of CO2 and H2O in
the cell according to eq. (6.10):
(6.10)
To calculate the Henrys constant, information of the partial pressure of CO2 was needed. In
this experiment, only the total pressure

in the cell was directly measured, which included

contributions from partial pressures of CO2 and H2O. Since the solubility of CO2 in water was
small, the vapour pressure of H2O could be considered to be the same as that of pure H2O at
the system temperature and calculated according to the formula suggested by Saul and
Wagner [246].
(6.11)
where
6.1.

, and

to

are parameters given in Table

is the temperature in the cell as measured by sensor T2. The partial pressure of CO2

was subsequently calculated from eq. (6.12):


(6.12)
Table 6.2 Constants in eq. (6.11)

-7.8582 1.83991 -11.781 22.6705 -15.939 1.77516

6. Solubility Measurements with the Static-synthetic Apparatus

158

Therefore, the most simplified form of the Henrys constant can be derived according to eq.
(6.13) assuming that the vapour and lliquid phase nonideality and the poynting factor are all
negligible,
(6.13)
6.3.2 Results from the graphical interpretation method
The results of an experiment at 313 K are shown in Fig. 6.5 as a plot of the total pressure
against cumulative moles of water injected. For other temperatures, the results are included in
Appendix V. To locate the bubble point, the two phase region was correlated with a second or
third order polynomial equation whereas the one phase region was represented by a linear
equation. By solving these two equations simultaneously, the bubble point was located and
Henrys constant was calculated according to the procedures outlined in section 6.3.1.
Henrys constants calculated from experimental data at five different temperatures are in
Table 6.3 together with the corresponding literature values.
Table 6.3 Comparison between experimental results and literature values

/K

Initial

/MPa

/MPa

_lit*/MPa

RD%

314.6

0.4032

241

238

1.1

335.0

0.1514

388

349

11.3

334.8

0.2087

386

348

10.9

355.0

0.1445

473

460

2.7

374.2

0.1369

538

511

5.3

394.2

0.1271

615

586

5.0

393.6

0.4074

602

585

3.0

394.4

0.9475

662

586

12.8

* Carroll et al. [17] and Crovetto [18]

6. Solubility Measurements with the Static-synthetic Apparatus

159

As can be seen from the table above, the differences between measured results and literature
values are quite large, except at T = 314.6 K. This can be attributed to the fact that as
temperature increases, the intersection of the two branches blurs. This is especially
problematic when arbitrary polynomial equations were used to fit the experimental data, i.e.
the choice of different orders of polynomial equation may result in large differences in the
location of bubble point. Therefore, fitting the data with a physical model rather than using
polynomial equations without any theoretical basis can enhance the accuracy of the
determination of the bubble point.
6.3.3 Calculation of the Henrys constant with theoretical modelling
(1) Vapour-liquid equilibria and phase non-ideality
For the temperature and pressure ranges studied in this work, the solvent was treated as
volatile, i.e. total pressure is the sum of partial pressures of CO2 and water. Non-ideality in
the gas phase was accounted for using the virial equation of state, while the liquid phase was
considered as ideal. In general, vapour-liquid equilibrium was described as the equality of the
fugacity for both components in the gas and liquid phases.
(6.14)
(6.15)
(6.16)
(6.17)
(6.18)
Here, subscript 1 refers to water, and 2 refers to CO2;
of H2O and CO2 in the vapour phase;

and

are the fugacity coefficients

is the activity coefficient of H2O in the liquid phase

on a mole-fraction scale normalized according to the convention of Raoults law (i.e.


as
while

);

is the fugacity coefficient of the pure solvent at its saturated vapour pressure,

is the activity coefficient of CO2 on a mole-fraction scale normalized according to

the convention of Henrys law (i.e.

as

). In addition,

is the partial molar

volume of component in the liquid phase at the composition, temperature and pressure in
question. Since the activity coefficient was considered to be unity on grounds of the low
concentration of dissolved CO2, the eqs. (6.15) and (6.17) were simplified to:

6. Solubility Measurements with the Static-synthetic Apparatus

160
(6.19)
(6.20)

The virial equation of state [270] is a good approximation to describe the gas phase
nonideality under low to moderate pressures (valid up to 10 atm), and was therefore chosen in
this work. Its truncated form after the second virial is:
(6.21)

In the above equation,

is the compressibility factor and

virial coefficient of a mixture. For a binary mixture,

is the second

is expressed as:
(6.22)

where

and

are second virial coefficients of pure components 1 and 2. They were

calculated from eq. (6.22) proposed by Bieling et al. [248]:


(6.23)
,

, and

are parameters listed respectively in Table 6.4.

Table 6.4 Constants in eq. (6.23)

CO2
H2O
(equivalent to

65.703
-53.53

184.854
-39.29

304.16
647.3

1.4
4.3

) is the interaction virial coefficient between components 1 and 2 and

may be calculated from eq. (6.23) as proposed by Plyasunov and Shock [271]:
(6.24)
Values for

are listed in Table 6.5.

Table 6.5 Constants in eq. (6.24)


0
1
2
3
-211.31 -729.48 1064.54 -656.13

6. Solubility Measurements with the Static-synthetic Apparatus

161

The fugacity coefficients for vapour phase CO2 and H2O were calculated according to eq.
(6.25):
(6.25)
(2) Mass balance calculations
Apart from obeying the phase equilibria equations, the system also needed to satisfy mass
balance. Since the global compositions of CO2 and H2O were known from experimental
data, the distribution of the two components between the two phases was derived as follows.
We define the

factor as:
(6.26)

Replacing eqs. (6.14), (6.16), (6.19), and (6.20) into eq. (6.26) for both water and CO2 leads
to:

(6.27)

(6.28)

Then , defined as the vapour fraction, is expressed in terms of the

factors by solving the

mass balance equations analytically:


(6.29)
where

and

are the overall mole fractions of components 1 and 2 respectively, shown as

below:
(6.30)
is the total amount of substance in the system. For CO2, it was calculated from the initial
conditions, i.e. cell volume, initial filling pressure and temperature. For H2O, the cumulative
amount for every injection step was obtained from the recorded volume, average temperature
and pressures during the injection process.

6. Solubility Measurements with the Static-synthetic Apparatus


At the bubble point,

162

and by substituting eqs. (6.27) and (6.28) into eq. (6.29), we

obtained:
(6.31)

is the system pressure at the bubble point. Meanwhile, since all the components were in the
liquid phase, and the liquid phase occupied the whole cell volume, the following equation can
be derived:
(6.32)
where

and

are respectively the molar volume of pure solvent and partial molar volume

of the gas in the solvent at infinite dilution. Here,


injected,
Finally,

is the cumulative amount of water

is the total amount of gas initially in the cell and

is the calibrated cell volume.

is the mean isothermal compressibility of the solution phase which was derived

from the measurements in the one phase region described later. The above equation includes
the effects from both the swelling (as a consequence of gas dissolution) and the compression
(owing to the pressure exerted on the liquid).
Eqs. (6.31) and (6.32) were solved together to find the solutions for

and total amount of

water in the system. For a more general case, i.e. in the two phase region, the calculation was
carried out as follows.
At a given temperature, an assumed value of Henrys constant would allow us to derive a
calculated value of the total pressure at any stage of the experiment, i.e. each injection stage.
In the analysis, the value of Henrys constant was optimised by minimisation of an objective
function based on the differences between the experimental and calculated total pressures.
Despite the straightforward methodology, the calculation of pressure from the estimated
Henrys constant was not trivial. This was due to the fact that, with the inclusion of nonideality in the vapour phase, the

factor becomes composition dependent so that an iterative

solution procedure is required. A double iteration loop was required to find the solution: for
each injection step, pressure was adjusted in the outer loop and the vapour phase
concentration was adjusted in the inner loop. Starting with a trial value of
for

or

, the

and an estimation

factors for the two components were calculated. Then eq.(6.29) was solved

6. Solubility Measurements with the Static-synthetic Apparatus

163

analytically to obtain the value for .

and

and

were derived based on the values of

factors, as per eq. (6.33):


(6.33)
The calculated values for vapour phase concentrations were found using:
(6.34)
or

was then compared with the initial estimation and the latter was adjusted until it

converged with the former. This was the inner loop of the whole calculation. Meanwhile, a
calculated value of the total pressure was determined according to the equation of state of the
vapour phase:
(6.35)

where

is given as:
(6.36)

The trial procedure was continued until it converged with the calculated

from the above

equation: this formed the outer loop of the whole procedure. During the numerical solution
process, we found that values for

converged quickly whereas the

values converged

much more slowly. To facilitate the convergence of , a step-wise, root-searching procedure


was applied, as detailed in Appendix VII.
The above numerical procedures were implemented as a Visual Basic Application (VBA) in
the MS Excel. The VBA code is given in Appendix VII. This solver was used to converge the
two loops and to find the pressure, vapour and liquid phases compositions for any given
temperature

, cumulative amount of water injected

, and estimated Henrys constant.

Several Henrys constants were tried in the calculation and the final optimized value for a
particular temperature was derived by minimising an objective function:

(6.37)

where

is variance of the th measurements of the systems total pressure. It is worth noting

that the procedures discussed above were applicable for the two phase region and the bubble

6. Solubility Measurements with the Static-synthetic Apparatus

164

point. When beyond the bubble point, i.e. all the components were in the liquid phase, the
system was described by the following equation:

(6.38)

As mentioned before,

was derived from the experimental measurements in the one-phase

region, and this value was used as a constant to simulate the two phase region.
The results for the Henrys constants were obtained by applying the above analysis to data in
the two phase region for various temperatures and are shown in Table 6.6.
Table 6.6 Comparison of the Henrys constants obtained using the model of section 6.3.3
with literature values [17]
/K

Initial

/MPa

/MPa

_lit*/MPa

RD%

314.6

0.4032

238

238

0.1

335.0

0.1514

362

349

3.7

334.8

0.2087

353

348

1.4

355.0

0.1445

456

460

-0.9

374.2

0.1369

518

511

1.3

394.2

0.1271

575

586

-2.0

393.6

0.4074

579

585

-1.0

394.4

0.9475

572

586

-2.4

The significant improvement in the quality of the results is apparent. Generally, the relative
deviation between the calculated and measured Henrys constant is within 2.5% with only
one outlier at 335 K. This may be caused by not fully evacuating the system before filling
with CO2. As the experiments progressed, more experience was developed to overcome this
issue. One way of purging out the remaining air was to repeatedly evacuate and refill the cell

6. Solubility Measurements with the Static-synthetic Apparatus

165

with CO2 3 to 5 times. Apart from the systematic errors discussed above, the results are also
subject to random errors which will be discussed in the next section 6.3.4.
6.3.4 Sources of uncertainties
(1) Calibration of the high pressure generator
The detailed procedures for calibration of high pressure generator are provided in Appendix
III. The purpose of the calibration was to establish the correlation between the number of
turns corresponding to a particular piston stroke and the volume displaced by the piston. This
was achieved by weighing different portions of water expelled from the pump. To convert the
mass to volume, Wagner-Pruss equation of state was employed. Subsequently a plot of
volume ( ) versus number of turns of the pump ( ) was obtained and fitted with a linear
equation. For each turn, the volume displaced was found to be (0.7000 0.0002) cm3.
(2) Calibration of the cell volume
Precise knowledge of the cell volume was crucial for the synthetic method, as it was used
both for the calculation of the total CO2 amount and the volume occupied by the vapour
phase in the two phase region. The calibration employed the already-calibrated high pressure
generator and the steps were similar to the actual experiment. The detailed information is
included in Appendix III. From the calibration, the cell volume was determined to be (67.024
0.002) cm3.
(3) Temperature and pressure measurements
As described earlier in section 6.2.1, the fluctuation of the cell temperature was controlled
within 0.02 K with four 100 W cartridge heaters and a Pt-100 RTD sensor both connected to
a temperature controller. Accurate measurement of the temperature was achieved by using
another Pt-100 RTD sensor and the uncertainty was estimated to be 0.021 K. The calibration
of this Pt-100 RTD sensor was through comparison of the water triple point with a standard
platinum resistance thermometer. Pressure of the initial CO2 and the total pressure of the
system during the experiment were measured with two separate transducers. The uncertainty
of the former was estimated to be 0.5 kPa, and of the latter 2 kPa.
6.3.5 Uncertainty analysis
The standard uncertainty in the Henrys constant calculation was calculated based on eq.
(6.39), taking account of the contribution from the uncertainties in the cell volume, volume of
the liquid dispelled by the high pressure generator, temperature and pressure measurements:

6. Solubility Measurements with the Static-synthetic Apparatus

166

(6.39)
is the combined standard uncertainty in the Henrys constant,

where

= 0.05 K,

= 0.002 cm3, and

= 2 kPa,

= 0.0002 cm3. Each contribution was

assessed with the simulation model in section 6.3.3 by varying the parameter to its maximum
error range while keeping others the same. The combined standard uncertainty in the Henrys
constants measurements was found to be approximately 2 MPa and is mainly due to the
contribution from

6.4 Measurement of the solubility of CO2 in aqueous MDEA system with the synthetic
apparatus
The successful determination of the Henrys constant for CO2 in H2O validated the
experimental procedures and the numerical model. The next step was to apply the synthetic
apparatus to determining the solubility of CO2 in amine systems. As chemical solvents,
amines have a much greater capacity to dissolve CO2 than pure water. Besides, for a given
temperature, the solubility of CO2 was no longer a linear function of its partial pressure
owing to the chemical reactions in the system, i.e. the Henrys constant which had been used
to describe the physical solubility in water was not suitable for the amine system. Instead, the
pressure of CO2 was plotted against the solubility in the solution. Generally for amine
systems, the solubility is expressed as the solution loading, defined as:
(6.40)

are the number of moles of CO2 and amine in the solution phase. 25 mass%

aqueous MDEA was chosen to be studied in the synthetic apparatus. MDEA is a tertiary
amine with no free hydrogen atom on the amine functional group, i.e. nitrogen atom. This
means it cannot react with CO2 directly to form carbamate (a compound with the functional
group R2N-COO-). Instead, MDEA protonates and creates a basic environment that enhances
the dissolution of CO2. The reactions in the solutions phase are:
(I)
(II)
(III)

6. Solubility Measurements with the Static-synthetic Apparatus

167
(IV)

At the gas-liquid interface, the CO2 dissolution is represented by:


(V)
6.4.1 Methodology for the calculation of CO2 solubility from experimental measurements
Because of the large solubility of CO2 in amine solution, the behaviour of the system was
considerably different from CO2-H2O. In the two phase region, contrary to the latter, the
introduction of amine solutions led to a decrease in the total pressure, as shown in Fig. 6.6.
6
5

P/MPa

4
3

One phase region


Two phase region

2
1
0
0

10

20

30

40

50

60

70

80

m/g

Figure 6.6 Total pressure against cumulative moles of solvent for MDEA-H2O-CO2 at
313 K

As can be seen from the above graph, in the two phase region, total pressure decreased slowly
with the introduction of the solvent into the vessel. This suggested that the solution loading
was also decreasing, as it varied directly with the pressure. The downhill trend was continued
until the system approaches the single-phase liquid region. Theoretically, the pressure should
only start to increase when the system enters the single-phase liquid region; however, owing
to the non-ideality in the experiments, such as air remained in the vessel and the
incompletely degassed solvent, the pressure generally started to rise before the system
reached the single-phase region. As can be seen from the graph above, the pressure started to
increase just after the cumulative mass of the solvent injected exceeded 50 g. These abnormal
points were not included in the calculation of CO2 solubility. In order to derive the partial
pressure and solution loading of CO2, the following calculation procedures were employed.

6. Solubility Measurements with the Static-synthetic Apparatus

168

Prior to the calculation, several assumptions were made to simplify the calculation while not
affecting the validity of the results. The first assumption was that there was negligible vapour
phase MDEA in the system. This was well justified as the vapour pressure of pure MDEA at
293 K is less than 1.3 Pa. Although the experiments were conducted at 313 K with 25 mass%
aqueous MDEA, the vapour pressure of MDEA was considered to be negligible compared to
that of water (at 293 K, the vapour pressure of MDEA is < 1.3 Pa vs. 233.9 Pa for water).
Secondly, it was assumed that the vapour pressure of H2O could be approximated using
Raoults law. This assumption can be improved with a better correlation such as the ones
taking into account the influence from phase non-ideality, or measuring the vapour pressure
of the 25 mass% MDEA directly. However, the latter option was not considered in this work
owing to the rather large uncertainty of the pressure transducer used in the current setup, i.e
2 kPa, especially compared to the low vapour pressure of the solvent. Finally, it was
assumed that the compressibility of the solvent was the same as that in the case of water, and
that the dissolution of the CO2 had negligible influence on the solvent volume. These two
assumptions were justified as follows: first, a sensitivity evaluation of the compressibility on
the results showed that the effect from it was very small (the increase of the compressibility
by 9 times will only result in 0.15% decrease in the solution volume); second, the majority of
the dissolved CO2 was in the ionic form, i.e. HCO3-, which has negligible influence on the
solution volume.
For each injection step, to convert the volume of solvent to amount, knowledge of the
solution density was essential: this information was obtained from the formulae suggested by
Hsu and Li [301]. Densities of pure components, i.e. MDEA and H2O, were correlated with a
temperature-dependant equation expressed as:

Here,

, and

(6.41)

are component-associated parameters with values listed in Table 6.7.

is the temperature of the liquid in K.


Table 6.7 Constants in eq. (6.41)
1

H2O
MDEA

0.863559
1.22864

1.21494 10-3
-5.4454 10-4

-2.5708 10-6
-3.3593 10-7

The density for the mixed solvent was calculated based on the molar weight and molar
volume of the mixture. The molar volume of the mixture was a combination of the molar

6. Solubility Measurements with the Static-synthetic Apparatus

169

volumes of the pure components and the Redlich-Kister-type equation for the excess molar
volume. For a binary system, the Redlich-Kester equation taken from Prausnitz et al. [115] is:

where

the mixture.

(6.42)

are the mole fractions of components 1 and 2.

is the number of components in

is defined as the pair parameter and is assumed to be temperature-dependent:


(6.43)

Values for parameters , , are given in Table 6.8.


Table 6.8 Constants in eq. (6.43)

-2.88774 10
-2.06623 10

0
1

6.95810 10-2
6.36707 10-2

-5.0304 10-7
0

Subsequently, the molar volume of binary liquid mixture was calculated based on those of the
pure components and the excess volumes of the mixture:
(6.44)
Finally, it was not difficult to obtain the density of the mixture from the average molar mass
and the molar volume of the mixture.
(6.45)
In the above equation,

is the molar mass of component .

From the correlation of the mixture density with temperature, the mass and volume of the
solvent in the cell at each stage were conveniently derived from the volume of liquid injected.
Meanwhile, in the gas phase the water vapour pressure was calculated from Raoult's law.
(6.46)
was correlated by Saul and Wagner [246] as below:
(6.47)
where,

6. Solubility Measurements with the Static-synthetic Apparatus

170

The amount of CO2 in the vapour phase was calculated from its equation of state with the
above information of its partial pressure, temperature and volume (total volume less the
volume taken up by the liquid phase). As the total amount of CO2 initially filling the cell was
known a priori, the moles of CO2 dissolved into the liquid phase was then obtained and used
to calculate solution loadings. The amount of amine in the system was derived from the total
amount of the solution and its mass concentration. One special point to note is that not all the
liquid volume displaced by the pump entered into the cell. This was because of the short
tubing connecting the valve and the vessel as well as the extra volume when opening the
needle valve, both of which were dead volumes in the system and were deducted from the
total volume displaced by the pump to obtain the true volume in the cell. The total dead
volume in the system was estimated to be 0.245 ml and was accounted in the final analysis by
subtracting it from the total volume injected.
6.4.2 Results
Three experiments were conducted at

= 316 K, with 25 mass% aqueous MDEA solution

and different initial CO2 pressures, namely: 1.334 MPa (Exp. 1), 0.797 MPa (Exp. 2), and
0.919 MPa (Exp. 3). Based on the above calculation steps, the partial pressures and loadings
of CO2 in the solution were derived and are presented in Table 6.9 and Fig. 6.7. With the
gradual introduction of amine solvent, the pressure in the cell decreased and so did the
solution loading. Because the methodology applied in this work was only suitable for the two
phase region, once the downhill trend of pressure started to reverse - an indication of the
system entering into the all-liquid phase region - the date points onwards were not used in
this analysis.

Table 6.9 CO2 (1) solubility data in a solution of water (2) (75 mass%) and MDEA (3) (25
mass%) at
Exp. 1
MDEA% =
24.78%
0.978
0.918
0.800
0.665
0.561
0.537

= 316 K

= 315.9 K
/kPa
518
228
112
50
31
28

Exp. 2
MDEA% =
25.23%
0.923
0.826
0.794
0.762
0.749
0.730

= 315.9 K
/kPa
264
120
99
81
76
68

Exp. 3
MDEA% =
24.35%
0.980
0.930
0.912
0.894
0.869
0.844

= 315.0 K
/kPa
412
279
238
199
168
140

6. Solubility Measurements with the Static-synthetic Apparatus


0.514
0.484
0.457
0.425
0.404
0.379
0.362
0.342

26
24
22
20
19
18
17
17

0.716
0.697
0.686
0.668
0.656
0.640
0.629
0.613
0.603
0.589
0.579
0.566
0.557
0.544
0.536
0.525
0.506
0.489
0.472
0.429
0.416
0.404
0.382
0.372
0.353
0.336
0.321
0.306
0.300
0.287
0.276
0.265

64
58
55
51
49
45
43
41
39
37
35
34
32
31
30
28
27
25
24
21
20
19
18
17
17
16
15
15
14
14
14
13

171
0.817
0.807
0.790
0.779
0.762
0.751
0.734
0.723
0.707
0.697
0.682
0.672
0.657
0.643
0.634
0.620
0.612
0.599
0.591
0.579
0.572
0.560
0.553
0.543
0.536
0.526
0.520
0.510
0.504
0.496
0.490
0.481
0.476
0.468
0.463
0.455
0.451
0.443
0.439
0.432

117
109
98
91
83
78
71
67
62
59
55
52
49
46
44
41
40
38
37
35
34
32
31
30
29
28
28
27
26
25
25
24
24
23
23
22
22
22
21
21

6. Solubility Measurements with the Static-synthetic Apparatus

172

1000

P_total / kPa

100

10

1
0.0

0.2

0.4

0.6

0.8

1.0

CO2

Figure 6.7 Comparison of CO2 solubility in 25 mass% MDEA at

= 313 K: ( ):

Silkenbumer et al., [71]; ( ): Sidi-Boumedine et al., [74], manually-operated apparatus; (


): Sidi-Boumedine et al., computer-operated apparatus; (
analytical-apparatus; (

): and at

): Sidi-Boumedine et al.,

= 316 K: this work, 1.334 MPa; ( ): this work, 0.797

MPa; ( ): this work, 0.919 MPa.


Since only total pressure was directly measured in this work and derived partial pressure of
CO2 was based on the assumption of the solvent vapour pressure, the comparison with
literature data was divided into two categories: one is with total pressure; the other is based
on the derived partial pressure. The experimental results were first plotted with total pressure
vs. CO2 loading and compared with the available literature sources in Fig. 6.7. Silkenbumer
et al. [71] employed an automatic, analytical apparatus to measure the CO2 solubility in 2.632
mol/kg MDEA solution at 313 K. Gas chromatography was used for the analysis of both gas
and liquid phases. Sidi-Boumedine et al. [74] measured the solubility of CO2 in 25 mass%
aqueous MDEA solutions with three types of apparatus: computer-operated, manuallyoperated, and analytical apparatus. Their analytical apparatus was only applied at pressures
above 500 kPa owing to the requirements of the GC sampling system; therefore, only one

6. Solubility Measurements with the Static-synthetic Apparatus

173

point is included in the above graph. It was also reported that all three types of equipment
produced good quality data with less scattering than other literature sources. As can be seen
from Fig. 6.7, the data from the synthetic apparatus of this work agree reasonably well with
the data of Sidi-Boumedine et al. The difference in the data sets of this work and those of
Sidi-Boumedine et al. may be attributed to the temperature difference in the measurements
(316 K in this work vs. 313 K in Sidi-Boumedine et al.).
Most of the available literature data were reported in the form of CO2 partial pressure.
Although not directly measured, it was also derived when calculating the solution loading. In
Fig. 6.8, the CO2 partial pressure is plotted against loading for the data from this work and
the available literature.
1000

P_CO2 / kPa

100

10

0.1

0.01
0.0

0.2

0.4

0.6

0.8

1.0

CO2

Figure 6.8 Comparison of CO2 solubility in 25 mass% MDEA at

= 313 K: ( ): Sidi-

Boumedine et al., computer-operated apparatus; ( ): Sidi-Boumedine et al., manuallyoperated apparatus; ( ): Jou and Mather, [119]; ( ): Austgen et al., [203]; and at
(

): this work, 1.334 MPa; ( ): this work, 0.797 MPa; ( ): this work, 0.919 MPa

= 316 K:

6. Solubility Measurements with the Static-synthetic Apparatus

174

From the above plot, it can be seen that the data from this work showed good agreement with
the majority of the previous data, except those from Jou and Mather [119] at high loading
range, i.e. over 0.8. Their data set showed higher loadings than all the other literature data at
the same pressures. A quantitative comparison of the data from this work and other literature
sources is presented in Chapter 5 Thermodynamic Models for the CO2 in Single and Blended
Amine Systems.
6.4.3 Other sources of uncertainties and limitations
The uncertainties in the temperature and pressure sensors have already been discussed in
section 6.3.4. The propagation of these errors to the derived loadings was estimated using the
calculation model for amine systems. The combined standard uncertainty of the solution
loading was also calculated using eq. (6.39) and was found to be 0.0007.
Another contribution to the uncertainty in the results was the estimated dead volume in the
system, which in fact has quite a large impact on the derived solution loadings. A change of
50%, i.e. from 1 ml to 0.5 ml, would change the loadings by as much as 7% particularly when
the injected solution volume was comparatively small, i.e. for the high loading measurements.
For the low loading measurements, the major error came mainly from the uncertainty in the
pressure transducer, which may result in large deviations in the low pressure measurements
(the absolute uncertainty of the pressure transducer is 2 kPa as discussed in section 6.3.4).
Despite the above possible uncertainties, the setup was proved to be capable of producing
reliable solubility data for CO2-MDEA-H2O. Certain caveats in the existing setup hinder a
wide application of the apparatus to measure all types of amines at the complete temperature
range relevant to the carbon capture process. However, it could be improved by replacing the
pressure transducer with one suitable for the measurement range and by reducing and/or
better characterising the dead volume.

7. Conclusions and Future Work

175

Chapter 7

Conclusions and Future Work


This chapter presents the conclusions and future work. It starts by laying out a summary of
the findings from this work according to the three main areas of research, namely, staticanalytical apparatus, quasi-chemical models, and synthetic setup. Subsequently, conclusions
are drawn based on the measurements from this work and relevant information collected from
literature on the amine systems studied herein. Finally, recommendations for future work are
laid out, mainly in two aspects: the parameters required to fully assess the potential of amine
solvents; and the potential amine blends worth investigating.

7. Conclusions and Future Work

176

7.1 Summary of this work


This work focused on the thermodynamic properties, especially the VLE, of amine-H2O-CO2
systems and can be divided into three main parts: experiments with the static-analytical
apparatus, modelling using quasi-chemical thermodynamic models, and experiments with a
static-synthetic apparatus.
7.1.1 Static-analytical apparatus
A new static-analytical apparatus was built in this work. The liquid phase compositions were
analysed using gas chromatography while the gas phase concentrations were derived from the
total and solvent vapour pressures. To overcome problems encountered with liquid-phase
analysis, an innovative liquid sampling system was devised which allowed sampling at
pressures up to 0.5 MPa. Calibration of the gas chromatograph was achieved through an insitu method based on knowledge of Henrys constant of CO2 in water. This apparatus and
the analytical methods were validated by measuring the solubility of CO2 in 30 mass%
aqueous MEA solutions at 313 K and 393 K. The satisfactory results (2.1% for

of data

of this work vs 3.6% for those of Jou et al. [73]) proved the feasibility of applying this
apparatus to measuring the solubility of CO2 in aqueous amine solutions.
Subsequently, the newly-built apparatus was employed to measure the solubility of CO2 in 30
mass% aqueous AMP solution. Often referred as the sterically-hindered counterpart of MEA,
AMP is one the most commonly-used amines in acid gas treating processes. However, to our
knowledge, there were very few publicly available solubility data for this system under
stripping conditions (T > 373 K); besides, the reported data show quite substantial
discrepancies [272]. To complement the existing database, the CO2 solubility of this solvent
system was measure from T = (313 to 393) K. In order to compare the cyclic capacity of
different amines, we defined in this work the equilibrium-equivalent CO2 partial pressure in
the stripper to be 100 kPa. Two partial pressures, namely 3 kPa and 10 kPa [277, 300] were
chosen as the equilibrium-equivalent absorber conditions. Evaluation of CO2 solubility in
potential aqueous amine-based solvents at low CO2 partial pressure, were considered for the
absorber to reflect the different extent of reaction. The measurements of this work confirmed
that AMP exhibited greater cyclic capacity than MEA at the same mass concentration. If the
equilibrium-equivalent CO2 partial pressure at the absorber is 10 kPa, the cyclic mole-loading
capacity of AMP is 2.4 times of that for MEA. When taking into account the difference in the

7. Conclusions and Future Work

177

molecular weight of these two amines, the cyclic mass-loading of AMP is only 60% higher
than that for MEA.
Despite the many advantages of AMP compared to conventional MEA including larger cyclic
capacity, low corrosiveness and better resistance to degradation, the undesirable property of
slow reaction kinetics makes the theoretical cyclic loading of AMP difficult to realise. A
commonly-used [63, 68, 236, 240] method of compensating for this drawback is through
blending the amine with another solvent with a fast reaction rate, so that the resulting blends
would potentially possess the advantages of both solvents. In this work, PZ was selected as
the activator to the AMP system and two different concentration mixtures were investigated,
including: 25 mass% AMP + 5 mass% PZ and 20 mass% AMP + 10 mass% PZ. If it is
assumed that the absorber and stripper temperatures are respectively 313 K and 393 K and
the equilibrium-equivalent CO2 partial pressure in the absorber is at 10 kPa, both these amine
blends possess reduced theoretical cyclic mole-loading capacity compared to the AMP at
equivalent mass concentration by approximately 8% and 12.5% respectively. This indicates
that at least a certain degree of trade-off is unavoidable between the cyclic capacity and
reaction rate. When cyclic mass-loading is used instead of mole-loading, both of the
AMP+PZ blends have 12.5% less CO2 capacity than AMP solution, while the differences
between the two blended systems are almost indistinguishable.
Apart from the primary and sterically-hindered amines, the third category studied in this work
was tertiary amines. DMMEA was selected to be studied in this work owing to its similarities
to AMP in terms of the functional groups and molecular weight so that the differences
between them can be largely attributed to the type of amine. Although the maximum
absorption capacity of the sterically-hindered and tertiary amines should be the same (since
both of the amines form predominantly carbonates/bicarbonates with CO2 instead of
carbamate), the measurements from this work showed a certain degree of difference in these
two amines - 8.3% less cyclic capacity of DMMEA compared to AMP when the equilibriumequivalent CO2 partial pressure is at 10 kPa. Compared to AMP, for which the cyclic capacity
is reduced when it is partially replaced with PZ, DMMEA appears to be subject to negligible
influence when it is replaced with PZ on an equal-mass basis. If 10 kPa is assumed to be the
equilibrium-equivalent CO2 partial pressure, the 25 mass% DMMEA + 5 mass% PZ mixture
actually has about a 5% greater cyclic capacity than amine blends containing 25 mass% AMP
and 5 mass% PZ.

7. Conclusions and Future Work

178

7.1.2 Quasi-chemical modelling


Another aspect of this work concerned the development of suitable quasi-chemical models
for the amine systems studied. In the first place, a - model was developed for MEA-H2OCO2 and AMP-H2O-CO2 systems separately. This model was based on the principle of using
activity and fugacity coefficients to describe the liquid and vapour phase non-ideality
respectively: the Debye-Hckel equation, as modified by Guggenheim, for the representation
of liquid phase non-ideality; and the virial equation of state truncated after the second term
for the vapour phase non-ideality. It was assumed that the gas phase non-ideality expression
contained no interaction parameter, whereas the number of adjustable parameters in the
modified Debye-Hckel expression was dependent on the number of species in the solution.
The values of these interaction parameters were obtained through regression of the
experimental data. For both amine systems, there were 30 possible adjustable parameters
initially and these were reduced to 10 and 14 for MEA-H2O-CO2 and AMP-H2O-CO2,
respectively, after a sensitivity analysis.
The Kent-Eisenberg model is a simpler but less rigorous model in which the liquid phase
non-ideality is lumped into selected amine-related equilibrium constants. This treatment of
non-ideality means that fewer adjustable parameters are required which reduces the
computational complexity compared to the - approach. In this work, the experimental data
produced with the static-analytical apparatus for the AMP, DMMEA and their PZ blended
systems were correlated with the Kent-Eisenberg model. The data for the single and blended
amine systems (e.g. AMP and AMP+PZ) were regressed simultaneously and for the sake of
simplicity the carbamate formation in the AMP based system was neglected (the carbamate
formation was not possible in DMMEA). After the sensitivity analysis, it was identified that
15 adjustable parameters were needed for both cases (AMP-based and DMMEA-based
systems).
The objective functions of these models could be based on either the solution loading or the
CO2 partial pressure. As the virial equation of state used in the - approach contains CO2
partial pressure as a parameter, to avoid additional iterative procedures the solution loading
was selected as the basis for the optimisation function; whereas, for the Kent-Eisenberg
model a pressure-based optimisation function was utilised.
In addition to the correlation of the solubility data, quasi-chemical models can provide
speciation predictions in the solution phase. The concentration profiles of the species as CO2

7. Conclusions and Future Work

179

loading increases are important for the reaction kinetics in the loaded solutions [192]. In
blended amine systems, the speciation becomes much more important since the amount of
free activator will have a strong effect on the enhancement.
Apart from correlation of the experimental data of the static-analytical apparatus, the two
models were also applied to the MDEA-H2O-CO2 system measured with the synthetic setup.
The experimental data of Sidi-Boumedine et al. were also included in the parameter
regression which covered two different amine concentrations and three temperatures. Both
models can satisfactorily correlate the experimental data. The

of the - approach is 5.8%

between the experimental and model loadings while for the Kent-Eisenberg model it is 18.4%
(or 6.8% on the solubility loading basis) on the pressure basis. This shows that even the value
of

based on CO2 partial pressure is much larger than the

between experimental

and predicted solution loadings; the correlation quality can be relatively similar.
7.1.3 Synthetic apparatus
A synthetic apparatus was also used in this work to study the solubility of CO2 in water and
amine solvents. The synthetic apparatus and the modelling procedures employed to treat the
experimental data are both innovative; therefore the initial objective was to validate them
using a well studied system CO2 and H2O. The validation was carried out at five different
temperatures from 313 K to 393 K and the raw experimental data were analysed to obtain
Henrys constants by means of a data reduction model implemented in Excel VBA. The
experimental results were compared with literature values from Carroll et al. [17]. The
apparatus and procedures were successfully validated as the majority of the relative
deviations between the experimental and literature data were within 2.5%.
Subsequently, we explored the application of the synthetic apparatus to aqueous amine
solutions. Considering its low vapour pressure and relatively simple solution chemistry,
MDEA was selected and its 25 mass% aqueous solution was measured at 316 K. Unlike the
CO2-H2O system, the solubility of CO2 in MDEA solution is large and this led to an initial
decrease in the total pressure as solvent was injected. Besides, as a chemical solvent, the
apparent solubility of CO2 in MDEA is a combination of chemical and physical interactions,
so the Henrys constant which was used to describe the physical solubility alone is not
applicable in this case. As a result, a different calculation procedure was employed in which
the solution loading was derived at each equilibrium point (corresponded to each injection
stage). This calculation procedure took into account the vapour phase non-ideality with the

7. Conclusions and Future Work

180

virial equation of state truncated after the second term and the temperature dependence of the
solution volume. The partial molar volume of CO2 in the solution was neglected in the
calculation of the CO2 loading in liquid phase as the molecular form only existed in minor
concentration. Three experiments with different initial CO2 filling pressure were conducted
and the results were compared favourably with available literature values.
From the solubility measurements with amine systems, several limitations of the synthetic
apparatus could be identified: firstly, the dead volume (~ 0.245 ml) in the system could affect
the results especially when the total injected volume in the vessel is small, despite the fact
that it is small compared to the total volume of the vessel (~ 67 ml); secondly, when a small
amount of liquid is injected (the exact quantity is dependent on the volatility of the solvent at
a certain temperature), the method is not suitable for amines with high volatility and
experiments at high temperatures, since the extent of solvent evaporation could lead to
significant deviation in the estimation of the solvent concentration and solvent vapour
pressure. In general, the accuracy of the results of the synthetic apparatus increases when
more solution is in the vessel; however, when the system approaches the bubble point, the
experimental results are extremely sensitive to impurities (e.g. air) in the system. If the total
pressure starts to increase before the system enters the single-phase liquid region, these
experimental points are not reliable due to the influence of these impurities.
7.2 Conclusions
The state-of-the-art process of the post-combustion capture is amine scrubbing which uses
15-30 mass% aqueous MEA solution to separate CO2 from flue gas in repeated cycles.
Despite being a mature process, a number of disadvantages, particularly the high energy
penalty, hinder the large-scale deployment of the technology. Besides, MEA is susceptible to
degradation and causes corrosion to the equipment. It has been estimated that the inclusion of
amine scrubbing into a coal-fired power plant would lead to at least a 30% reduction in its
thermal efficiency [22] while the cost of electricity could potentially increase by as much as
80% [273]. Many efforts have been devoted to reduce the energy penalty and additional costs,
of which developing advanced solvents has received significant attention [274-276].
This work attempted to advance the understanding of the relationships between the types of
amines and their cyclic capacity through vapour-liquid equilibrium measurements. Together
with the kinetics and other process-related information, including degradation, this will be
useful in the selection of efficient amine solvents. Solubility data of CO2 in amine solvents is

7. Conclusions and Future Work

181

also vital for process design. Considering the scarcity of the CO2 solubility data in certain
amines, we aimed to produce reliable experimental data and built a new static-analytical
apparatus. We also developed two quasi-chemical models to interpret the experimental data.
In the following sections, the conclusions from this work will be laid out in detail.
7.2.1 CO2 absorption in 30 mass% aqueous AMP and AMP+PZ blends
From the experimental measurements of this work, we conclude that the cyclic loading
capacity of AMP is considerably larger than that of MEA: the cyclic mole-loading capacity of
the former is between 100% (PCO2_absorber = 3 kPa) and 140% (PCO2_absorber = 10 kPa) greater
than the latter; while the increase in cyclic mass-loading capacity of AMP compared to MEA
is between 25% (PCO2_absorber = 3 kPa) and 60% (PCO2_absorber = 10 kPa). The increased cyclic
loading capacity means that a smaller solvent recycling rate is needed which brings benefit to
the process economics. Besides, the smaller heat of absorption with CO 2 of aqueous AMP
solution compared to MEA solution has been reported by several authors [277-279] indicates
that the energy penalty in the reboiler for AMP is also less than that for MEA system. In
addition, AMP is more resistant to oxidative and thermal degradations compared to MEA
[280-282]. These advantages make AMP a promising candidate for improving the process
economics compared to the MEA system.
Apart from the above characteristics, the absorption rate is also an important factor that will
affect the economics of the process. The absorption rate is controlled by thermodynamics,
which establishes the driving force for the absorption; reaction kinetics, which enhances the
absorption rate far beyond what could be achieved with physical absorption alone; and mass
transfer, especially when the reaction rate of the gas with the alkanoamine is fast enough to
deplete the amine in the boundary layer, such as is the case with primary amines, secondary
amines and PZ. As stated in a DOE report [192], reaction kinetics as well as chemical
equilibria are important for the CO2 absorption into primary or secondary amines; whereas
for tertiary or sterically-hindered amines, the CO2 absorption process is controlled by the
effective rate constant of reaction with the amine and the chemical equilibria play a minor
role. Many authors reported the kinetic data of various amines, as summarised partially in
Table 7.1.

7. Conclusions and Future Work

182

Table 7.1 Literature data [192] on the reaction kinetics of CO2 in aqueous amines at 298 K
Amine
MEA
DGA
DEA
MDEA
TEA
PZ
AMP
DMMEA

K298K (m3kmol-1s-1)
3703-8400
3990-5923
351-1500
2.4-6.2
1.6-50
68000
612-681
27.4

Mechanism
First order
First order
Zwitterion
Second order
Second order
First order
Zwitterion
First order

References
[208, 281]
[208, 282]
[283, 284]
[285, 286]
[287, 288]
[77]
[207, 226]
[233]

As a sterically-hindered amine, the reaction kinetics of AMP are almost an order of


magnitude smaller than those of primary amines, such as MEA and DGA. The reaction
kinetics of AMP with CO2 are a main limiting factor for the CO2 absorption process while the
chemical equilibria only plays a minor role. Therefore, it is of critical importance to enhance
the reaction rate of the system in order to make the solvent commercially viable. One
convenient way is through blending it with another amine which possesses a fast reaction rate.
From Table 7.1, we can clearly see that PZ is a good candidate for an activator. Lensen [217]
discussed the promoter effect of PZ to other amines and concluded that it is a more effective
activator than MEA or DEA. In fact, it can even be used to activate MEA, as reported by
Rochelle et al. [192]. The use of PZ as a promoter of AMP has also been extensively studied
by many researchers [218, 224, 290-292]. Samanta and Bandyopadhyay [223] investigated
the CO2 absorption rates of different AMP and PZ blends. They used an enhancement factor
to quantitatively represent the ratio of the reaction rate of the amine systems relative to that
only involves physical absorption. It was concluded that replacing 2 mass%, 5 mass% and 8
mass% of AMP with PZ while maintaining the total amine concentration at 30 mass% led to
increases in reaction rate by 2.3, 3.6 and 4.6 respectively. With the information in Table 7.1,
we can simply estimate that the reaction rate of 22 mass% AMP and 8 mass% PZ ( 3575)
has been raised to over half of that for 30 mass% MEA ( 6000) and is twice of that for DEA.
Samanta and Bandyopadhyay [223] also discussed the reaction mechanisms of PZ-promoted
AMP solution. It was believed that the reaction between CO2 and PZ mainly involves the
formation of a zwitterions followed by the deprotonation of the zwitterions by a base to
produce PZ-carbamate and protonated base [77, 222, 293]. The strength of the base and its
concentration determine the contribution of the base to the deprotonation [294, 295].
Therefore, the presence of AMP as bulk solvent facilitates the deprotonation of PZ

7. Conclusions and Future Work

183

zwitterions; meanwhile, the fast reaction between CO2 and PZ enhances the overall reaction
of the blended system.
Although the absorption of CO2 with PZ activated AMP has been broadly studied, the
solubility of the system is not well covered. Only two papers [218, 296] have been found on
this system mainly at absorber conditions (< 373 K). In this work, we measured the CO2
solubility in two different blends, namely 25 mass% AMP + 5 mass% PZ and 20 mass%
AMP + 10 mass% PZ at both absorber and stripper conditions. Both of the mixtures showed a
reduced cyclic capacity compared to the AMP only system at equal mass concentration:
assuming the equilibrium-pressure of CO2 in the absorber is 10 kPa, replacing 5 mass% of
AMP with 5 mass% of PZ leads to the reduction in the cyclic mole-loading capacity by about
8%, compared to that of 30 mass% AMP. A further replacement of 5 mass% reduces the
capacity by another 4.5%, i.e. 12.5% smaller compared to that of 30 mass% AMP. However,
compared to 30 mass% MEA, both amine blends have significant advantages over it with
approximately twice of the cyclic mole-loading capacity of the primary amine.
Based on the results from this work and information collected from other literature sources on
the kinetics, resistance to degradation, etc. we can conclude that AMP+PZ is a very
promising candidate as an alternative to conventional MEA to improve the process
economics. To fully quantify the potential of the process, a reliable process model is essential
[15].
7.2.2 CO2 absorption in 30 mass% aqueous DMMEA and DMMEA+PZ blends
The CO2 solubility measurements from this work for the 30 mass% aqueous DMMEA show
that its cyclic capacity is similar to that of 30 mass% aqueous AMP solution and is much
larger than the loading capacity of the MEA system (for mole-loading, the cyclic capacity of
DMMEA is between 90% and 120% larger than MEA; while for mass-loading, the former is
1.25 to 1.47 times of the latter).
As a tertiary amine, DMMEA does not react with CO2 to form carbamate; instead, it acts as a
catalyst and increase the basicity of the solvent. From Table 7.1, we can notice that the
reaction rate of DMMEA is even slower than AMP, although it is much higher than MDEA.
Henni et al. [233] studied the reaction kinetics of DMMEA using the stopped-flow technique
between 298 K and 313 K. They found that the reaction between DMMEA and CO2 can be
well explained by the base-catalysed mechanism and represented with a pseudo-first-order
rate expression.

7. Conclusions and Future Work

184

Apparently, in order to make DMMEA a commercially attractive candidate, the reaction rate
of the system needs to be greatly enhanced. In our work, PZ was again chosen as the activator.
Only one study conducted by Brder and Svendsen [244] has investigated the blended amine
system of DMMEA and PZ, along with several other mixed amines, in order identify the
most promising candidates. It was concluded from their work that DMMEA and AMP are the
best bulk amines while MAPA and PZ are both good candidates for activators and perform
better than MEA. Comparatively, the PZ-promoted MDEA system has been well researched
[99, 220, 224, 235, 257, 295]. Mamun et al. [274] compared the amine mixtures of
MDEA+PZ, MDEA+MEA, and MDEA+AEEA, of which the MDEA+PZ blend showed the
highest absorption rate at the same total amine concentration. Puxty and Rowland [224] built
a diffusion and chemical model for AMP+PZ and MDEA+PZ and compared the absorption
performance of the two systems. It was concluded that both AMP and MDEA serve as a pH
buffer and maintain the free PZ level in the solution and enhance the mass transfer. AMP is
more effective than MDEA owing to its higher pH value. In this work, we also measured the
pH of some selected amines as shown in Fig. 7.1 At 296 K, the pHs for both 30 mass% AMP
and DMMEA are approximately 12 while MDEA is a poorer base with pH around 11.5.
Compared with 5 mass% PZ, 30 mass% MDEA almost has a similar pH. This is not desirable
as the contribution of the bases to the deprotonation of PZ-carbamate was believed to be
relied upon their base strength. A less effective buffer results in rapid depletion of PZ and the
disappearance of the rate enhancement effect with the increase in solution loading.
12.6
12.4
12.2

pH

12
AMP

11.8

DMMEA

11.6

MDEA

11.4

PZ

11.2
11
0%

10%

20%

30%

40%

mass% of amine

Figure 7.1 pH of AMP, DMMEA, MDEA and PZ at 296 K

50%

60%

7. Conclusions and Future Work

185

The solubility measurements from this work showed similar cyclic capacity for DMMEA and
DMMEA+PZ blends. This is an important merit as the blended system has significantly
improved reaction kinetics compared to DMMEA alone. A further cross comparison also
showed us similar cyclic capacity of AMP+PZ and DMMEA+PZ.
7.2.3 Relationship between the structure of amine and the solubility of CO2
From the CO2 solubility measurements of this work we can conclude that: both AMP as a
sterically-hindered amine and DMMEA as a tertiary amine have more than twice the cyclic
capacity of that for MEA at the same mass concentration; Although replacing equal-mass of
the AMP with PZ leads to a reduced cyclic capacity compared to AMP alone, the blended
systems still possess considerably larger loading capacity than MEA. DMMEA and its PZ
blend also showed approximately twice the cyclic mole-loading capacity of MEA at the same
mass concentration.
7.2.4 Conclusions from the synthetic apparatus
This work also proved the feasibility of applying a synthetic apparatus to measure the CO 2
solubility in amine solutions. Some complications hindered the reliability of the results,
especially the dead volume in the system and the requirement for accurate pressure
measurement below ambient pressures. All in all, the synthetic apparatus could serve as a
useful complement to the analytical apparatus, especially if these factors can be mitigated.
Sometimes, when the analytical determination of the amine concentration is difficult, e.g.
MDEA, with its high boiling point, is difficult to analyse using GC, the synthetic method will
be useful [74].
7.3 Future work
As extensions to this work, two aspects are worth investigating: one is to continue measuring
the CO2 solubility of more potential amines or amine blends, as accurate CO2 solubility
information is important not only for amine selection, but also for process design; the other
aspect involves a better understanding of other important properties that will affect the
economics of the capture process, notably the reaction kinetics.
Taking into account both factors, we propose that a more systematic amine selection process
which assesses both the kinetic and thermodynamic properties could be carried out similar to
that which was employed in the work of Mamun et al. [274]: in the first place, an amine
selection process could be conducted in a screening apparatus to semi-quantitatively evaluate

7. Conclusions and Future Work

186

the absorption rate and capacity of a wide range of amines; subsequently, a more accurate and
reliable solubility measurements would be carried out in the static-analytical apparatus built
in this work.
The static-analytical apparatus built in this work could be upgraded to allow measurement at
a wider range of pressure conditions. This can be relatively conveniently done by fitting a
ROLSI valve which has been successfully implemented to amine systems in previous
studies [74, 297, 298]. Automatic sampling valves including GSV and LSV were initially
considered and used in this work; however, due to the adsorptive nature of amines and the
formation of carbamate in some amines, the approach was not successful (details in Chapter
3). The ROLSI valve [299] is believed to be able to overcome this issue by avoiding any
cold section in the sampling system. However, to allow sampling with the ROLSI valve,
the equilibrium system pressure needs to be at least 2 to 3 bar; thus the addition of an inert
gas to increase the system total pressure is unavoidable.
In this work, we evaluated the CO2 solubility in AMP, which is a sterically-hindered amine,
and DMMEA, which is a tertiary amine. Both sterically-hindered and tertiary amines are
believed to offer significantly greater loading capacity than primary or secondary amines due
to the formation of bicarbonate/carbonate (stoichiometry for amine: CO 2 is 1:1) instead of
carbamate (the stoichiometry for amine: CO2 is 2:1). The results from this work confirmed
this through measurements of the CO2 solubility at absorber (313 K to 353 K) and stripper
conditions (373 K to 393 K). A standardised approach was used to compare the solubility of
different amine system as detailed in Chapter 4. Besides, blended amine system, mainly PZ
activated AMP or DMMEA mixtures were also investigated for their CO2 absorption capacity.
The reason to choose PZ as the activator in this work is because it has been identified as an
effective promoter to other amines [217]. In the future work, it will be interesting to
investigate a wider range of promoters which have been identified as superior rate-enhancing
additives by previous studies, e.g. MAPA [244]. Moreover, the screening experiments could
also be used to guide the selection of suitable amine blends as the amines with fast kinetics
often would also be a good promoter to other amines. One general rule for the selection of
suitable amine blends would possibly entail a bulk amine possessing high cyclic capacity
together with an activator with fast reaction kinetics. In the meantime, resistance to
degradation, low corrosion to equipment, and low vapour pressure are also desirable
properties for the amine and amine blends, but are beyond the scope of this work.

7. Conclusions and Future Work

187

The collaboration with the Molecular Systems Engineering group also at Imperial College
London on the development of promising amine blends is also a topic worth considering in
the future work. On the one hand, despite the significantly reduced adjustable parameters
required in the SAFT-VR model compared to the industrially prevalent - approach (or
quasi-chemical model), experimental data are still essential for the regression of these
parameters and to validate the model. On the other hand, the predictability of the SAFT-VR
model can be used to guide the experimental measurements especially in terms of the
selection of amine blends.

8. Nomenclature

Chapter 8

Nomenclature
activity of water
AMP

2-amino-2-methyl-1-propanol

AGR

acid gas removal

average absolute deviation

second virial coefficient


pure component virial coefficient
cross virial coefficient
CLC

chemical looping cycle

DMMEA

2-Dimethylaminoethanol
response factor

FID

flame ionization detector

FTIR

fourier transform infrared spectroscopy

188

8. Nomenclature
GC

gas chromatography

GSV

gas sampling valve


Henrys law constant for CO2 in water
ionic strength of the solution

IGCC

Integrated Gasification Combined Cycle


equilibrium constant

LSV

liquid sampling valve


molality of species i (in molkg-1)

MDEA

N-Methyldiethanolamine

MEA

monoethanolamine
molar mass of water
total pressure
partial pressure of component i

PZ

piperazine
ideal gas constant (= 8.314)

RD

relative deviation or repeatability of measurements

SAFT-VR

Statistical Associating Fluid Theory Variable Range

SS

stainless steel
temperature (in K)

TCD

thermal conductivity detector


partial molal volume of component i
partial molal volume of CO2 in water at infinite dilution

189

8. Nomenclature
x

liquid mole fraction


mole fractions of components i in vapour phase
ionic charge of component i

ZEC

Zero Emission Carbon

Greek symbols

loading (mol CO2/ mol amine)


average loading

ij

binary interaction parameter of species i and j

activity coefficient

fugacity coefficient

density

Subscripts and superscripts


c

critical

saturation

water

infinite dilution

liquid phase

initial

sol

solvent

190

8. Nomenclature

191

Bibliography
[1] IPCC (2001) Working Group I: The Scientific Basis. Switzerland
[2] IPCC (2007) Fourth Assessment Report (AR4) Climate Change 2007: The Physical
Science Basis. UK
[3] International Energy Agency (2010) World energy outlook 2010 Edition. London,
Organisation for Economic Co-operation and Development.
[4] International Energy Agency (2010) Energy technology perspectives 2010 Edition.
[Online] London, Organisation for Economic Co-operation and Development. Available
from: http://www.iea.org/techno/etp/index.asp [Accessed 10th January 2012].
[5] U.S. Energy Information Administration (2011) International Energy outlook. [Online]
Report Number: DOE/EIA-048. Available from:
http://www.eia.gov/forecasts/ieo/index.cfm [Accessed 10th January 2012].
[6] World Energy Council (2007) Transport Technologies and Policy Scenarios to 2050.
[Online] UK. Available from:
http://www.worldenergy.org/documents/transportation_study_final_online.pdf [Accessed
10th January 2012].
[7] Zhou, N., Fridley, D., McNeil, M., Zheng, N., Ke, J., and Levine, M. (2011) Chinas
Energy and Carbon Emissions Outlook to 2050. Lawrence Berkeley National Laboratory.
[Online] Report number: LBNL-4472E. Available from:
http://china.lbl.gov/publications/Energy-and-Carbon-Emissions-Outlook-of-China-in-2050
[Accessed 10th January 2012].

Bibliography

192

[8] Gambhir, A., Hirst, N., Brown, T., Riahi, K., Schulz, N., Faist, M., Foster, S., Jennings,
M., Munuera, L., Tong, D., and Tse, L (2012) China's energy technologies to 2050.
[Online] Report number: Grantham Report GR2. Available from:
http://www3.imperial.ac.uk/climatechange/publications/institutepublications [Accessed
10th January 2012].
[9] Hirst, N., Dunnett, A., Faist, M., Foster, S., Jennings, M., Munuera, L., and Tong, D.
(2011) An assessment of China's 2020 carbon intensity target. [Online] Report number:
Grantham Report GR1. Available from:
http://www3.imperial.ac.uk/climatechange/publications/institutepublications [Accessed
10th January 2012].
[10] International Energy Agency (2009) Technology Roadmaps Carbon Capture and
storage. [Online] France. Available from:
http://www.iea.org/papers/2009/CCS_Roadmap.pdf [Accessed 16th January 2012].
[11] Global CCS Institute (2011) Global Status of Large-Scale Integrated CCS Projects December 2011 Update. [Online] Australia. Available from:
http://www.globalccsinstitute.com/publications/global-status-large-scale-integrated-ccsprojects-december-2011-update [Accessed 16th January 2012].
[12] Global CCS Institute (2011) The Global Status of CCS: 2011. [Online] Australia.
Available from:
http://cdn.globalccsinstitute.com/sites/default/files/publications/22562/global-status-ccs2011.pdf [Accessed 16th January 2012].
[13] Bottoms, R. R. (1930) Separating acid gases. U. S. Patent 1783901.
[14] Department of Energy (2007) Carbon Dioxide Capture from Existing Coal-Fired Power
Plants. [Online] Report number: DOE/NETL-401/110707. Available from:
http://www.netl.doe.gov/energyanalyses/pubs/CO2%20Retrofit%20From%20Existing%20Plants%20Revised%20Novem
ber%202007.pdf [Accessed 16th January 2012].
[15] MacDowell, N. (2010) The integration of advanced molecular thermodynamics and
process modelling for the design of amine-based CO2 capture processes. PhD Diss.,
Imperial College London, London.

Bibliography

193

[16] Kent, R.L., and Eisenberg, B. (1976) Better data for amine treating. Hydrocarbon
Process, 55 (2), 87-90.
[17] Carroll, J. J., Slupsky, J. D. and Mather, A. E. (1991) The Solubility of Carbon Dioxide
in Water at Low Pressure. J. Phys. Chem. Ref. Data, 20, 1201-1209.
[18] Crovetto, R. and Mather, A. E. (1994) Evaluation of Solubility Data of the System CO2
+ H2O from 273 K to the critical Point of Water, in Carbon Dioxide in Water and
Aqueous Electrolyte Solutions, IUPAC Solubility Series, Vol 62, Scharlin, P. (ed),
Oxford University Press, Oxford, 5-12.
[19] Guggenheim, E.A. and Stokes, R.H. (1958) Activity coefficients of 2:1 and 1:2
electrolytes in aqueous solution from isopiestic data. Trans. Faraday Soc., 54, 1646-1649.
[20] Florin, N. and Fennell, P. (2010) Carbon capture technology: future fossil fuel use and
mitigating climate change. [Online] Grantham Institute for Climate Change. Report
number: Briefing paper No 3. Available from:
http://www3.imperial.ac.uk/climatechange/publications [Accessed 16th January 2012].
[21] Rochelle, G. T. (2009) Amine Scrubbing for CO2 Capture. Science, 325, 1652-1654.
[22] Global CCS Institute (2012) CO2 Capture Technologies: Post Combustion Capture
(PCC). [Online] Australia. Available from:
http://www.globalccsinstitute.com/publications/co2-capture-technologies-postcombustion-capture-pcc [Accessed 18th January 2012].
[23] Fluor. Available: http://www.fluor.com/econamine/Pages/default.aspx [Accessed 10th
March 2012].
[24] Mitchell, R. (2008) Mitsubishi Heavy Industries carbon capture technology. Carbon
Capture Journal. [Online] 1, 2-5. Available from:
http://www.carboncapturejournal.com/displaynews.php?NewsID=97 [Accessed 10th
March 2012].
[25] Alstom. Available: http://www.alstom.com/power/news-and-events/pressreleases/alstom-chilled-ammonia-process-selected-for-leading-cO2-capture-plant-inromania/ [Accessed 10th March 2012]
[26] Brouwer, J.P., Feron, P.H.M. and Asbroek, N.A.M. ten (2005) Amino-acid salts for
CO2 capture from flue gases. Fourth Annual Conference on Carbon Capture &
Sequestration, Alexandria Virginia, 5th May, USA.

Bibliography

194

[27] Olajire, A. (2010) CO2 capture and separation technologies for end-of-pipe applications
A review. Energy, 35, 2610-2628.
[28] Davison J. and Thambimuthu, K. (2009) An overview of technologies and costs of
carbon capture in power generation. Proceeding of the Institution of Mechanical
Engineers, part B, Journal of engineering manufacture, 223 (3), 201212
[29] Global CCS Institute (2012) CO2 Capture Technologies: Pre Combustion Capture.
[Online] Australia. Available from: http://www.globalccsinstitute.com/publications/co2capture-technologies-pre-combustion-capture [Accessed 10th March 2012].
[30] Texas Clean Energy Project (TCE) available from:
http://sequestration.mit.edu/tools/projects/tcep.html [Accessed 10th March 2012].
[31] Hatfield information available from:
http://powerassetmodelling.co.uk/html/hatfield_igcc_.html [Accessed 10th March 2012].
[32] Greengen information available from:
http://sequestration.mit.edu/tools/projects/greengen.html [Accessed 10th March 2012].
[33] ZeroGen project information available from:
http://www.zerogen.com.au/project/overview.aspx [Accessed 10th March 2012].
[34] Weyburn-Midale project available from: http://www.ptrc.ca/weyburn_overview.php
[Accessed 10th March 2012].
[35] Enid Fertilizer project available from: http://www.globalccsinstitute.com/projects/12561
[Accessed 10th March 2012].
[36] Lee, A. (2011) High temperature adsorption materials and their performance for precombustion capture of carbon dioxide. Energy Procedia, 4, 1199-1206.
[37] Global CCS Institute (2012) CO2 Capture Technologies: Oxy Combustion with CO2
Capture. [Online] Australia. Available from:
http://www.globalccsinstitute.com/publications/co2-capture-technologies-oxycombustion-co2-capture [Accessed 10th March 2012].
[38] Vattenfall (2009) Newsletter on Carbon Capture and Storage at Vattenfall, Bridging to
the Future. [Online] Report number: No. 14. Available from:
http://www.vattenfall.com/en/ccs/file/Bridging_to_the_future_May_2011_18032681.pdf
[Accessed 10th March 2012].

Bibliography

195

[39] Lacq CCS Pilot Plant: Operational Feedback of the Surface Facilities One Year after
Start-up (2011) IEAGHG 2nd Oxy-Fuel Combustion Conference. 12-16 September,
Yeppoon, Queensland, Australia.
[40] Farley M. BCURA Coal Science lecture (2008). Available from:
http://www.bcura.org/csl08.pdf [Accessed 10th March 2012].
[41] CIUDEN Oxy-CFB Boiler Demonstration Project (2011) The 36th International
Technical Conference on Clean Coal & Fuel Systems. 5-9th June, Clearwater, Florida
USA. Available from:
http://www.fwc.com/publications/tech_papers/files/TP_CCS_11_04.pdf [Accessed 10th
March 2012].
[42] Callide Coal project. Available from: http://www.powertechnology.com/projects/callide-coal/ [Accessed 10th March 2012].
[43] FuturGen information available from:
http://www.scientificamerican.com/article.cfm?id=back-to-the-future-with-futuregencoal-plant [Accessed 10th March 2012].
[44] Pires, J., Martins, F. G., Alvim-Ferraz, M. C. M., Alvim-Ferraz, M. (2011) Recent
developments on carbon capture and storage: An overview. Chemical Engineering
Research and Design, 89, 1446-1460.
[45] MacDowell, N., Florin, N., Buchard, A., Hallett, J., Galindo, A., Jackson, G., Adjiman,
C. S., Williams, C. K., Shah, N. and Fennell, P. (2010) An overview of CO2 capture
technologies. Energy & Environmental Science, 3, 1645-1669.
[46] Anthony, E. J. (2011) Ca looping technology: current status, developments and future
directions. Greenhouse Gases: Science and Technology, 1, 36-47.
[47] Gao, L., Paterson, N., Fennell, P., Dugwell, D. and Kandiyoti, R. (2008) The Zero
Emission Carbon Concept (ZECA): Extents of Reaction with Different Coals in
Steam/Hydrogen, Tar Formation and Residual Char Reactivity. Energy & Fuels, 22,
2504-2511.
[48] Hossain, M. and de Lasa, H. I. (2008) Chemical-looping combustion (CLC) for inherent
CO2 separationsa review. Chemical Engineering Science, 63, 4433-4451.

Bibliography

196

[49] Alstom information available from:


http://www.netl.doe.gov/technologies/coalpower/ewr/co2/oxycombustion/prototype.html [Accessed 10th March 2012].
[50] Sayari, A., Belmabkhout, Y. and Serna-Guerrero, R. (2011) Flue gas treatment via CO2
adsorption. Chemical Engineering Journal, 171, 760-774.
[51] Xu, X., Song, C., Andrsen, J. M., Miller, B. G. and Scaroni, A. W. (2003) Preparation
and characterization of novel CO2 molecular basket adsorbents based on polymermodified mesoporous molecular sieve MCM-41. Microporous and mesoporous materials,
62, 29-45.
[52] Xu, X., Song, C., Andersn, J.M., Miller, B.G. and Scaroni, A.W. (2002) Novel
polyethylenimine-modified mesoporous molecular sieve of MCM-41 type as highcapacity adsorbent for CO2 capture, Energy & Fuels, 16, 14631469.
[53] Xu, X., Song, C., Miller, B.G. and Scaroni, A.W. (2005) Influence of moisture on CO2
separation from gas mixture by a nanoporous adsorbent based on polyethyleniminemodified molecular sieve MCM-41. Ind. Eng. Chem. Res., 44, 81138119.
[54] Ma, X., Wang, X. and Song, C. (2009) Molecular basket sorbent for separation of CO2
and H2S from various gas streams, J. Am. Chem. Soc., 131, 57775783.
[55] Jadhav, P. D., Chatti, R. V., Biniwale, R. B., Labhsetwar, N. K., Devotta, S. and Rayalu,
S. S. (2007) Monoethanol amine modified zeolite 13X for CO2 adsorption at different
temperatures, Energy & Fuels, 21, 35553559.
[56] Fisher, J. C., Tanthana, J. and Chuang, S. S. C. (2009) Oxide-supported
tetraethylenepentamine for CO2 capture. AIChE J., 28, 589598.
[57] Leal, O., Bolivar, C., Ovalles, C., Garcia, J. J. and Espidel, Y. (1995) Reversible
adsorption of carbon dioxide on amine surface-bonded silica gel. Inorg. Chim. Acta, 240,
183189.
[58] Huang, H. Y., Yang, R. T., Chinn, D. and Munson, C. L. (2003) Amine-grafted MCM48 and silica xerogel as superior sorbents for acidic gas removal from natural gas, Ind.
Eng. Chem. Res., 42, 24272433.
[59] Knowles, G. P., Graham, J. V., Delaney, S. W. and Chaffee, A. L. (2005)
Aminopropylfunctionalized mesoporous silica as CO2 adsorbents. Fuel Process. Technol.,
86, 14351448.

Bibliography

197

[60] Liu, N., Assink, R. A., Smarsly, B. and Brinker, C. J. (2003) Synthesis and
characterization of highly ordered functional mesoporous silica thin films with positively
chargeable -NH2 groups. Chemical Communications, 10, 1146-1147.
[61] Isaacs, E. E., Otto, F. D. and Mather A. E. (1980) Solubility of mixture of hydrogen
sulfide and carbon dioxide in a monoethanolamine solution at low partial pressures. J.
Chem. Eng. Data, 25 (2), 118-120.
[62] Huang, Y., Soriano, A. N., Caparanga, A. R. and Li, M. (2011) Kinetics of absorption of
carbon dioxide in 2-amino-2-methyl-l-propanol + N-methyldiethanolamine + water.
Journal of the Taiwan Institute of Chemical Engineers, 42, 76-85.
[63] Chakravarty, T., Phukan, U. K. and Weilund, R. H. (1985) Reaction of acid gases with
mixtures of amines. Chem. Eng. Prog., 81, 32-36.
[64] Sartori, G. and Savage, D. W. (1983) Sterically hindered amines for carbon dioxide
removal from gases. Ind. Eng. Chem. Fundamen., 22 (2), 239-249.
[65] Chakma, A. and Meisen, A. (1987) Solubility of carbon dioxide in aqueous
methyldiethanolamine and N,N-bis(hydroxyethyl)piperazine solutions. Ind. Eng. Chem.
Res., 26 (12), 2461-2466.
[66] Baek, J. and Yoon, J. (1998) Solubility of carbon dioxide in aqueous solutions of 2amino-2-methyl-1, 3-propanediol. J. Chem. Eng. Data, 43, 4, 635-637.
[67] Mamun, S., Jakobsen, J. P. and Svendsen, H. F. (2006) Experimental and Modeling
Study of the Solubility of Carbon Dioxide in Aqueous 30 Mass % 2-((2Aminoethyl)amino)ethanol Solution. Ind. Eng. Chem. Res., 45, 8, 2505-2512.
[68] Vaidya, P. D. and Kenig, E. Y. (2007) Absorption of CO2 into aqueous blends of
alkanolamines prepared from renewable resources. Chem. Eng. Sci., 62, 24, 7344-7350.
[69] Hla, E., Pick, J., Fried, V. and Vilim, O. (1957) Vapour-liquid equilibrium. United
States, Pergamon Press.
[70] Mamun, S., Nilsen, R., and Juliussen, H.F.S. (2005) Solubility of Carbon Dioxide in 30
mass % Monoethanolamine and 50 mass % Methyldiethanolamine Solutions. J. Chem.
Eng. Data, 50, 630-634.
[71] Silkenbumer, D., Rumpf, B. and Lichtenthaler, R.N. (1998) Solubility of Carbon
Dioxide in Aqueous Solutions of 2-Amino-2-Methyl-1-Propanol and N-

Bibliography

198

Methyldiethanolamine and their Mixtures in the Temperature Range from 313 to 353 K
and Pressure up to 2.7 MPa. Ind. Eng. Chem. Res., 37, 3133-3141.
[72] Li, M. H. and Chang, B. C. (1994) Solubilities of Carbon Dioxide in Water +
Monoethanolamine + 2-Amino-2-methyl-1-propanol. J. Chem. Eng. Data, 39, 448-452.
[73] Jou, F. Y., Mather, A. E., and Otto, F. D. (1995) The solubility of CO2 in a 30 mass
percent monoethanolamine solution. The Canadian journal of chemical engineering, 73
(1), 140-147.
[74] Sidi-Boumedine, R. Horstmann, S., Fischer, K., Provost, E., Frst, W. and Gmehling, J.
(2004) Experimental determination of carbon dioxide solubility data in aqueous
alkanolamine solutions. Fluid phase equilibria, 218(1), 85-94.
[75] Kuranov, G., Rumpf, B., Smirnova, N.A. and Maurer, G. (1996) Solubility of Single
Gases Carbon Dioxide and Hydrogen Sulfide in Aqueous Solutions of NMethyldiethanolamine in the Temperature Range 313-413K at Pressures up to 5 MPa.
Ind. Eng. Chem. Res., 35, 1959-1966.
[76] Dicko, M., Coquelet, C., Jarne, C., Northrop, S., and Richon, D. (2010) Acid gases
partial pressures above 50 wt% aqueous methyldiethanolamine solution: Experimental
work and modelling. Fluid phase equilibria, 289, 99-109.
[77] Bishnoi, S. and Rochelle, G. T. (2000) Absorption of carbon dioxide into aqueous
piperazine: reaction kinetics, mass transfer and solubility. Chem. Eng. Sci., 55, 55315543.
[78] Mondal, M. K. (2009) Solubility of Carbon Dioxide in an Aqueous Blend of
Diethanolamine and Piperazine. J. Chem. Eng. Data., 54, 9, 2381-2385.
[79] Abu-Arabi (1988) Physical Solubility of Hydrogen Sulfide and Carbon Dioxide in
Alkanolamine Solutions. PhD Diss., Oklahoma State Univ., Stillwater.
[80] Bhairi, A. M. (1984) Experimental equilibrium between acid gases and ethanolamine
solutions. PhD Diss., Oklahoma State Univ., Stillwater.
[81] Dingman, J. C., Jackson, J. L., Moore, T. F. and Branson J. A. (1983) Equilibrium Data
for the H2S-CO2-Diglycolamine Agent-Water System. Proc. 62nd Ann. Gas Proc. Assoc.
Convention, 256-268.

Bibliography

199

[82] Jones, J. H., Froning, H. R. and Claytor Jr., E. E. (1959) Solubility of Acidic Gases in
Aqueous Monoethanolamine. J. Chem. Eng. Data, 4, 1, 85-92.
[83] Lawson, J. D. and Garst, A. W. (1976) Gas Sweetening Data: Equilibrium Solubility of
Hydrogen Sulfide and Carbon Dioxide in Aqueous Monoethanolamine and Aqueous
Diethanolamine Solutions. J. of Chem. Eng. Data, 21, 1, 20-30.
[84] Maddox, R. N., Bhairi, A. H., Diers, J. R. and Thomas, P.A. (1987) Equilibrium
Solubility of Carbon Dioxide or Hydrogen Sulfide in Aqueous Solutions of
Monoethanolamine, Diglycolamine, Diethanolamine and Methyldiethanolamine. Report
number: RR-104, Gas Processors Association, Tulsa, OK.
[85] Muhlbauer, H. G. and Monaghan, P. R. (1957) Sweetening natural gas with
ethanolamine solutions. Oil Gas J., 55, 139-145.
[86] Scott, W. W. (1939) Standard Methods of Chemical Analysis, 5th ed. New York. Vol. 1,
PP 235-237.
[87] Dawodu, O. F. and Meisen, A. (1996) Degradation of alkanolamine blends by CO2. Can.
J. Chem. Eng. 74, 960-966.
[88] Haji-Sulaiman, M. Z. and Aroua, M. K. (1994) Equilibrium of CO2 in Aqueous
Diethanolamine (DEA) and Amono-methyl-propanol (AMP) Solutions. Chem. Eng.
Commun. 140, 1, 157-171.
[89] Huttenhuis et al. (2009) Solubility of Carbon Dioxide and Hydrogen Sulfide in
Methyldiethanolamine Solutions. Ind. & Eng. Chem. Res., 48 (8), 4051-4059.
[90] Jenab, M. H., Abdi, M. A., Najibi, S. H., Vahidi, M. and Matin, N. S. (2005) Solubility
of carbon dioxide in aqueous mixtures of N-methyldiethanolamine + piperazine +
sulfolane. J. Chem. Eng. Data, 50, 2, 583-586.
[91] Kennard, M. L. and Meisen, A. (1984) Solubility of Carbon Dioxide in Aqueous
Diethanolamine Solutions at Elevated Temperature and Pressures. J. Chem. Eng. Data,
29, 309.
[92] Kundu, M., Mandal, B. P. and Bandyopadhyay, S. S. (2003) Vapour-Liquid Equilibria of
CO2 in Aqueous Solutions of 2-Amino-2-Methyl-1-Propanol. J. Chem. Eng. Data, 48,
789-796.

Bibliography

200

[93] Lee, J. I., Otto F. D. and Mather A. E. (1972) Solubility of carbon dioxide in aqueous
diethanolamine solutions at high pressures. J. Chem. Eng. Data, 17, 465-468.
[94] Park, M. K. and Sandall, O. C. (2001) Solubility of Carbon Dioxide and Nitrous Oxide
in 50 mass Methyldiethanolamine. J. Chem. Eng. Data, 46 (1), 166168.
[95] Park, S. H., Lee, K. B., Hyun, J. C. and Kim, S. H. (2002) Correlation and Prediction of
the Solubility of Carbon Dioxide in Aqueous Alkanolamine and Mixed Alkanolamine
Solutions. Ind. Eng. Chem. Res., 41, 16581665.
[96] Reed, R. M. and Wood, W. R. (1941) Recent design developments in amine gas
purification plants. Trans. AIChE., 37, 363-383.
[97] Saha, A. K., Bandyopadhyay, S. S. and Biswas, A. K. (1993) Solubility and diffusivity
of N2O and CO2 in aqueous solutions of 2-amino-2-methyl-1-propanol. J. Chem. Eng.
Data, 38, 7882.
[98] Shen, K. P. and Li, M. H. (1992) Solubility of carbon dioxide in aqueous mixtures of
monoethanolamine with methyldiethanolamine. J. Chem. Eng. Data, 37, 96-100.
[99] Xu, G. W., Zhang, C. F., Qin, S. J., Gao, H. W. and Liu, H. B. (1998) GasLiquid
Equilibrium in a CO2MDEAH2O System and the Effect of Piperazine on It. Ind. Eng.
Chem. Res. 37, 14731477.
[100] Horstmann, S., Birtke, G. and Fischer, K. (2004) Vapour-liquid equilibrium and excess
enthalpy data for the binary systems propane + dimethyl ether and propene + dimethyl
ether at temperatures from (298 to 323) K. J. Chem. Eng. Data, 49, 38-42.
[101] Isaacs, Otto & Mather (1980) Solubility of mixture of hydrogen sulfide and carbon
dioxide in a monoethanolamine solution at low partial pressures. J. Chem. Eng. Data, 25
(2), 118-120.
[102] Jou, F. Y., Carroll, J. J., Mather, A. E. and Otto F. D. (1993) The solubility of carbon
dioxide and hydrogen sulfide in a 35 wt% aqueous solution of methyldiethanolamine.
Can. J. Chem. Eng., 71, 264-268.
[103] Lee, J. I., Otto F. D. and Mather A. E. (1976) Equilibrium between carbon dioxide and
aqueous monoethanolamine solutions. J. Appl. Chem. Biotechnol, 26, 541-549.
[104] Martin, J. L., Otto, F. D. and Mather, A. E. (1978) Solubility of hydrogen sulphide and
carbon dioxide in a diglycolamine solution. J. Chem. Eng. Data, 23, 163-164.

Bibliography

201

[105] Teng, T. T. and Mather A. E. (1989) Solubility of H2S, CO2 and their mixtures in an
AMP solution. Can. J. Chem. Eng., 67, 846-850.
[106] Jane, I. S. and Li, M. H. (1997) Solubilities of mixtures of carbon dioxide and hydrogen
sulfide in water + diethanolamine + 2-amine-2-methyl-1-propanol. J. Chem. Eng. Data,
42 (6), 98-105.
[107] Lee, J. I., Otto F. D. and Mather A. E. (1974) The solubility of H2S and CO2 in aqueous
monoethanolamine solutions. Can. J. Chem. Eng., 52, 803-805.
[108] Rho, S. W., Yoo, K. P., Lee, J. S., Nam, S. C., Son, J. E. and Min B. M. (1997)
Solubility of CO2 in aqueous methyldiethanolamine solutions. J. Chem. Eng. Data, 42
(6), 1161-1164.
[109] Seo, D.J. and Hong, D.J. (1996) Solubilities of Carbon Dioxide in Aqueous Mixtures of
Diethanolamine and 2-Amino-2-methyl-1-Propanol. J. Chem. Eng. Data, 41, 258-260.
[110] Frazier, R. E. (1993) Acid Gas-Diethanolamine Vapor-Liquid Equilibrium Data by
Fourier Transform Infrared Spectroscopy. PhD Diss., Texas A&M Univ., College
Station.
[111] Hou, S. X, Duan, Y. Y. and Wang, X. D. (2007) Vapor-liquid equilibria predictions for
new refrigerant mixtures based on group contribution theory. Ind. Eng. Chem. Res.,
46(26), 9274-9284.
[112] Rogers, W. J., Bullin, J. A., Davison, R. R., Frazier, R. E. and Marsh, K. N. (1998)
FTIR methodfor VLE measurements of acid-gas-alkanolamine systems.
Thermodynamics, 43(12), 3223-3231.
[113] Rebolledo-Libreros, M. E. and Trejo, A. (2004) Gas solubility of CO2 in aqueous
solutions of N-Methyldiethanolamine and diethanolamine with 2-amino-2-methyl-1propanol. Fluid phase equilibria, 218 (2), 261-267.
[114] Ruska, W. E. A., Hurt, L. J. and Kobayashi, R. (1970) Circulating pump for high
pressure and -200 to +400 C application. Rev. Sci. Instrum. 41(10), 1444-1446.
[115] Prausnitz, J. M., Lichtenthaler, R. N. and de Azevedo, E. G. (1986) Molecular
Thermodynamics of Fluid-Phase Equilibria, 2nd ed.; Prentice-Hall: Englewood Cliffs,
NJ.

Bibliography

202

[116] Bougie, F. and Iliuta, M. C. (2010) CO2 absorption into mixed aqueous solutions of 2amino-2-hydroxymethyl-1,3-propanediol and piperazine. Ind. Eng. Chem. Res., 49 (3), 11501159.
[117] Atwood, K. Arnold, M. R. and Kindrick, R. C. (1957) Equilibria for the system
ethanolamine-hydrogen sulfidewater. Ind. Eng. Chem. Res., 49(9), 14391444.
[118] Isaacs, E. E., Otto, F. D. and Mather, A. E. (1980) Solubility of mixture of hydrogen
sulfide and carbon dioxide in a monoethanolamine solution at low partial pressures. J.
Chem. Eng. Data, 25 (2), 118-120.
[119] Jou, F. Y., Mather, A. E. and Otto, F. D. (1982) Solubility of H2S and CO2 in aqueous
methyldiethanolamine solutions. Ind. Eng. Chem. Process Des. DeV., 21, 539-544.
[120] Lal, D., Otto, F. D. and Mather, A. E. (1985) The solubility of H2S and CO2 in a
diethanolamine solution at low partial pressures. Can. J. Chem. Eng., 63, 681-685.
[121] Leibush A. G. and Shneerson, A. L. (1950) The absorption of hydrogen sulphide and of
its mixtures with carbon dioxide by ethanolamines. J. Appl. Chem. (Russ.), 23, 149-157.
[122] Mason, J. W. and Dodge, B. F. (1936) Equilibrium absorption of carbon dioxide by
solution of ethanolamines. AIChE J., 32, 27-48.
[123] Nasir, P. and Mather, A. E. (1977) The Measurement and Prediction of the Solubility of
Acid Gases in Monoethanolamine Solutions at Low Partial Pressures, The Can. J.
Chem. Eng., 55(6), 715-717.
[124] Riegger, E. Tartar, H. V. and Lingafelter, E. C. (1944) Equilibria between hydrogen
sulfide and aqueous solutions of monoethanolamine at 25 C,45 C and 60 C. J. Am.
Chem. Soc., 66, 2024-2027.
[125] Kim, I., Svendsen, H. F. and Borresen, E. (2008) Ebulliometric determination of
vapour-liquid equilibria for pure water, monoethanolamine, N-Methyldiethanolamine,
3-(Methylamino)-propylamine, and their binary and ternary solutions, J. Chem. Eng.
Data, 53(11), 2521-2531.
[126] Jensen, M. B., Jrgensen, E., Faurholt, C. and Srensen, N. A. (1954) Reactions
between Carbon Dioxide and Amino Alcohols. I. Monoethanolamine and
Diethanolamine. Acta Chem. Scand., 8(1), 1137-1140.

Bibliography

203

[127] Aroua, M. K. Benamor, A. and Haji-Sulaiman, M. Z. (1999) Equilibrium constant for


carbamate formation from monoethanolamine and its relationship with temperature. J.
Chem. Eng. Data, 44 (5) 887-891.
[128] Browning, G. J. and Weiland, R. H. (1994) Physical solubility of carbon dioxide in
aqueous alkanolamines via nitrous oxide analogy. J. Chem. Eng. Data, 39, 817822.
[129] Arcis, H., Rodier, L., Ballerat-Busserolles, K. and Coxam, J. Y. (2009) Enthalpy of
solution of CO2 in aqueous solution of methyldiethanolamine at T = 372.9 K and
pressures up to 5 MPa. J. Chem. Thermodyn., 41 (7), 836-841.
[130] Austgen, D. M., Rochelle, G., Peng, X. and Chen, C. C. (1989) Model of Vapor-Liquid
Equilibria for Aqueous Acid Gas-Alkanolamine Systems Using the Electrolyte-NRTL
Equation. Ind. Eng. Chem. Res., 28(7), 1060-1073.
[131] Benamor, A. and Aroua, M. K. (2005) Modeling of CO2 solubility and carbamate
concentration in DEA, MDEA and their mixtures using the DeshmukhMather model.
Fluid Phase Equilibria, 231(2), 150162.
[132] Danckwerts, P. V. and McNeil, K. M. (1967) The absorption of carbon dioxide into
aqueous amine solutions and the effects of catalysis, Trans. Instn. Chem. Engrs, 45,
T32T49.
[133] Deshmukh, R. D. and Mather, A. E. (1981) A Mathematical Model for Equilibrium
Solubility of Hydrogen Sulfide and Carbon Dioxide in Aqueous Alkanolamine
Solutions. Chem. Eng. Sci. 36(2), 355-362.
[134] Edwards, T. J., Maurer, G., Newman, J. and Prausnitz, J. M. (1978) Vapor-liquid
equilibria in multicomponent aqueous solutions of volatile weak electrolytes. AIChE J.
24(6), 966976.
[135] Gabrielsen, J., Michelsen, M. L., Stenby, E. H. and Kontogeorgis, G. M. (2005) A
model for estimating CO2 solubility in aqueous alkanolamines. Ind. Eng. Chem. Res. 44,
33483354.
[136] Kaewsichan, L., Al-Bofersen, O., Yesavage, V. F. and Selim, M. S. (2001) Prediction
of the solubility of acid gases in monoethanolamine (MEA) and methyldiethanolamine
(MDEA) solutions using the electrolyte-UNIQUAC model. Fluid Phase Equilibria,
183-184, 159-171.

Bibliography

204

[137] Kent, R. L. and Eisenberg, B. (1976) Better Data for Amine Treating. Hydro. Proc.,
55(2), 87-90.
[138] Li, C. X. and Frst, W. (2000) Representation of CO2 and H2S solubility in aqueous
MDEA solution using an electrolyte equation of state. Chem. Eng. Sci., 55 (15), 29752988.
[139] Kuranov, G., Rumpf, B., Maurer, G. and Smirnova, N. (1997) VLE modeling for
aqueous systems containing methyldiethanolamine, carbon dioxide and hydrogen
sulfide. Fluid Phase Equilibria, 136(1-2), 147-162.
[140] Liu, Y., Zhang, L. and Watanasiri, S. (1999) Representing Vapor-Liquid Equilibrium
for an Aqueous MEA-CO2 System Using the Electrolyte Nonrandom-Two-Liquid
Model. Ind. Eng. Chem. Res., 38 (5), 2080-2090.
[141] Schmidt, K. A. G., Maham, Y. and Mather, A. E. (2007) Use of the NRTL equation for
simultaneous correlation of vapour-liquid equilibria and excess enthalpy applications to
aqueous alkanolamine systems. Journal of Thermal Analysis and Calorimetry, 89 (1),
61-72.
[142] Vrachnos, A., Kontogeorgis, G. and Voutsas, E. (2006) Thermodynamic modeling of
acidic gas solubility in aqueous solution of MEA, MDEA and MEA-MDEA blends. Ind.
Eng. Chem. Res., 45(14), 5148-5154.
[143] Kritpiphat, W. and Tontiwachwuthikul, P. (1996) New modified Kent-Eisenberg model
for predicting carbon dioxide solubility in aqueous 2-amino-2-methyl-1-propanol
(AMP) solutions. Chem. Eng. Commun., 144, 77-83.
[144] Tontiwachwuthikul, P., Meisen, A. and Lim, C. J. (1991) Solubility of CO2 in 2Amino-2-methyl-1-propanol solutions. J. Chem. Eng. Data, 36(1), 130-133.
[145] Xu, S., Wang, Y. W., Otto, F. D. and Mather, A. E. (1992) Representation of
equilibrium solubility properties of CO2 with aqueous solutions of 2-amino-2-methyl-1propanol. Chem. Eng. Proc., 31, 7-12.
[146] Haji-Sulaiman, M., Aroua, M. K. and Benamor, A. (1998) Analysis of Equilibrium
Data of CO2 in Aqueous Solutions of Diethanolamine (DEA), Methyldiethanolamine
(MDEA) and Their Mixtures Using the Modified Kent Eisenberg Model. Chem. Eng.
Res. Des., 76 (8), 961-968.

Bibliography

205

[147] Hu, W. and Chakma, A. (1990) Modelling of equilibrium solubility of CO2 and H2S in
aqueous amino methyl propanol (AMP) solutions. Chem. Eng. Commun., 94, 53-61.
[148] Li, M. H. and Shen, K. P. (1993) Calculation of equilibrium solubility of carbon
dioxide in aqueous mixtures of monoethanolamine with methyldiethanolamine. Fluid
Phase Equilibria, 85, 129-140.
[149] Yang, Z. Y., Soriano, A. N., Caparanga, A. R. and Li, M. H. (2010) Equilibrium
solubility of carbon dioxide in (2-amino-2-methyl-1-propanol + piperazine + water). J.
Chem. Thermo., 42, 659-665.
[150] Park, S. H., Lee, K. B., Hyun, J. C. and Kim, S. H. (2002) Correlation and Prediction of
the Solubility of Carbon Dioxide in Aqueous Alkanolamine and Mixed Alkanolamine
Solutions. Ind. Eng. Chem. Res., 41, 1658-1665.
[151] Edwards, T. J., Maurer, G., Newman, J. and Prausnitz, J. M. (1978) Vapor-liquid
equilibria in multicomponent aqueous solutions of volatile weak electrolytes. AIChE J.
24(6), 966976.
[152] Nakamura, R., Breedveld, G. J. F. and Prausnitz, J. M. (1976) Thermodynamic
Properties of Gas Mixtures Containing Common Polar and Nonpolar Components. Ind.
Eng. Chem. Proc. Design Develop., 15, 557-564.
[153] Gokcen, N. A. (1996) Gibbs-Duhem-Margules laws. Journal of Phase Equilibria, 17(1),
50-51.
[154] Bromley, L. A. (1972) Approximate Individual Ion Values of (or B) in Extended
Debye-Hckel Theory for Uni-univalent Aqueous Solutions at 298.15 K. J. Chem.
Thermo., 4, 669-673.
[155] Edwards, T. J., Newman, J. and Prausnitz, J. M. (1975) Thermodynamics of Aqueous
Solutions Containing Volatile Weak Electrolytes. AIChE J., 21(2), 248-259.
[156] Pitzer, K. S. and Mayorga, G. (1973) Thermodynamics of Electrolytes, II. Activity and
osmotic coefficients for strong electrolytes with one or both ions univalent. J. Phys.
Chem., 77(19), 2300-2308.
[157] Arcis, H., Rodier, L., Ballerat-Busserolles, K. and Coxam, J. Y. (2009) Enthalpy of
solution of CO2 in aqueous solution of methyldiethanolamine at T = 372.9 K and
pressures up to 5 MPa. J. Chem. Thermodyn., 41 (7), 836-841.

Bibliography

206

[158] Posey, M. L. and Rochelle, G. T. (1997) A thermodynamic model of


methyldiethanolamine-CO2-H2S-water. Ind. Eng. Chem. Res., 36, 3944-3953.
[159] Brown, K. M. (1973) Numerical solution of nonlinear algebraic equations (edited by
Byrne G. D. and Hall C. A.), p. 281. Academic Press, New York.
[160] Peng, D. and Robinson, D. B. (1976) A new two-constant equation of state. Ind. Eng.
Chem. Fundam., 15(1), 59-64.
[161] Chakravarty, T., Phukan, U. K. and Weiland, R. H. (1985) Reaction of acid gases with
mixture of amines. Chem. Eng. Prog. 81, 32-36.
[162] Weiland, R. H., Chakravarty, T. and Mather, A. E. (1993) Solubility of carbon dioxide
and hydrogen sulfide in aqueous alkanolamines. Ind. Eng. Chem. Res., 32(7), 14191430.
[163] Pahlavanzadeh, H., Nourani, S. and Saber, M. (2011) Experimental analysis and
modelling of CO2 solubility in AMP (2-amino-2-methyl-1-propanol) at low CO2 partial
pressure using models of Deshmukh-Mather and the artificial neural network. J. Chem.
Thermodyn, 43 (12), 1775-1783.
[164] Chen, C. C. (1982) A local composition model for the excess gibbs energy of aqueous
electrolyte systems. I: single completely dissociated electrolyte single solvent systems.
AIChE J., 28(4), 588-596.
[165] Renon, H. and Prausnitz, J. M. (1968) Local composition on thermodynamic excess
functions for liquid mixtures. AIChE J., 14(1), 135-144.
[166] Chen, C. C. and Evans, L. N. (1986) A local composition model for the excess gibbs
energy of aqueous electrolyte systems. AIChE J., 32, 444454.
[167] Soave, G. (1971) Equilibrium constants from a modified Redlich-Kwong equation of
state. Chem. Eng. Sci., 27, 1197-1203.
[168] Hilliard, M. (2008). A Predictive Thermodynamic Model for an Aqueous Blend of
Potassium Carbonate, Piperazine, and Monoethanolamine for Carbon dioxide. Ph.D.
Thesis. The University of Texas at Austin.
[169] Posey, M. L. and Rocehlle, G. T. (1997) A thermodynamic model of
methyldiethanolamine-CO2-H2S-water. Ind. Eng. Chem. Res., 36, 3944-3953.

Bibliography

207

[170] Schmidt, K. A. G., Maham, Y. and Mather, A. E. (2007) Use of the NRTL equation for
simultaneous correlation of vapour-liquid equilibria and excess enthalpy applications to
aqueous alkanolamine systems. Journal of Thermal Analysis and Calorimetry, 89 (1),
61-72.
[171] Bollas, G., Chen, C. C. and Barton, P. (2008) Refined electrolyte-NRTL model:
activity coefficient expressions for application to multi-electrolyte systems. AIChE. J.,
54, 16081624.
[172] Hessen, E. T., Haug-Warberg, T. and Svendsen, H. F. (2010) The refined e-NRTL
model applied to CO2-H2O-alkanolamine systems. Chem. Eng. Sci., 65(11), 3638-3648.
[173] Zong, L. and Chen, C. C. (2011) Thermodynamic modelling of CO2 and H2S
solubilities in aqueous DIPA solution, aqueous sulfolane-DIPA solution, and aqueous
sulfolane-MDEA solution with electrolyte NRTL model. Fluid Phase Equilibria,
306(2), 190-203.
[174] Gross, J. and Sadowski, G. (2001) Perturbed-chain SAFT: An equation of state based
on a perturbation theory for chain molecules. Ind. Eng. Chem. Res., 40(4), 1244-1260.
[175] Zhang, Y., Que, H. and Chen, C. C. (2011) Thermodynamic modelling for CO2
absorption in aqueous MEA solution with electrolyte NRTL model. Fluid Phase
Equilibria, 311, 67-75.
[176] Galindo, A., Davies, L. A., Gil-Villegas, S. and Jackson, G. (1998) The
thermodynamics of mixtures and the corresponding mixing rules in the SAFT-VR
approach for potentials of variable range. Molecular Physics, 93(2), 241-252.
[177] Wertheim, M. S. (1984) Fluids with highly directional attractive forces. I. Statistical
thermodynamics. Journal of Statistical Physics, 35(1-2), 19-34.
[178] Wertheim, M. S. (1984) Fluids with highly directional attractive forces. II.
Thermodynamic perturbation theory and integral equations, Journal of Statistical
Physics, 35(1-2), 35-47.
[179] Wertheim, M. S. (1984) Fluids with highly directional attractive forces. III. Multiple
attraction sites, Journal of Statistical Physics, 42(3-4), 459-476.
[180] Wertheim, M. S. (1984) Fluids with highly directional attractive forces. IV.
Equilibrium polymerization, Journal of Statistical Physics, 42(3-4), 477-492.

Bibliography

208

[181] Chapman, W. G. and Gubbins, K. E. (2001) Theory and simulation of associating


liquid mixtures. Fluid Phase Equilibria, 29, 337-346.
[182] Galindo, A., Gil-Villegas, A., Whitehead, P. J. and Jackson G. (1998) Prediction of
phase equilibria for refrigerant mixtures of Difluoromethane (HFC-32), 1,1,1,2Tetrafluoroethane (HFC-134a), and Pentafluoroethane (HFC-125a) using SAFT-VR. J.
Phys. Chem. B, 102(39), 7632-7639.
[183] Keskes, E., Adjiman, C., Galindo, A. and Jackson, G. (2006) Integrating advanced
thermodynamics and process and solvent design for gas separation. Computer Aided
Chemical Engineering, 21, 743-748.
[184] Kaewsichan, L., Al-Bofersen, O., Yesavage, V. F. and Selim, M. S. (2001) Prediction
of the solubility of acid gases in monoethanolamine (MEA) and methyldiethanolamine
(MDEA) solutions using the electrolyte-UNIQUAC model. Fluid Phase Equilibria,
183-184, 159-171.
[185] Vrachnos, A., Kontogeorgis, G. and Voutsas, E. (2006) Thermodynamic modeling of
acidic gas solubility in aqueous solution of MEA, MDEA and MEA-MDEA blends. Ind.
Eng. Chem. Res., 45 (14), 5148-5154.
[186] Kuranov, G., Rumpf, B., Maurer, G. and Smirnova, N. (1997) VLE modeling for
aqueous systems containing methyldiethanolamine, carbon dioxide and hydrogen
sulfide. Fluid Phase Equilibria, 136, 147-162.
[187] Li, C. and Frst, W. (2000) Representation of CO2 and H2S solubility in aqueous
MDEA solution using an electrolyte equation of state. Chem. Eng. Sci., 55 (15), 29752988.
[188] Hampe, E. M. and Rudkevich, D. M. (2003) Exploring reversible reactions between
CO2 and amines. Tetrahedron, 59, 9619-9625.
[189] National Institute of Standards and Technology (NIST) Thermophysical properties of
fluid systems. [online] USA. Available from: http://webbook.nist.gov/chemistry/fluid/
[Accessed 23th March 2012].
[190] Guiochon, G. and Guillemin, C.L. (1988) Quantitative gas chromatography: for
laboratory analyses and on-line process control. Elsevier.
[191] Kohl, A. L. and Nielsen, R. B. (1997) Gas purification. Gulf Professional Publishing.

Bibliography

209

[192] Rochelle, G. T., Bishnoi, S., Chi, S., Dang H. Y. and Santos, J. (2001) Research needs
for CO2 capture from flue gas by aqueous absorption/stripping. Final report for DOE
contract DE-AF26-99FT01029.
[193] Mahajani, V. V. and Joshi, J. B. (1988) Kinetics of reactions between carbon dioxide
and alkanolamines. Gas Separation and Purification, 2(2), 50-64.
[194] Versteeg, G. F. and van Swaaij, W. P. M. (1987) On the kinetics between CO2 and
alkanolamines both in aqueous and non-aqueous solutions-II. Tertiary amines. Chem.
Eng. Sci., 43(3), 587-591.
[195] Aboudheir, A., Tontiwachwuthikul, P. and Chakma, A. (2003) Kinetics of the reactive
absorption of carbon dioxide in high CO2-loaded, concentrated aqueous
monoethanolamine solutions. Chem. Eng. Sci., 58, 5195-5210.
[196] Danckwerts, P. V. (1979) The reaction of CO2 with ethanolamines. Chem. Eng. Sci.,
34(4), 443-446.
[197] Caplow, M. (1968) Kinetics of carbamate formation and breakdown. J. Am. Chem.
Soc., 90(24), 6795-6803.
[198] Crooks, J. E. and Donnellan, J. P. (1989) Kinetics and mechanism of the reaction
between carbon dioxide and amines in aqueous solution. Journal of Chemical Society of
Perkin Transaction II, 331-333.
[199] Lyudkovskaya, M. A. and Leibush, A. G. (1949) Solubility of carbon dioxide in
solutions of ethanolamines under pressure. Zhur. Priklad. Khim., 22, 558-567.
[200] Atadan, E. M. (1954) Absorption of carbon dioxide by aqueous monoethanolamine
solutions. PhD Diss., University of Tennessee, Knoxville, TN.
[201] Goldman, A. M. and Leibush, A. G. (1959) Solubility of carbon dioxide in aqueous
solutions of monoethanolamine in the temperature range 75 C-140 C. Trudy Gosudarst.
Nauch. Issledov Proekt. Inst. Azot. Prom., 10, 54-82.
[202] Murzin, V. I. and Leites, I. L. (1971) Partial pressure of carbon dioxide over its dilute
solutions in aqueous aminoethanol, Zhur. Fiz. Khim., 45, 417-420.
[203] Austgen, D. M., Rochelle, G. T. and Chen, C. C. (1991) Model of vapor-liquid
equilibria for aqueous acid gas-alkanolamine systems. 2. Representation of hydrogen

Bibliography

210

sulfide and carbon dioxide solubility in aqueous MDEA and carbon dioxide solubility in
aqueous mixtures of MDEA with MEA or DEA. Ind. Eng. Chem. Res., 30(3), 543-555.
[204] Murrieta-Guevara, F., Rebolledo-Libreros, E. and Trejo, A. (1993) Gas solubility of
carbon dioxide and hydrogen sulfide in mixtures of sulfolane with monoethanolamine.
Fluid Phase Equilibria, 86, 225-231.
[205] Portugal, A. F., Sousa, J. M., Magalhaes, F. D. and Mendes, A. (2009) Solubility of
carbon dioxide in aqueous solutions of amino acid salts. Chem. Eng. Sci., 9 (1), 19932002.
[206] Aronu, U. E., Gondal, S., Hessen, E. T., Haug-Warberg, T., Hartono, A., Hoff, K. A.
and Svendsen, H. F. (2011) Solubility of CO2 in 15, 30, 45 and 60 mass% MEA from 40
to 120 C and model representation using the extended UNIQUAC framework. Chem.
Eng. Sci., 66(24), 6393-6406.
[207] Xu, S., Wang, Y. W., Otto, F. D. and Mather, A. E. (1996) Kinetics of the reaction of
carbon dioxide with 2-amino-2-methyl-1-propanol solutions. Chem. Eng. Sci., 51(6),
841-850.
[208] Littel, R. J., Versteeg, G. F. and van Swaaij, P. M. (1992) Kinetics of CO2 with primary
and secondary amines in aqueous solutions-I. Zwitterion deprotonation kinetics for DEA
and DIPA in aqueous blends of alkanolamines. Chem. Eng. Sci., 47(8), 2027-2035.
[209] Aboudher, A., Tontiwachwuthikul, P., Chakma, A. and Idem, R. (2003) Kinetics of the
reactive absorption of carbon dioxide in high CO2-loaded, concentrated aqueous
monoethanolamine solutions. Chem. Eng. Sci., 58(23-24), 5195-5210.
[210] de Silva, E. F. and Svendsen, H. F. (2004) Ab initio study of the reaction of carbamate
formation from CO2 and alkanolamines. Ind. Eng. Chem. Res., 43(13), 3413-3418.
[211] Sharma, M. M. (1961) Kinetics reactions of carbonyl sulphide and carbon dioxide with
amines and catalysis by Brnsted bases of the hydrolysis of COS. Trans. Faraday Soc.,
61, 681-688.
[212] Sharma, M. M. (1964) Kinetics of gas absorption. PhD Diss., University of Cambridge,
OA.
[213] Chakraborty, A. K., Astarita, G. and Bischoff, K. B. (1986) CO2 Absorption in
Aqueous Solutions of Hindered Amines. Chem. Eng. Sci., 41(4), 997-1003.

Bibliography

211

[214] Hook, R. J. (1997) An investigation of some sterically hindered amines as potential


carbon dioxide scrubbing compounds. Ind. Eng. Chem. Res., 36, 1779-1790.
[215] Chakraborty, A. K., Astarita, G., Bischoff, K. B. and Damewood, J. R. (1988).
Molecular orbital approach to substituent effects in amine-CO2 interactions. Journal of
American Chemical Society, 110(21), 6947-6954.
[216] Mimura, T., Suda, T., Iwaki, I., Honda, A. and Kumazawa, H. (1998) Kinetics of
reaction between carbon dioxide and stericallt hindered amines for carbon dioxide
recovery from power plant flue gases. Chem. Eng. Comm., 170(1), 245-260.
[217] Lensen, R. (2004) The promoter effect of piperazine on the removal of carbon dioxide.
[Online] Available from: http://www.bsdfreaks.nl/files/hoofd6.pdf [Accessed 25th
March 2012].
[218] Dash, S. K., Samanta, A., Samanta, A. N. and Bandyopadhyay, S. S. (2011) Absorption
of carbon dioxide in piperazine activated concentrated aqueous 2-amino-2-methyl-1propanol solvent. Chem. Eng. Sci., 66, 3223-3233.
[219] Cullinane, J. T. and Rochelle, G. T. (2006) Kinetics of carbon dioxide absorption into
aqueous potassium carbonate and piperazine. Ind. Eng. Chem. Res., 45(8), 2531-2545.
[220] Bishnoi, S. and Rochelle, G. T. (2004) Absorpion of carbon dioxide in aqueous
piperazine/methyldiethanolamine. AIChE J., 48(12), 2788-2799.
[221] Bishnoi, S. and Rochelle, G. T. (2000) Absorption of carbon dioxide into aqueous
piperazine: reaction kinetics, mass transfer and solubility. Chem. Eng. Sci., 55, 55315543.
[222] Samanta, A. and Bandyopadhyay, S. S. (2007) Kinetics and modelling of carbon
dioxide adsorption into aqueous solutions of piperazine. Chem. Eng. Sci., 62, 7312-7319.
[223] Samanta, A. and Bandyopadhyay, S. S. (2009) Absorption of carbon dioxide into
aqueous solutions of piperazine activated 2-amino-2-methyl-1-propanol. Chem. Eng.
Sci., 64, 1185-1194.
[224] Puxty, G. and Rowland, R. (2011) Modelling CO2 mass transfer in amine mixtures: PZAMP and PZ-MDEA. Environ. Sci. Technol., 45(6), 2398-2405.
[225] Roberts, B. E. and Mather, A. E. (1988) Solubility of CO2 and H2S in a Hindered
Amine Solution. Chem. Eng. Comm., 64, 105-111.

Bibliography

212

[226] Blauwhoff, P. M. M., Versteeg, G. F. and van Swaaij, W. P. M. (1984) A study on the
reaction between CO2 and alkanolamines in aqueous solutions. Chem. Eng. Sci., 39(2),
207-225.
[227] Barth, D., Tondre, C. and Delpuech, J. J. (1984) Kinetics and mechanisms of the
reactions of carbon dioxide with alkanolmaines: a discussion concerning the cases of
MDEA and DEA. Chem. Eng. Sci., 39(12), 1753-1757.
[228] Barth, D., Tondre, C., Lappai, G. and Delpuech, J. J. (1981) Kinetic study of carbon
dioxide reaction with tertiary amines in aqueous solutions. J. Phys. Chem., 85(24), 36603667.
[229] Sada, E., Kumazawa, H., Han, Z. W. and Matsuyama, H. (2004) Chemical kinetics of
the reaction of carbon dioxide with ethanolamines in nonaqueous solvents. AIChE J.,
31(8), 1297-1303.
[230] Park, S. W., Lee, J. W., Choi, B. S. and Lee, J. W. (2006) Absorption of carbon dioxide
into non-aqueous solutions of N-Methyldiethanolamine. Chemistry and Materials
Science, 23(5), 806-811.
[231] Kierzkowska-Pawlak, H. and Zarzycki, R. (2002) Solubility of carbon dioxide and
nitrous oxide in water + methyldiethanolamine and ethanol + methyldiethanolamine
solutions. J. Chem. Eng. Data, 47(6), 1506-1509.
[232] Donaldson, T. L. and Nguyen, Y. N. (1980) Carbon dioxide reaction kinetics and
transport in aqueous amine membranes. Ind. Eng. Chem. Fundamen., 19(3), 260-266.
[233] Henni, A., Li J. and Tontiwachwuthikul, P. (2008) Reaction kinetics of CO2 in aqueous
1-amino-2-propanol, 3-amino-1-propanol, and dimethylmonoethanolamine solutions in
the temperature range of 298-313 K using the stopped-flow technique. Ind. Eng. Chem.
Res., 47(7) 2213-2220.
[234] Lawal, A. O. and Idem, R. O. (2006) Kinetics of the oxidative degradation of CO2
loaded and concentrated aqueous MEA-MDEA blends during CO2 absorption from flue
gas streams. Ind. Eng. Chem. Res., 45(8), 2601-2607.
[235] Ali, B. S. and Aroua, M. K. (2004) Effect of piperazine on CO2 loading in aqueous
solutions of MDEA at low pressure. International Journal of Thermodynamics, 25(6),
1863-1870.

Bibliography

213

[236] Zhang, X., Zhang, C. F. and Liu, Y. (2002) Kinetics of absorption of CO2 into aqueous
solution of MDEA blended with DEA. Ind. Eng. Chem. Res., 41(5), 1135-1141.
[237] Pacheco, M. A., Kaganoi, S. and Rochelle, G. T. (2000) CO2 absorption into aqueous
mixtures of diglycolamine and methyldiethanolamine. Chem. Eng. Sci., 55(21), 51255140.
[238] Glasscock, D. A., Critchfield, J. E. and Rochelle, G. T. (1991) CO2
absorption/desorption in mixtures of methyldiethanolamine with monoethanolamine or
diethanolamine. Chem. Eng. Sci., 46(11), 2829-2845.
[239] Rinker, E. B., Ashour, S. S. and Sandall, O. C. (2000) Absoprtion of carbon dioxide
into aqueous blends of diethanolamine and methyldiethanolamine. Ind. Eng. Chem. Res.,
39(11), 4346-4356.
[240] Dubois, L. and Thomas, D. (2009) CO2 absorption into aqueous solutions of
monoethanolamine, methyldiethanolamine, piperazine and their blends. Chem. Eng.
Technol., 32(5), 710-718.
[241] Liao, C. H. and Li, M. H. (2002) Kinetics of absorption of carbon dioxide into aqueous
solutions of monoethanolamine + N-methyldiethanolamine. Chem. Eng. Sci., 57, 45694582.
[242] Ramachandran, N., Aboudher, A., Idem, R. and Tontiwachwuthikul, P. (2006) Kinetics
of the aborption of CO2 into mixed aqueous loaded solutions of monoethanolamine and
methyldiethanolamine. Ind. Eng. Chem. Res., 45, 2608-2616.
[243] Appl, M., Wagner, U., Henrici, H.J., Kuessnet, K., Volkamer, F. and Ernst-Neust, N.
(1982) Removal of CO2 and/or H2S and/or COS from gases containing these
constituents. US Patent Nr 4336233.
[244] Brder, P. and Svendsen, H. S. (2011) Solvent composition for postcombustion CO2
capture. 1st Post Combustion Capture Conference, 17th-19th May, [Online] Abu Dhabi.
Available from:
http://www.ieaghg.org/docs/General_Docs/PCCC1/Abstracts_Final/pccc1Abstract00075
.pdf [Accessed 26th March 2012].
[245] Little, R. J., van Swaaij, W. P. M. and Versteeg, G. F. (1990) Kinetics of carbon
dioxide with tertiary amines in aqueous solution. AIChE J., 36(11), 1633-1640.

Bibliography

214

[246] Saul, A. and Wagner, W. (1987) International Equation for the Saturation Properties of
Ordinary Water Substance. J. Phys. Chem. Ref. Data, 16, 893-901.
[247] Brelvi, S. W. and OConnell, J. P. (1972) Corresponding states correlations for liquid
compressibility and partial molal volumes of gases at infinite dilution in liquids. AIChE
J., 18, 1239-1243.
[248] Bieling, V., Kurz, F., Rumpf, B., and Maurer, G. (1995) Simultaneous Solubility of
Ammonia and Carbon Dioxide in Aqueous Solutions of Sodium Sulfate in the
Temperature Range 313-393 K and Pressures up to 3 MPa. Ind. Eng. Chem. Res., 34,
1449-1460.
[249] Plyasunov, A. V. and Shock, E. L. (2003) Second Cross Virial Coefficients for
Interactions Involving Water. Critical Data Compilation. J. Chem. Eng. Data, 48, 808821.
[250] Guggenheim, E. A. and Stokes, R. H. (1958). Activity coefficients of 2:1 and 1:2
electrolytes in aqueous solution from isopiestic data. Trans. Faraday Soc., 54, 1646-1649.
[251] Lewis, G. N., Randall, M., Pitzer, K. S., and Brewer, L. (1961) Thermodynamics
second ed. McGraw Hill.
[252] Process Systems Enterprise Ltd (2004) gPROMS advanced user guide, Process
Systems Enterprise Ltd, London, UK.
[253] Process Systems Enterprise Ltd (2010) gPROMS v 3.3.1. Available from:
http://www.psenterprise.com/ [Accessed 27th March 2012].
[254] Poplsteinova, J.J., Krane, J., and Svendsen, H.F., 2005. Liquid-phase composition
determination in CO2H2Oalkanolamine systems: an NMR study. Ind. Eng. Chem. Res.,
[255] Bttinger, W., Maiwald, M. and Hasse, H. (2008) Online NMR spectroscopic study of
species distribution in MEA-H2O-CO2 and DEA-H2O-CO2. Fluid Phase Equilibria,
263(2), 131-143.
[256] Hilliard, M. (2008) A Predictive Thermodynamic Model for an Aqueous Blend of
Potassium Carbonate, Piperazine, and Monoethanolamine for Carbon dioxide. Ph.D.
Diss. The University of Texas at Austin.

Bibliography

215

[257] Kamps, A. P., Xia, J. and Maurer, G. (2003) Solubility of CO2 in (H2O+piperazine) and
in (H2O+MDEA+piperazine). AIChE J., 49(10), 2662-2670.
[258] Ermatchkov, V., Kamps A. P., Maurer G. (2002) Chemical equilibrium constants for
the formation of carbamates in (carbon dioxide + piperazine + water) from 1H-NMRspectroscopy. J Chem Thermodyn., 35, 1277-1289.
[259] Cullinane, J. T. and Rochelle, G. T. (2005) Thermodynamics of aqueous potassium
carbonate, piperazine, and carbon dioxide. Fluid Phase Equilibria, 227(2), 197-213.
[260] Hamborg, E.S. and Versteeg, G.F. (2009) Dissociation Constants and Thermodynamic
Properties of Amines and Alkanolamines from (293 to 353) K. J. Chem. Eng. Data, 54,
1318-1328.
[261] Rumpf, B. and Maurer, G. (1993) An Experimental and Theoretical Investigation on
the Solubility of Carbon Dioxide in Aqueous Solutions of Strong Electrolytes. Ber.
Bunsen-Ges. Phys. Chem., 97, 85-97.
[262] Bunsen, R. (2006) Ueber das Gesetz der Gasabsorption. European Journal of Organic
Chemistry, 93(1), 1-50.
[263] Haar, L., Gallagher, J. S. and Kell, G. S. (1984) NBS/NRC Steam Tables.
Thermodynamic and Transport Properties and Computer Programs for Vapor and Liquid
States of Water in SI Units, Hemisphere.
[264] Redlich, O. and Kwong, J. N. S. (1949) On the thermodynamics of solutions. V An
equation of state. Fugacities of gaseous solutions. Chem. Rev., 44, 233-244.
[265] Novk, J., Fried, V. and Pick, J. (1961) Coll Czech. Chem. Commun. 26, 2266-2270.
[266] Crovetto, R. and Wood, R. H. (1992) Solubility of CO2 in water and density of aqueous
CO2 near the solvent critical temperature. Fluid Phase Equilibria, 74(2), 271-288.
[267] Van Dusen, M. S. (1925) Platinum-resistance thermometry at low temperatures. JACS,
47, 326-332.
[268] Span, R. and Wagner, W. (1996) A new equation of state for carbon dioxide covering
the fluid region from the triple pont temperature to 110 K at pressures up to 800 MPa.
JPCRD, 25(3), 1053-1121.

Bibliography

216

[269] Wagner, W. and Pruss, A. (2002) The IAPWS formulation (1995) for the
thermodynamic properties of ordinary water substance for general and scientific use. J.
Phys. Chem. Ref. Data, 31, 387-535.
[270] Kamps, A. P., Balaban, A., Jdecke, M., Kuranov, G., Smirnova, N. A. and Maurer, G.
(2001) Solubility of single gases carbon dioxide and hydrogen sulphide in aqueous
solutions of N-methyldiethanolamine at temperatures from 313 to 393 K and pressures
up to 7.6 MPa: New experimental data and model extension. Ind. Eng. Chem. Res., 40(2),
696-706.
[271] Plyasunov, A. V. and Shock, E. L. (2003) Second cross virial coefficients for
interactions involving water. Critical data compilation. J. Chem. Eng. Data, 48(4), 808821.
[272] Tong, D., Trusler, J. P., Maitland, G. C., Gibbins, J. and Fennell, P. S. (2012) Solubility
of carbon dioxide in aqueous solution of monoethanolamine or 2-amino-2-methyl-1propanol: Experimental measurements and modelling. International Journal of
Greenhouse Gas Control, 6, 37-47.
[273] Tamms, K. and Gilliland, D (2011) Mountaineer project: CCS business case report.
Project no.: PRO 004. [Online] Available from:
http://cdn.globalccsinstitute.com/sites/default/files/publications/32971/global-ccsinstitute-business-case-final.pdf [Accessed 29th March 2012]
[274] Mamun, A., Svendsen, H. F., Hoff, K. A. and Juliussen, O. (2007) Selection of new
absorbents for carbon dioxide capture. Energy Conversion and Management, 48, 251258.
[275] Singh, P., Niederer, J. P. M. and Versteeg, G. F. (2007) Structure and activity
relationships for amine based CO2 absorbents-I. International Journal of Greenhouse Gas
control, 1, 5-10.
[276] Singh, P., Niederer, J. P. M. and Versteeg, G. F. (2009) Structure and activity
relationships for amine-based CO2 absorbents-II. Chem. Eng. Res. Des., 87(2), 135-144.
[277] Chen, X., Closman, F. and Rochelle, G. T. (2011) Accurate screening of amines by the
wetted wall column. Energy Procedia, 4, 101-108.

Bibliography

217

[278] Kim, S. T., Kang, J. W., Lee, J. S. and Min, B. M. (2011) Analysis of the heat of
reaction and regeneration in alkanolamine-CO2 system. Korean J. Chem. Eng., 28(12),
2275-2281.
[279] Dave, N., Do, T., Puxty, G., Rowland, R., Feron, P. H. M. and Attalla, M. I. (2009)
CO2 capture by aqueous amines and aqueous ammonia-A comparison. Energy Procedia,
1(1), 949-954.
[280] Lepaumier, H., Picq, D. and Carrette, P. L. (2009) New amines for CO2 capture. I.
Mechanisms of amine degradation in the presence of CO2. Applied Chemistry, 48, 90619067.
[281] Lepaumier, H., Picq, D. and Carrette, P. L. (2009) New amines for CO2 capture. II.
Oxidative degradation mechanisms. Ind. Eng. Chem. Res., 48, 9068-9075.
[282] Islam, M. S., Yusoff, R., Ali, B. S., Islam, M. N. and Chakrabarti, M. H. (2011)
Degradation studies of amines and alkanolamines during sour gas treatment process.
International Journal of the Physical Science, 6(25), 5877-5890.
[281] Sada, E., Kumazawa, H. and Butt, M. A. (1976) Gas absorption with consecutive
chemical reaction: Absorption of carbon dioxide into aqueous amine solutions. Can. J.
Chem. Eng., 54(5), 421-424.
[282] Hikita, H., Asai, S., Ishikawa, H. and Honda, M. (1977) The kinetics of reactions of
carbon dioxide with monoisopropanolamine, diglycolamine, and ethylenediamine by a
rapid mixing method. Chem. Eng. J., 14, 27-30.
[283] Davis, R. A. and Sandall, O. C. (1993) Kinetics of the reaction of carbon dioxide with
secondary amines in polyethylene glycol. Chem. Eng. Sci., 48(18), 3187-3193.
[284] Rinker, E. B., Ashour, S. S. and Sandall, O. C. (1996) Kinetics and modelling of carbon
dioxide absorption into aqueous solutions of diethanolamine. Ind. Eng. Chem. Res., 35,
1107-1114.
[285] Haimour, N., Bidarian, A. and Sandall, O. C. (1987) Kinetics of the reaction between
carbon dioxide and methyldiethanolamine. Chem. Eng. Sci., 42(6), 1393-1398.
[286] Rinker, E. B., Ashour, S. S. and Sandall, O. C. (1995) Kinetics and modelling of carbon
dioxide absorption into aqueous solutions of N-methyldiethanolamine. Chem. Eng. Sci.,
50(5), 755-768.

Bibliography

218

[287] Crooks, J. E. and Donnellan, J. P. (1990) Kinetics of the reaction between carbon
dioxide and tertiary amines. J. Org. Chem., 55, 1372-1374.
[288] Hikita, H., Asai, S., Ishikawa, H. and Honda, M. (1977) The kinetics of reactions of
carbon dioxide with monoethanolamine, diethanolamine, and triethanolamine by a rapid
mixing method. Chem. Eng. J., 13, 7-12.
[289] Bosch, H., Versteeg, G. F. and van Swaaij, W. P. M. (1984) A study on the reaction
between CO2 and alkanolamines in aqueous solutions. Chem. Eng. Sci., 39(2), 207[290] Seo, D. J. and Hong, W. H. (2000) Effect of piperazine on the kinetics of carbon
dioxide with aqueous solutions of 2-amino-2-methyl-1-propanol. Ind. Eng. Chem. Res.,
39, 2062-2067.
[291] Sun, W. C., Yong, C. B. and Li, M. H. (2005) Kinetics of the absorption of carbon
dioxide into mixed aqueous solutions of 2-amino-2-methyl-1-propanol and piperazine.
Chem. Eng. Sci., 60, 503-516.
[292] Choi, W. J., Cho, K. C., Lee, S. S., Shim, J. G., Hwang, H. R., Park, S. W. and Oh, K.
J. (2007) Removal of carbon dioxide by absorption into blended amines: kinetics of
absorption into aqueous AMP/HMDA. AMP/MDEA, and AMP/piperazine solutions.
Green Chemistry, 9, 594-598.
[293] Derks, P. W. J., Kleingeld, C., van Aken, C., Hogendoorn, J. A. and Versteeg, G. F.
(2006) Kinetics of absorption of carbon dioxide in aqueous piperazine solutions. Chem.
Eng. Sci., 61, 6837-6854.
[294] Hagewiesche, D. P., Ashour, S. S., Al-Ghawas, H. A. and Sandall, O. C. (1995)
Absorption of carbon dioxide into aqueous blends of monoethanolamine and Nmethyldiethanolamine. Chem. Eng. Sci., 50, 1071-1079.
[295] Bishnoi, S. (2000) CO2 absorption and solution equilibrium in piperazine activated
methyldiethanolamine. PhD Diss., The University of Texas at Austin.
[296] Yang, Z. Y., Soriano, A. N., Caparanga, A. R. and Li, M. H. (2010) Equilibrium
solubility of carbon dioxide in (2-amino-2-methyl-1-propanol + piperazine + water). J.
Chem. Therm., 42, 659-665.
[297] Valtz, A., Hegarty, M. and Richon, D. (2003) Experimental determination of the
solubility of aromatic compounds in aqueous solutions of various amines. Fluid Phase
Equilibria, 210, 257-276.

Bibliography

219

[298] Valtz, A., Coquelet, C. and Richon, D. (2005) Volumetric properties of the
monoethanolamine-methanol mixtures at atmospheric pressure from 283.15 to 353.15 K.
Thermochimica Acta., 428, 185-191.
[299] ROLSI valve, Transvalor. Available from: http://www.rolsi.com/English.htm
[Accessed 29th March 2012].
[300] Singh, P., Brilman, D. W. F. and Groeneveld, M. J. (2011) Evaluation of CO2 solubility
in potential aqueous amine-based solvents at low CO2 partial pressure. International
Journal of Greenhouse Gas Control, 5(1), 61-68.
[301] Hsu, C. H. and Li, M. H. (2009) Densities of Aqueous Blended Amines. J. Chem. Eng.
Data, 42(3), 502-507.

Appendices

Appendices

220

Appendices

221

I Summary of typical amines

Class

Name (Abbr.)
Monoethanolamine (MEA)

Structure

Primary Amine
Diglycolamine (DGA)

Diethanolamine (DEA)
Secondary Amine
N-ethylethanolamine (EEA)

Diisopropanolamine (DIPA)

Triethanolamine (TEA)
Tertiary Amine
Methyldiethanolamine (MDEA)

2-Dimethylaminoethanol (DMMEA)

OH

N,N-diethylethanolamine (DEEA)
N

Isobutanolamine (AMP)
Hindered Amine
2-piperidineethanol (PE)

Piperazine (PZ)
Diamine
Aminoethylethanolamine (AEEA)

Appendices

222

II Sample gPROMS code (MEA, DeshmukhMather model)


# Chemical reactions in aqueous MEA #
# ---- Ionization of water ---- #
#

2H2O = OH- + H3O+

(1)

# ---- Dissociation of dissolved CO2 through carbonic acid ---- #


#

CO2 + 2H2O = HCO3- + H3O+

(2)

# ---- Dissociation of bicarbonate ---- #


#

HCO3- + H2O = CO3-- + H3O+

(3)

# ---- Dissociation of protonated MEA ---- #


#

RR'NH2+ + H2O = RR'NH + H3O+ (4)

# ---- Reversion to bicarbonate ---- #


#

RR'NCOO- + H2O = RR'NH + HCO3- (5)

#......................#
# Components list #
#......................#
#

1: CO2

2: H2O

3: MEA(RR'NH) #

4: OH-

5: H3O+

6: HCO3-

7: CO3--

8: RR'NH2+

9: RR'NCOO-

#
#
#

PARAMETER
T1, T2 AS REAL
NoReaction AS INTEGER DEFAULT 5
MW_MEA

AS INTEGER DEFAULT 61

Appendices

223

NoMolecules AS INTEGER DEFAULT 3


NoSPECIES AS INTEGER DEFAULT 9
No1 AS INTEGER DEFAULT 25
No2 AS INTEGER DEFAULT 63
No AS INTEGER DEFAULT 12
B AS REAL DEFAULT 1.2
iteration AS INTEGER DEFAULT 50
iteration1 AS INTEGER DEFAULT 49
a_1,a_2,b_1,b_2,c_1,c_2,d_1,d_2 AS REAL
R

AS REAL DEFAULT 8.314

Tc

AS ARRAY(NoMolecules) OF REAL

Pc

AS ARRAY(NoMolecules) OF REAL

Omega

AS ARRAY(NoMolecules) OF REAL

VARIABLE
K1, K2 AS ARRAY (NoReaction) OF No_type
H1, H2 AS No_type
m1 AS ARRAY (NoSpecies,No1,iteration) OF Molarity
m2 AS ARRAY (NoSpecies,No2,iteration) OF Molarity
Gamma1 AS ARRAY (NoSpecies,No1,iteration1) OF Coefficient
Gamma2 AS ARRAY (NoSpecies,No2,iteration1) OF Coefficient
Beta1, Beta2 AS ARRAY (No) OF No_type

A1, A2 AS No_type
I1 AS ARRAY (No1,iteration1) Of No_type
I2 AS ARRAY (No2,iteration1) Of No_type
P1 AS ARRAY (No1) Of No_type
P2 AS ARRAY (No2) Of No_type
Loading1 AS ARRAY (No1,iteration1) Of No_type
Loading2 AS ARRAY (No2,iteration1) Of No_type

Appendices

224

Loading_E1 AS ARRAY (No1) Of No_type


Loading_E2 AS ARRAY (No2) Of No_type
c, d AS ARRAY (No) OF No_type # parameters in binary interaction parameters need to be regressed
from experimental data
B11_1,B11_2,B22_1,B22_2,B12_1,B12_2

AS No_type

W_MEA1 AS No_type
Bm1

AS ARRAY(No1) OF No_type

Bm2

AS ARRAY(No2) OF No_type

Pyvot_1

AS ARRAY(No1) OF No_type

Pyvot_2

AS ARRAY(No2) OF No_type

Pw_s_1,Pw_s_2,v_CO2_1,v_CO2_2 AS No_type
Vapour_Fraction1

AS ARRAY(NoMolecules-1,No1) OF fraction

Vapour_Fraction2

AS ARRAY(NoMolecules-1,No2) OF fraction

Fugacity_Coefficient1 AS ARRAY(NoMolecules-1,No1) OF coefficient


Fugacity_Coefficient2 AS ARRAY(NoMolecules-1,No2) OF coefficient

SET
Tc(1) := 304.15;
Tc(2) := 647.30;
Tc(3) := 741.9;

Pc(1) := 7374600;
Pc(2) := 22048000;
Pc(3) := 4160000;

Omega(1) := 0.225;
Omega(2) := 0.344;
Omega(3) := 0.8373; # MEA

Appendices

225

a_1 := 65.703;
a_2 := -53.53;
b_1 := -184.854;
b_2 := -39.29;
c_1 := 304.16;
c_2 := 647.3;
d_1 := 1.4;
d_2 := 4.3;

T1 := 313.15;
T2 := 393.15;

EQUATION
Vapour_fraction1(2,) = 1-Vapour_fraction1(1,);
Vapour_fraction2(2,) = 1-Vapour_fraction2(1,);
# Pyvoting term #
v_CO2_1 = 0.000375*T1^2-0.177362*T1+52.1675;
v_CO2_2 = 0.000375*T2^2-0.177362*T2+52.1675;
Pw_s_1 = Pc(2)/1000*EXP(Tc(2)/T1*(-7.8582*(1-T1/Tc(2))+1.83991*(1-T1/Tc(2))^1.5-11.781*(1T1/Tc(2))^3+22.6705*(1-T1/Tc(2))^3.5-15.939*(1-T1/Tc(2))^4+1.77516*(1-T1/Tc(2))^7.5));
Pw_s_2 = Pc(2)/1000*EXP(Tc(2)/T2*(-7.8582*(1-T2/Tc(2))+1.83991*(1-T2/Tc(2))^1.5-11.781*(1T2/Tc(2))^3+22.6705*(1-T2/Tc(2))^3.5-15.939*(1-T2/Tc(2))^4+1.77516*(1-T2/Tc(2))^7.5));
#First Temperature
# Equilibrium Constant
K1(1) = EXP((-13445.9 / T1) - 22.4773 * LOG(T1) + 140.932);
K1(2) = EXP((-12092.1 / T1) - 36.7816 * LOG(T1) + 235.482);
K1(3) = EXP((-12431.7 / T1) - 35.4819 * LOG(T1) + 220.067);
K1(4) = EXP((-17.33 / T1) + 0.05764 * T1 - 38.846);
K1(5) = EXP(-1545.3/T1+2.151);

H1 = EXP(192.876-9624.4/T1+0.01441*T1-28.749*log(T1));

Appendices

226

##------------------------------------Vapour Phase Non-ideality Calculation------------------------------------##


# Virial Equation of State
#~~~ First Temperature ~~~#
B11_1 = a_1 + b_1*(c_1/T1)^d_1;
B22_1 = a_2 + b_2*(c_2/T1)^d_2;
B12_1 = -211.31-729.48*(298.15/T1-1)-1064.54*(298.15/T1-1)^2-656.13*(298.15/T1-1)^3;
# First condition #
For ll := 1 to No1
Do
Bm1(ll)
=
Vapour_fraction1(1,ll)*Vapour_fraction1(1,ll)*B11_1+2*Vapour_fraction1(1,ll)*Vapour_fraction1(2,l
l)*B12_1+Vapour_fraction1(2,ll)*Vapour_fraction1(2,ll)*B22_1;
Fugacity_Coefficient1(1,ll)
=
EXP(P1(ll)/Vapour_fraction1(1,ll)/1000/R/T1*(2*(Vapour_fraction1(1,ll)*B11_1+Vapour_fraction1(2,
ll)*B12_1)-Bm1(ll)));
Fugacity_Coefficient1(2,ll)
=
EXP(P1(ll)/Vapour_fraction1(1,ll)/1000/R/T1*(2*(Vapour_fraction1(1,ll)*B12_1+Vapour_fraction1(2,
ll)*B22_1)-Bm1(ll)));
Pyvot_1(ll) = EXP(v_CO2_1*(P1(ll)/Vapour_fraction1(1,ll)/1000-Pw_s_1)/R/T1);
End

##------------------------------------Liquid Phase Non-ideality Calculation------------------------------------##


# Debye-Huckel Constant related to Temperature
A1 = 1.313+1.335*10^(-3)*(T1-273.15)+1.164*10^(-5)*(T1-273.15)^2;
# Binary Interaction Parameters
Beta1(1) = c(1) + d(1) * T1;
Beta1(2) = c(2) + d(2) * T1;
Beta1(3) = c(3) + d(3) * T1;
Beta1(4) = c(4) + d(4) * T1;
Beta1(5) = c(5) + d(5) * T1;
Beta1(6) = c(6) + d(6) * T1;

Appendices

227

Beta1(7) = c(7) + d(7) * T1;


Beta1(8) = c(8) + d(8) * T1;
Beta1(9) = c(9) + d(9) * T1;
Beta1(10) = c(10) + d(10) * T1;
Beta1(11) = c(11) + d(11) * T1;
Beta1(12) = c(12) + d(12) * T1;
# Experimental Points
For j :=1 to No1
Do
# Molality Calculated by setting Activity Coefficient to zero
m1(1,j,1) = P1(j)*Fugacity_Coefficient1(1,j)/H1/1000/Pyvot_1(j);
m1(2,j,1) = 1;
m1(3,j,1) = 1000 * (W_MEA1/MW_MEA)/(1-W_MEA1) - m1(8,j,1) - m1(9,j,1);
m1(4,j,1)*m1(5,j,1)*10^14 = K1(1)*10^14 ;
m1(5,j,1)*m1(3,j,1)*10^10 = K1(4)*m1(8,j,1)* 10^10;
m1(6,j,1)*m1(5,j,1)*10^7 = K1(2)*m1(1,j,1)*10^7;
m1(7,j,1)*m1(5,j,1)*10^11 = K1(3)*m1(6,j,1)*10^11 ;
m1(8,j,1) = m1(6,j,1) + 2 * m1(7,j,1);
m1(9,j,1) * K1(5) * 10^3 = m1(3,j,1) *m1(6,j,1) * 10^3;
# Iteration to Convergy
For l := 2 to Iteration
Do
I1(j,l-1) = 0.5 * (m1(4,j,l-1) + m1(5,j,l-1) + m1(6,j,l-1) + 4 * m1(7,j,l-1) + m1(8,j,l-1)+m1(9,j,l-1));
Loading1(j,l-1) = (m1(1,j,l-1) + m1(6,j,l-1) + m1(7,j,l-1))/(1000 * (W_MEA1/MW_MEA)/(1-W_MEA1));
# Activity Coefficient Calculated from Molarity
Gamma1(1,j,l-1) = Exp(2 * (Beta1(1) * m1(8,j,l-1) + Beta1(4) * m1(3,j,l-1) + Beta1(7) * m1(6,j,l-1) +
Beta1(8) * m1(7,j,l-1)+Beta1(9)*m1(9,j,l-1)));
Gamma1(2,j,l-1) = 1;
Gamma1(3,j,l-1) = Exp(2 * (Beta1(4) * m1(1,j,l-1) + Beta1(5) * m1(6,j,l-1) + Beta1(6) * m1(7,j,l-1)));
Gamma1(4,j,l-1) = 1;

Appendices

228

Gamma1(5,j,l-1) = 1;
Gamma1(6,j,l-1) = Exp(-A1 * I1(j,l-1) ^ 0.5 / (1 + B * I1(j,l-1) ^ 0.5) + 2 * (Beta1(2) * m1(8,j,l-1) +
Beta1(5) * m1(3,j,l-1) + Beta1(7) * m1(1,j,l-1)+Beta1(10)*m1(9,j,l-1)));
Gamma1(7,j,l-1) = Exp(-A1 * 4 * I1(j,l-1) ^ 0.5 / (1 + B * I1(j,l-1) ^ 0.5) + 2 * (Beta1(3) * m1(8,j,l-1) +
Beta1(6) * m1(3,j,l-1) + Beta1(8) * m1(1,j,l-1)+Beta1(11)*m1(9,j,l-1)));
Gamma1(8,j,l-1) = Exp(-A1 * I1(j,l-1) ^ 0.5 / (1 + B * I1(j,l-1) ^ 0.5) + 2 * (Beta1(1) * m1(1,j,l-1) +
Beta1(2) * m1(6,j,l-1) + Beta1(3) * m1(7,j,l-1)+Beta1(12)*m1(9,j,l-1)));
Gamma1(9,j,l-1) = Exp(-A1 * I1(j,l-1) ^ 0.5 / (1 + B * I1(j,l-1) ^ 0.5) + 2 * (Beta1(9) * m1(1,j,l-1) +
Beta1(10)* m1(6,j,l-1) + Beta1(12) * m1(8,j,l-1)));
# New Molarity Calculated from Activity Coefficient
m1(1,j,l)*Gamma1(1,j,l-1) = P1(j)*Fugacity_Coefficient1(1,j)/H1/1000/Pyvot_1(j); #*Gamma1(1,j,l-1)
m1(2,j,l) = 1;
m1(3,j,l) = 1000 * (W_MEA1/MW_MEA)/(1-W_MEA1) - m1(8,j,l) - m1(9,j,l);
m1(4,j,l)*m1(5,j,l)*10^14 = K1(1)*10^14 ;
m1(5,j,l)*m1(3,j,l)*Gamma1(3,j,l-1)*10^10 = K1(4)* (m1(8,j,l)*Gamma1(8,j,l-1))*10^10;
m1(6,j,l)*m1(5,j,l)*Gamma1(6,j,l-1)*10^7 = K1(2) *10^7 *(m1(1,j,l)*Gamma1(1,j,l-1));
m1(7,j,l)*Gamma1(7,j,l-1)*m1(5,j,l)*10^11 = K1(3)*10^11*(m1(6,j,l)*Gamma1(6,j,l-1));
m1(8,j,l) = m1(6,j,l) + 2 * m1(7,j,l);
m1(9,j,l)*Gamma1(9,j,l-1) * K1(5) * 10^3 = m1(3,j,l)*Gamma1(3,j,l-1) *m1(6,j,l)*Gamma1(6,j,l-1) *
10^3;
End
End

# Second Condition
# Equilibrium Constant
K2(1) = EXP((-13445.9 / T2) - 22.4773 * LOG(T2) + 140.932);
K2(2) = EXP((-12092.1 / T2) - 36.7816 * LOG(T2) + 235.482);
K2(3) = EXP((-12431.7 / T2) - 35.4819 * LOG(T2) + 220.067);
K2(4) = EXP((-17.33 / T2) + 0.05764 * T2 - 38.846);
K2(5) = EXP(-1545.3/T2+2.151);

H2 = EXP(192.876-9624.4/T2+0.01441*T2-28.749*log(T2));

Appendices

229

##------------------------------------Vapour Phase Non-ideality Calculation------------------------------------##


# Virial Equation of State
#~~~ First Temperature ~~~#
B11_2 = a_1 + b_1*(c_1/T2)^d_1;
B22_2 = a_2 + b_2*(c_2/T2)^d_2;
B12_2 = -211.31-729.48*(298.15/T2-1)-1064.54*(298.15/T2-1)^2-656.13*(298.15/T2-1)^3;
# First condition #
For ll := 1 to No2
Do
Bm2(ll)
=
Vapour_fraction2(1,ll)*Vapour_fraction2(1,ll)*B11_2+2*Vapour_fraction2(1,ll)*Vapour_fraction2(2,l
l)*B12_2+Vapour_fraction2(2,ll)*Vapour_fraction2(2,ll)*B22_2;
Fugacity_Coefficient2(1,ll)
=
EXP(P2(ll)/Vapour_fraction2(1,ll)/1000/R/T2*(2*(Vapour_fraction2(1,ll)*B11_2+Vapour_fraction2(2,
ll)*B12_2)-Bm2(ll)));
Fugacity_Coefficient2(2,ll)
=
EXP(P2(ll)/Vapour_fraction2(1,ll)/1000/R/T2*(2*(Vapour_fraction2(1,ll)*B12_2+Vapour_fraction2(2,
ll)*B22_2)-Bm2(ll)));
Pyvot_2(ll) = EXP(v_CO2_2*(P2(ll)/Vapour_fraction2(1,ll)/1000-Pw_s_2)/R/T2);
End
##------------------------------------Liquid Phase Non-ideality Calculation------------------------------------##

# Debye-Huckel Constant related to Temperature


A2 = 1.313+1.335*10^(-3)*(T2-273.15)+1.164*10^(-5)*(T2-273.15)^2;
# Binary Interaction Parameters
Beta2(1) = c(1) + d(1) * T2;
Beta2(2) = c(2) + d(2) * T2;
Beta2(3) = c(3) + d(3) * T2;
Beta2(4) = c(4) + d(4) * T2;
Beta2(5) = c(5) + d(5) * T2;
Beta2(6) = c(6) + d(6) * T2;

Appendices

230

Beta2(7) = c(7) + d(7) * T2;


Beta2(8) = c(8) + d(8) * T2;
Beta2(9) = c(9) + d(9) * T2;
Beta2(10) = c(10) + d(10) * T2;
Beta2(11) = c(11) + d(11) * T2;
Beta2(12) = c(12) + d(12) * T2;
# Experimental Points
For j :=1 to No2
Do
# Molarity Calculated by setting Activity Coefficient to zero
m2(1,j,1) = P2(j)*Fugacity_Coefficient2(1,j)/H2/1000/Pyvot_2(j);
m2(2,j,1) = 1;
m2(3,j,1) = 1000 * (W_MEA1/MW_MEA)/(1-W_MEA1) - m2(8,j,1) - m2(9,j,1);
m2(4,j,1)*m2(5,j,1)*10^14 = K2(1)*10^14 ;
m2(5,j,1)*m2(3,j,1)*10^10 = K2(4)*m2(8,j,1)* 10^10;
m2(6,j,1)*m2(5,j,1)*10^7 = K2(2)*m2(1,j,1)*10^7;
m2(7,j,1)*m2(5,j,1)*10^11 = K2(3)*m2(6,j,1)*10^11 ;
m2(8,j,1) = m2(6,j,1) + 2 * m2(7,j,1);
m2(9,j,1) * K2(5) * 10^3 = m2(3,j,1) *m2(6,j,1) * 10^3;
# Iteration to Convergy
For l := 2 to Iteration
Do
I2(j,l-1) = 0.5 * (m2(4,j,l-1) + m2(5,j,l-1) + m2(6,j,l-1) + 4 * m2(7,j,l-1) + m2(8,j,l-1)+ m2(9,j,l-1));
Loading2(j,l-1) = (m2(1,j,l-1) + m2(6,j,l-1) + m2(7,j,l-1))/(1000 * (W_MEA1/MW_MEA)/(1-W_MEA1));
# Activity Coefficient Calculated from Molarity
Gamma2(1,j,l-1) = Exp(2 * (Beta2(1) * m2(8,j,l-1) + Beta2(4) * m2(3,j,l-1) + Beta2(7) * m2(6,j,l-1) +
Beta2(8) * m2(7,j,l-1)+Beta2(9)*m2(9,j,l-1)));
Gamma2(2,j,l-1) = 1;
Gamma2(3,j,l-1) = Exp(2 * (Beta2(4) * m2(1,j,l-1) + Beta2(5) * m2(6,j,l-1) + Beta2(6) * m2(7,j,l-1)));
Gamma2(4,j,l-1) = 1;

Appendices

231

Gamma2(5,j,l-1) = 1;
Gamma2(6,j,l-1) = Exp(-A2 * I2(j,l-1) ^ 0.5 / (1 + B * I2(j,l-1) ^ 0.5) + 2 * (Beta2(2) * m2(8,j,l-1) +
Beta2(5) * m2(3,j,l-1) + Beta2(7) * m2(1,j,l-1)+Beta2(10)*m2(9,j,l-1)));
Gamma2(7,j,l-1) = Exp(-A2 * 4 * I2(j,l-1) ^ 0.5 / (1 + B * I2(j,l-1) ^ 0.5) + 2 * (Beta2(3) * m2(8,j,l-1) +
Beta2(6) * m2(3,j,l-1) + Beta2(8) * m2(1,j,l-1)+Beta2(11)*m2(9,j,l-1)));
Gamma2(8,j,l-1) = Exp(-A2 * I2(j,l-1) ^ 0.5 / (1 + B * I2(j,l-1) ^ 0.5) + 2 * (Beta2(1) * m2(1,j,l-1) +
Beta2(2) * m2(6,j,l-1) + Beta2(3) * m2(7,j,l-1)+Beta2(12)*m2(9,j,l-1)));
Gamma2(9,j,l-1) = Exp(-A2 * I2(j,l-1) ^ 0.5 / (1 + B * I2(j,l-1) ^ 0.5) + 2 * (Beta2(9) * m2(1,j,l-1) +
Beta2(10)* m2(6,j,l-1) + Beta2(12) * m2(8,j,l-1)));

# New Molarity Calculated from Activity Coefficient


m2(1,j,l)*Gamma2(1,j,l-1) = P2(j)*Fugacity_Coefficient2(1,j)/H2/1000/Pyvot_2(j); #*Gamma2(1,j,l-1)
m2(2,j,l) = 1;
m2(3,j,l) = 1000 * (W_MEA1/MW_MEA)/(1-W_MEA1) - m2(8,j,l) - m2(9,j,l);
m2(4,j,l)*m2(5,j,l)*10^14 = K2(1)*10^14 ;
m2(5,j,l)*m2(3,j,l)*Gamma2(3,j,l-1)*10^10 = K2(4)* (m2(8,j,l)*Gamma2(8,j,l-1))*10^10;
m2(6,j,l)*m2(5,j,l)*Gamma2(6,j,l-1)*10^7 = K2(2) *10^7 *(m2(1,j,l)*Gamma2(1,j,l-1));
m2(7,j,l)*Gamma2(7,j,l-1)*m2(5,j,l)*10^11 = K2(3)*10^11*(m2(6,j,l)*Gamma2(6,j,l-1));
m2(8,j,l) = m2(6,j,l) + 2 * m2(7,j,l);
m2(9,j,l)*Gamma2(9,j,l-1) * K2(5) * 10^3 = m2(3,j,l)*Gamma2(3,j,l-1) *m2(6,j,l)*Gamma2(6,j,l-1) *
10^3;
End
End
#~~~~~~~~~~~~~~~~~~~~~~~ For Parameter Estimization ~~~~~~~~~~~~~~~~~~~~~~~#

For ii := 1 to No1
Do

Loading_E1(ii) = Loading1(ii,Iteration1);
End

For ii := 1 to No2

Appendices

232

Do
Loading_E2(ii) = Loading2(ii,Iteration1);
End

ASSIGN
# Experimental Data

P1
:=
[3.89,19.05526,71.44,100.94,159.34,161.4635,211.858,296.9635,408.10668,19914,14945,9969,598
6,2992,993,593,293,103,36.1,8.09,2.57,0.604,0.0677,0.00896,0.00147]; # 25
P2
:=
[17.38,44.3922,76.5693,90.5652,91.8595,121.5595,171.1595,186.4705,188.7595,190.9755,236.759
5,295.1426,17723,14741,9770,5809,2804,822,422,222,122,46.8,2.29,0.0984,0.0221,0.00202,7.354,9
.314,9.045,15.51,19.62,25.2,27.71,39.18,40.4,43.49,51.82,58.57,62.88,74.95,77.59,83.61,92.79,137.
9,191.9,71.8769,102.714,128.389,128.389,167.806,262.183,191.843,229.332,229.332,286.658,342.6
76,358.314,374.665,447.881,391.763,468.32,468.32,468.32]; # 45
Vapour_fraction1(1,)
:=
[0.3312,0.7056,0.8996,0.9268,0.9523,0.9529,0.9637,0.9738,0.9808,0.9957,0.996333333,0.9969,0.9
97666667,0.997333333,0.993,0.988333333,0.976666667,0.936363636,0.171904762,0.035173913,0.
011681818,0.00287619,0.000225667,0.0000448,0.00000735];
Vapour_fraction2(1,)
:=
[0.0964,0.2085,0.3123,0.3494,0.3526,0.4188,0.5036,0.5250,0.5280,0.5309,0.5839,0.6362,0.984611
111,0.982733333,0.977,0.968166667,0.934666667,0.822,0.703333333,0.555,0.406666667,0.133714
286,0.007633333,0.000246,6.31429E-05, 0.00000808,0.042799609,0.053595268, 0.052128058,
0.086176242,0.106578304,0.13286234,0.144187741,0.192388903,0.197198223,0.209126755,0.239
585741,0.262598637,0.276577963,0.3130482,0.320540362,0.337028378,0.360685688,0.456063763
,0.538485282,0.304128979,0.384446091,0.438413653,0.438413653,0.505035122,0.614525493,0.53
8426564,0.582368357,0.582368357,0.635439065,0.675708291,0.685408991,0.694950151,0.731424
157,0.704327221,0.740099245,0.740099245,0.740099245];
# Optimised parameters
c
:=
[-9.828144E-01,-1.969601E-01,-3.275116E-01,-1.700729E-01,0.000000E+00,-2.026647E01,0.000000E+00,4.925087E-01,0.000000E+00,-1.557938E-01,0.000000E+00,0.000000E+00];
d
:=
[3.153199E-03,4.311135E-04,0.000000E+00,2.098788E-04,0.000000E+00,0.000000E+00,
0.000000E+00,0.000000E+00,0.000000E+00,0.000000E+00,0.000000E+00,0.000000E+00];
W_MEA1 := 0.30;

Appendices

233

III Calibration of the synthetic apparatus


1) Calibration of the High Pressure Generator
In order to establish the linear relation existing between the number of turns corresponding to
a particular piston stroke and the volume displaced by the piston, it is necessary to execute
calibration measurements on the pump. The principle of these experiments is based upon
weighing the different portions of a suitable liquid which is successively expelled from the
pump. By means of the density of the liquid, which has to be very well known under the
defined state conditions of the calibration run, the volume of the liquid samples can be
determined. This enables us to establish the relation between the cumulative number of turns
of the piston and the corresponding volume displaced.
Since water is the solvent being used in the solubility experiments as well as a liquid with
very well known thermo-physical properties, it was also employed in the calibration runs.
Prior to filling the pump, analytical grade water was thoroughly degassed in an ultrasonic
bath under reduced pressure at around 40C. The experiments were carried out at ambient
temperature. In order to minimise the risk of formation of bubbles within the pump cylinder,
the water had been pressurised up to

= 5.0 MPa to collapse possible bubbles that may have

been formed during the filling of the pump. Moreover, the calibration runs were performed
under an elevated pressure of

= 1.0 MPa. Three series of calibration measurements

covering the whole volume range of the pump, each containing between 6 to 8 single points,
have been executed.
A Mettler-Toledo balance (Model PR5003, 0.001 g readability up to 1010 g, 0.01 g
readability with loads up to 5100 g, Mettler-Toledo Ltd., Beaumont Leys Leicester, UK) had
been used to perform the weighing. The weighed values of the water samples had been
buoyancy corrected, using the density of air as derived from the BIPM formula of humid air 7.
The density of water needed to perform the buoyancy correction and used for converting the
masses to volumes had been calculated by means of the Wagner-Pruss equation of state
[269]. The averaged slope of the straight regression lines drawn through the plots of V versus
N, where V is the volume and N is the number of turns of the handle connected to the piston,
was found to be (0.7000 0.0002) cm3.

Davis, R. S., Equation for the Determination of the Density of Moist Air (1981/91), Metrologia 29
(1992) 67-70

Appendices

234

2) Cell volume calibration


Firstly, the whole system was dried and evacuated. Next degassed, deionised water was
drawn from the liquid reservoir into the high pressure generator. Afterwards, the valve
between the injection pump and the cell was opened to enable the flow of the water into the
cell. The liquid was injected step by step while the pressure inside the cell was recorded
together with the volume of water injected. Then, the total pressure in the cell was plotted
against the cumulative amount of water injected. The cell volume was calculated from the
knowledge of the calibration of the pump.
Since the volume of each stroke of the high pressure generator was known to be 0.7000
0.0002 ml, the cell volume can be calibrated against it by injecting pure water into a vacuum
without filling CO2 in advance. Experiment was conducted at about 100 C and the total
pressure and volume of water in each injection were recorded. Then the pressure measured by
P1 was plotted against cumulative moles of water injected, and is shown in Fig. III.1.
6
5

y = 382.32x - 1384.1

P/MPa

4
Two phase region

One phase region

Two phase region average


Linear (One phase region)

1
0
0

n/mol

Figure III.1 Cell Volume Calibration


In the above figure, the blue points represent the components inside the cell exist both in
liquid phase and vapour phase, whereas the green points show that the content reached allliquid phase which led to the sharp increase in the total pressure. The points in the two phase
region are correlated by taking the average of all the measured data and the one phase region
is simulated by linear regression. These two branches intersect at the bubble point of the pure

Appendices

235

water. By calculating the volume of the water in the cell at the bubble point, the volume of
the cell can be determined. The procedures are as follows.
The amount of water injected each time was recorded in volume

, and it can be

converted to number of moles according to the following Equation 3.1:


(Equation 3.1)
where

was calculated from the Wagner-Pruss equation of state [269] at the pump

temperature measured by
and

and the average pressure

between two equilibrium states

was the molecular weight of H2O.

The intersection of the two branches was found by solving the two correlated equations
simultaneously. Once the bubble point was located, known as BP (

BP,

BP),

the volume of

the water in the cell could be calculated by reversing equation 3.1.

(Equation 3.2)
where

was also calculated by using the Wagner-Pruss equation of state [269] at the cell

temperature measured by

and the bubble point total pressure.

and

were

respectively the moles of water in the cell and the total pressure at the bubble point.
The cell volume was found to be (67.9570.0002) ml at 98.03 C. This volume can be
converted to that at different temperatures by reference to the thermal expansion of the cell
material (Appendix VI).

Appendices
IV Experimental procedure of the synthetic apparatus

236

Appendices

237

Execution of solubility measurements using the high-pressure


solubility cell according to the synthetic method
Refer to the attached diagram.
A) Filling the high pressure pump with solvent
a) Evacuate the line attached to the pump by opening valves V3 and V1 (with all other
valves closed) and apply vacuum via the line attached to V3.
b) Close valves V3 and V1 and slowly open V5 (with V6 also open) to allow solvent to
fill the lines up to V1 until the pressure reached its atmospheric value. Then unwind
the piston to fill the cylinder.
c) After the cylinder is completely filled, close valve V5.
B) Filling the cell with gas
a) Start the vacuum pump while having all valves closed.
b) Open valves V2, V4 and V3 to evacuate the cell and the dead volume up to the gas
cylinder valves. Since at this stage the pressure can not be sensed, ensure that a
sufficient long period of time is allowed for this process.
c) Close V4 and slowly open V7 or V8 to pressurize the system. In order to prevent
condensation in cases where the critical temperature is above ambient, the
temperature of the cell should be higher than that of the cylinder. Close V7 or V8.
d) After filling wait for thermal and hydrostatic equilibrium to be achieved and measure
the initial filling pressure p2 of the gas as V2 is closed.
C) Execution of solubility measurements
a) Evacuate the dead volume by opening V3 (while having V1, V2 and V4 closed) and
applying vacuum.
b) Close V3 and open V1 to allow solvent into the dead volume. Now advance the piston
of the liquid injector to allow water to raise the liquid pressure p1 to an initial value
just greater than the gas pressure p2. Then open valve V4 by one turn.
c) In the actual experimental run, solvent is injected in steps, while stirring the cell
contents, and the equilibrium pressures achieved after each injection are recorded. A
typical step size is 10 revolutions of the pump, or the amount needed to increase the
pressure by, say 10%, whichever is the smaller. During any long pause in the
measurements, V1 might be closed.
D) Draining the cell
a) It is not recommended to disconnect any of the high pressure fittings.
b) Disconnect the vacuum pump (11) and put the end of the hose connected to V3 into a
suitable receiver inside the fume cupboard. With valves V1 and V2 closed, open V4
and V3 to relieve the system pressure. Then heat the cell to force out the bulk of the
solvent under its own vapour pressure.

Appendices

238

c) Further cleaning of the cell may be achieved by alternatively evacuating and then
back filling with pure water through valve V3. This is especially important for less
volatile solvents. Periodically close V3 and admit some gas through V2 and V7 or V8
up to a pressure of a few bar, then discharge slowly through V3 and continue cycles of
evacuation and flushing.
d) Finally, dry the cell at an elevated temperature and under continuous evacuation.

Appendices

239

V Plots of experimental results from the synthetic apparatus

Total Pressure (MPa)

5
Two phase region

4
3
2

One phase region

1
0
0

Cumulative moles of water injected (mol)

V.1 Total Pressure against Cumulative Moles of Water injected at 333 K

Total Pressure (MPa)

5
4

Two phase region

3
2
One phase region

1
0
0

Cumulative moles of water injected (mol)

V.2 Total Pressure against Cumulative Moles of Water injected at 353 K

Appendices

240

5
5

Total Pressure (MPa)

4
Two phase region

4
3
3
2

One phase region

2
1
1
0
0

Cumulative moles of water injected (mol)

V.3 Total Pressure against Cumulative Moles of Water injected at 373 K

9
8

Total Pressure (MPa)

7
Two phase region

6
5
4
3

One phase region

2
1
0
0

Cumulative moles of water injected (mol)

V.4 Total Pressure against Cumulative Moles of Water injected at 393 K

Appendices

241

VI Thermal expansion of Hastelloy8

/ C
21
93
204
316
427
538

T/K
294.15
366.15
477.15
589.15
700.15
811.15

L/10-6 C V/10-6 C
11.2
11.2
12
12.8
13.2
13.4

33.6
33.6
36
38.4
39.6
40.2

Approximation of volumetric
thermal expansivity of hastelloy
41
40
<V> over [294.15 K,
366.15 K]
<V> over [294.15 K,
477.15 K]
<V> over [294.15 K,
589.15 K]
<V> over [294.15 K,
700.15 K]
<V> over [294.15 K,
811.15 K]

39

V / 10-6 K-1

38
37
36
35
34
33
200 300 400 500 600 700 800 900
T/K

Technical data Blue sheet for Nickel-Base Alloy Allegheny Ludlum AL 276TM (UNS
Designation N10276), Allegheny Ludlum, Pittsburgh, PA
8

Appendices

242

VII Sample VBA code for calculation of the Henrys constant from the synthetic apparatus

Sub PressureCalculation()
Set objExcel = CreateObject("Excel.Application")
objExcel.Application.Visible = True
Set
Workbook
=
objExcel.Workbooks.Open("C:\Documents
Settings\dt207\Desktop\Thesis\Synthetic method\simulation Master Version")
Set Worksheet = Workbook.Worksheets("40-4bar")

'parameters
R = 8.314
T = Worksheet.Cells(2, 2).Value
n2 = Worksheet.Cells(24, 2).Value
Vm_CO2 = Worksheet.Cells(25, 2).Value
Bij = Worksheet.Cells(26, 2).Value
Bii = Worksheet.Cells(27, 2).Value
Bjj = Worksheet.Cells(28, 2).Value
Theta = Worksheet.Cells(31, 2).Value
Pw_s = Worksheet.Cells(32, 2).Value
Kappa = Worksheet.Cells(35, 2).Value
Vm_H2O = Worksheet.Cells(37, 2).Value
Vcell = Worksheet.Cells(39, 2).Value

Dim KH(25), Overallobj(25)


For j = 0 To 9
KH(j) = Worksheet.Cells(20 + j, 5).Value

and

Appendices

243

Dim n1(11)
Dim Z1(11), Z2(11), Z(11), Phi_s(11), Phi1(11), Phi2(11), K1(11), K2(11), Bm(11),
Beta(11), nL(11), x1(11), x2(11), y1(11), y2(11), VL(11), P1(11), P(11), P_calculation(11)
Dim obj(11)

'Loop to calculate point 1 to 9


For i = 0 To 7
'Initial value
n1(i) = Worksheet.Cells(5 + i, 6).Value
P(i) = Worksheet.Cells(5 + i, 9).Value
y1(i) = 0.5
y2(i) = 1 - y1(i)
Z1(i) = n1(i) / (n1(i) + n2)
Z2(i) = n2 / (n1(i) + n2)
Bm(i) = y1(i) ^ 2 * Bjj + 2 * y1(i) * y2(i) * Bij + y2(i) ^ 2 * Bii
Z(i) = 1 + (Bm(i) * P(i) / R / T)
Phi_s(i) = Exp(Bm(i) * Pw_s / R / T)
Phi1(i) = Exp(P(i) / R / T * (2 * (y1(i) * Bjj + y2(i) * Bij) - Bm(i)))
Phi2(i) = Exp(P(i) / R / T * (2 * (y1(i) * Bij + y2(i) * Bii) - Bm(i)))
K1(i) = (Pw_s * Phi_s(i) * Exp(Vm_H2O / R / T * (P(i) - Pw_s))) / (P(i) * Phi1(i))
K2(i) = (KH(j) * Exp(Vm_CO2 / R / T * (P(i) - Pw_s))) / (P(i) * Phi2(i))
Beta(i) = ((1 - K2(i)) + Z1(i) * (K2(i) - K1(i))) / ((K1(i) - 1) * (K2(i) - 1))
nL(i) = (1 - Beta(i)) * (n1(i) + n2)
x2(i) = Z2(i) / (1 + Beta(i) * (K2(i) - 1))
x1(i) = 1 - x2(i)
VL(i) = (x1(i) * nL(i) * Vm_H2O + x2(i) * nL(i) * Vm_CO2) * (1 - P(i) * Kappa)
P_calculation(i) = Beta(i) * (n1(i) + n2) * R * T * Z(i) / (Vcell - VL(i))

Appendices

244

'Iteration loops for y1 and P


Dim Accuracy, Step
Accuracy = Worksheet.Cells(41, 2).Value
Step = Worksheet.Cells(42, 2).Value
obj(i) = P_calculation(i) - P(i)

Do While Abs(obj(i)) >= Accuracy: Rem


If obj(i) < 0 Then
P1(i) = P(i) - Step: Rem
Else
P1(i) = P(i) + Step: Rem
End If: Rem

P(i) = P1(i): Rem


y1(i) = x1(i) * K1(i): Rem
y2(i) = 1 - y1(i): Rem
Z1(i) = n1(i) / (n1(i) + n2): Rem
Z2(i) = n2 / (n1(i) + n2): Rem
Bm(i) = y1(i) ^ 2 * Bjj + 2 * y1(i) * y2(i) * Bij + y2(i) ^ 2 * Bii: Rem
Z(i) = 1 + (Bm(i) * P(i) / R / T): Rem
Phi_s(i) = Exp(Bm(i) * Pw_s / R / T): Rem
Phi1(i) = Exp(P(i) / R / T * (2 * (y1(i) * Bjj + y2(i) * Bij) - Bm(i))): Rem
Phi2(i) = Exp(P(i) / R / T * (2 * (y1(i) * Bij + y2(i) * Bii) - Bm(i))): Rem
K1(i) = (Pw_s * Phi_s(i) * Exp(Vm_H2O / R / T * (P(i) - Pw_s))) / (P(i) * Phi1(i)) : Rem
K2(i) = (KH(j) * Exp(Vm_CO2 / R / T * (P(i) - Pw_s))) / (P(i) * Phi2(i)) : Rem
Beta(i) = ((1 - K2(i)) + Z1(i) * (K2(i) - K1(i))) / ((K1(i) - 1) * (K2(i) - 1)): Rem
nL(i) = (1 - Beta(i)) * (n1(i) + n2): Rem

Appendices

245

x2(i) = Z2(i) / (1 + Beta(i) * (K2(i) - 1)): Rem


x1(i) = 1 - x2(i): Rem
VL(i) = (x1(i) * nL(i) * Vm_H2O + x2(i) * nL(i) * Vm_CO2) * (1 - P(i) * Kappa): Rem
P_calculation(i) = Beta(i) * (n1(i) + n2) * R * T * Z(i) / (Vcell - VL(i)): Rem
obj(i) = P_calculation(i) - P(i): Rem
Loop

Worksheet.Cells(44 + i, 2).Value = x1(i)


Worksheet.Cells(44 + i, 3).Value = x2(i)
Worksheet.Cells(44 + i, 4).Value = y1(i)
Worksheet.Cells(44 + i, 5).Value = y2(i)
Worksheet.Cells(44 + i, 6).Value = Z1(i)
Worksheet.Cells(44 + i, 7).Value = Z2(i)
Worksheet.Cells(44 + i, 8).Value = Z(i)
Worksheet.Cells(44 + i, 9).Value = Bm(i)
Worksheet.Cells(44 + i, 10).Value = Phi_s(i)

Worksheet.Cells(57 + i, 1).Value = Phi1(i)


Worksheet.Cells(57 + i, 2).Value = Phi2(i)
Worksheet.Cells(57 + i, 3).Value = K1(i)
Worksheet.Cells(57 + i, 4).Value = K2(i)
Worksheet.Cells(57 + i, 5).Value = Beta(i)
Worksheet.Cells(57 + i, 6).Value = nL(i)
Worksheet.Cells(57 + i, 7).Value = VL(i)
Worksheet.Cells(57 + i, 8).Value = P(i)
Worksheet.Cells(57 + i, 9).Value = P_calculation(i)
Worksheet.Cells(57 + i, 10).Value = obj(i)

Appendices

246

Next
Overallobj(j) = (P(0) - Worksheet.Cells(5, 9).Value) ^ 2 + (P(1) - Worksheet.Cells(6,
9).Value) ^ 2 + (P(2) - Worksheet.Cells(7, 9).Value) ^ 2 + (P(3) - Worksheet.Cells(8,
9).Value) ^ 2 + (P(4) - Worksheet.Cells(9, 9).Value) ^ 2 + (P(5) - Worksheet.Cells(10,
9).Value) ^ 2 + (P(6) - Worksheet.Cells(11, 9).Value) ^ 2 + (P(7) - Worksheet.Cells(12,
9).Value) ^ 2
Worksheet.Cells(20 + j, 6).Value = Overallobj(j)
Next
End Sub

You might also like