Sei sulla pagina 1di 282

Green Energy and Technology

Wojciech M. Budzianowski Editor

Energy Efficient
Solvents for CO2
Capture by Gas–
Liquid Absorption
Compounds, Blends and Advanced
Solvent Systems
Green Energy and Technology
More information about this series at http://www.springer.com/series/8059
Wojciech M. Budzianowski
Editor

Energy Efficient
Solvents for CO2 Capture
by Gas–Liquid Absorption
Compounds, Blends and Advanced Solvent
Systems

123
Editor
Wojciech M. Budzianowski
Consulting Services
Wrocław
Poland

and

Renewable Energy and Sustainable


Development (RESD) Group
Wrocław
Poland

ISSN 1865-3529 ISSN 1865-3537 (electronic)


Green Energy and Technology
ISBN 978-3-319-47261-4 ISBN 978-3-319-47262-1 (eBook)
DOI 10.1007/978-3-319-47262-1
Library of Congress Control Number: 2016956846

© Springer International Publishing AG 2017


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer International Publishing AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Greenhouse gas emissions limit the expansion of many industries and especially
considering the whole life cycle it may be difficult and costly to remarkably reduce
existing CO2 sources. It is therefore expected that CO2 separation from various
gases may be required in the future in order to mitigate emissions and associated
climate change. CO2 capture may be required not only in coal-fired power plants
but also in many other industries. Besides, it may be needed not only for fossil fuel
processing but also for biomass processing. In addition, by efficiently separating
and concentrating CO2 a new carbon resource may be created for CO2 utilisation
technologies. Consequently, CO2 capture is a very important process that will play
a significant role in the future. Its high efficiency achieved by applying innovative
techniques will greatly contribute to commercial success of many CO2 emitting and
CO2 utilising industries.
This edited book is dedicated to developments of solvents for CO2 capture by
gas–liquid absorption. The entire process involving these solvents needs to be
characterised by very high energy efficiency to enable implementation of this
technology in actual industries. Book chapters put emphasis on compounds, their
blends and advanced solvent-based capture processes (ASBCPs). Single compound
solvents are usually simple and easy to handle. Solvent blends are used for
enhancing CO2 capturing processes and are often characterised by reduced energy
requirements. Finally, emerging innovative advanced solvent-based capture pro-
cesses make use of novel phenomena and innovative approaches capable of creating
disruptive innovations. The ASBCPs involve for instance two immiscible liquid
phases or microencapsulated solvents which have very favourable operating char-
acteristics. Therefore, ASBCPs may greatly contribute to the increase of energy
efficiency of CO2 capture plants.
The book includes a selection of Chapters dedicated to energy efficient solvents
for CO2 capture by gas–liquid absorption. It also discusses the whole technical
context of using CO2 capture solvents in practical applications. Chapter
“Introduction to Carbon Dioxide Capture by Gas–liquid Absorption in Nature,

v
vi Preface

Industry, and Perspectives for the Energy Sector and Beyond” introduces CO2
capture by gas–liquid absorption in nature, industry and discusses challenges in the
energy sector. Chapter “Assessment of Thermodynamic Efficiency of Carbon
Dioxide Separation in Capture Plants by Using Gas–liquid Absorption” assesses
thermodynamic efficiency of a CO2 separation process relying on gas–liquid
absorption. In Chapter “Process Implications of CO2 Capture Solvent Selection” the
process implications in the selection of energy efficient solvents with the overall
message that the process should not limit the choice of solvent but be designed to
make best use of the advantages of the energy efficient solvent choice and minimise
the disadvantages. Chapter “Useful Mechanisms, Energy Efficiency Benefits, and
Challenges of Emerging Innovative Advanced Solvent Based Capture Processes”
analyses advanced solvent-based capture processes for energy efficient carbon
dioxide capture by gas–liquid absorption. Chapter “Phase Change Solvents for CO2
Capture Applications” investigates solvent systems that separate into two phases
upon absorption of CO2 having significant potential as energy efficient solvents due
to only the CO2 rich stream being heated in the regenerator resulting in reduced
sensible and latent heating requirements. Chapter “Aqueous Amino Acid Salts and
Their Blends as Efficient Absorbents for CO2 Capture” is dedicated to aqueous
amino acid salts and their blends having high CO2 loading, high reaction rate, being
less corrosive, less toxic, and requiring less regeneration energy compared to
commercial amines. Further, Chapter “Ionic Liquids: Advanced Solvents for CO2
Capture” relates to ionic liquids (ILs) which have become more attractive for CO2
capture because of their excellent properties and potential energy saving efficiency.
It reviews and analyses the research progress on CO2 capture with ILs including the
absorption capacity, the absorption mechanism and the process simulation and
assessment. Furthermore, Chapter “Amine-Blends Screening and Characterization
for CO2 Post-combustion Capture” provides the CO2 loading and heat of absorption
experimental data of amine blends for CO2 capture which can be useful for future
industrial application in the selection of amine blend solvents. Chapter “Post-
combustion Carbon Dioxide Capture with Aqueous (Piperazine + 2-Amino-2-
Methyl-1-Propanol) Blended Solvent: Performance Evaluation and Analysis of
Energy Requirements” critically discusses the properties of the AMP+PZ blended
solvent characterised by a regeneration energy demand of 2.9–3.7 GJ/tCO2. Chapter
“Energy Efficient Absorbents for Industry Promising Carbon Dioxide Capture”
reviews and analyses energy efficient absorbents for industrially relevant carbon
dioxide capture systems. Chapter “The Absorption Kinetics of CO2 into Ionic
Liquid—CO2 Binding Organic Liquid and Hybrid Solvents” is dedicated to
absorption kinetics of CO2 into ionic liquids and hybrid solvents as measure to
achieve higher efficiency energy utilisation in carbon capture. Finally, Chapter
“Solubility of Carbon Dioxide in Aqueous Solutions of Linear Polyamines” pro-
vides information on the solubility of carbon dioxide in aqueous solutions of linear
polyamines. Overall, this book provides a useful engineering resource for the
Preface vii

development of energy efficient CO2 capture solvent systems involving gas–liquid


absorption. The invited leading authors from different continents give their own
perspectives on the associated problems and hence the whole picture is well bal-
anced and up-to-date.

Wrocław, Poland Wojciech M. Budzianowski


Contents

Introduction to Carbon Dioxide Capture by Gas–Liquid


Absorption in Nature, Industry, and Perspectives for the
Energy Sector and Beyond . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Wojciech M. Budzianowski
Assessment of Thermodynamic Efficiency of Carbon Dioxide
Separation in Capture Plants by Using Gas–Liquid Absorption . . . . . . . 13
Wojciech M. Budzianowski
Process Implications of CO2 Capture Solvent Selection . . . . . . . . . . . . . . 27
Leigh T. Wardhaugh and Ashleigh Cousins
Useful Mechanisms, Energy Efficiency Benefits, and Challenges of
Emerging Innovative Advanced Solvent Based Capture Processes . . . . . 69
Wojciech M. Budzianowski
Phase Change Solvents for CO2 Capture Applications. . . . . . . . . . . . . . . 99
Kathryn A. Mumford, Kathryn H. Smith and Geoffrey W. Stevens
Aqueous Amino Acid Salts and Their Blends as Efficient
Absorbents for CO2 Capture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
Azmi Mohd Shariff and Muhammad Shuaib Shaikh
Ionic Liquids: Advanced Solvents for CO2 Capture . . . . . . . . . . . . . . . . . 153
Xiangping Zhang, Lu Bai, Shaojuan Zeng, Hongshuai Gao,
Suojiang Zhang and Maohong Fan
Amine-Blends Screening and Characterization for CO2
Post-combustion Capture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
Abdullah Al Hinai, Nabil El Hadri and Mohammad Abu Zahra
Post-combustion Carbon Dioxide Capture with Aqueous
(Piperazine + 2-Amino-2-Methyl-1-Propanol) Blended Solvent:
Performance Evaluation and Analysis of Energy Requirements . . . . . . . 191
Sukanta K. Dash

ix
x Contents

Energy Efficient Absorbents for Industry Promising


Carbon Dioxide Capture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
Y.S. Yu, T.T. Zhang and Z.X. Zhang
The Absorption Kinetics of CO2 into Ionic Liquid—CO2
Binding Organic Liquid and Hybrid Solvents . . . . . . . . . . . . . . . . . . . . . 241
Ozge Yuksel Orhan, Cyril Sunday Ume and Erdogan Alper
Solubility of Carbon Dioxide in Aqueous Solutions
of Linear Polyamines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
Jian Chen, Ruilei Zhang, Zhongjie Du and Jianguo Mi
Introduction to Carbon Dioxide Capture
by Gas–Liquid Absorption in Nature,
Industry, and Perspectives for the Energy
Sector and Beyond

Wojciech M. Budzianowski

Abstract The study describes how the process of CO2 capture by gas–liquid
absorption is used in nature, industry, and discusses perspectives for its use in the
energy sector and beyond. The process is used in nature on a large scale greatly
contributing to the global carbon cycle. CO2 is captured from air by oceans and
stroma (aqueous fluid) of plant leaf cells during photosynthesis. Step-change bio-
logical developments are associated with the evolution of enzymes that catalyse
CO2 dissolution in stroma and incorporate it in the cell’s Calvin cycle. In industry
the first implementation of CO2 capture was in natural gas upgrading (in 1920s)
followed by its use in Enhanced Oil Recovery (EOR) (in 1970s). Developments
relied on using more efficient physical and chemical solvents and their blends. Only
relatively recently CO2 capture has been proposed as a countermeasure to mitigate
CO2 emissions. Perspectives for employing gas-liquid absorption for decarboni-
sation of the energy sector and modern industries are strengthened by recent solvent
based research outcomes. Developments in gas–liquid absorption are essential for
wide-scale deployment of CO2 capture and are therefore approached through all
Chapters of this edited book. Promising developments relate to e.g. energy efficient
solvents relying on innovative advanced solvent based capture processes.

1 Introduction

CO2 capture by gas–liquid absorption is an old process widely employed in nature,


industry, and in recent years considered for the mitigation of CO2 emissions from
the energy sector and beyond. In nature CO2 capture by gas–liquid absorption is an
important process of the global carbon cycle [1]. CO2 is absorbed from air by
oceans and transformed there into carbohydrates (mainly cellulose and chitin) and

W.M. Budzianowski
Consulting Services, Wrocław, Poland
W.M. Budzianowski (&)
Renewable Energy and Sustainable Development (RESD) Group, Wrocław, Poland
e-mail: wojciech.budzianowski@gmail.com

© Springer International Publishing AG 2017 1


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_1
2 W.M. Budzianowski

minerals (mainly calcium carbonate). The scale of CO2 absorption by oceans and
other hydrosphere was in 2009 about 338 PgCO2/yr [1, 2]. On land CO2 is mainly
absorbed by stroma (aqueous fluid) present in plant leaf cells via photosynthesis
thus contributing to biomass productivity. The scale of CO2 absorption by bio-
sphere and soils in 2009 amounted to about 450 PgCO2/yr [1, 2]. A remarkable part
of these natural CO2 transfers from the atmosphere was achieved by CO2 capture by
gas–liquid absorption meaning that this process is essential in natural conditions for
capturing CO2 from air.
In industry CO2 capture by gas–liquid absorption was historically applied to
natural gas upgrading. Other mostly niche applications related e.g. to chemical and
food industry. The scale of these processes was historically relatively small com-
pared to CO2 removals by oceans and biosphere as well as to anthropogenic CO2
emissions.
Only relatively recently CO2 capture has been proposed as a measure to coun-
teract CO2 atmospheric emissions from fossil fuel combustion with the aim to
mitigate climate change. The scale of CO2 emissions from fossil fuel combustion is
big enough to seriously affect the global carbon cycle and in 2009 it amounted to 29
Pg CO2/yr [2, 3]. This newly emerged CO2 capture application has attracted sig-
nificant global attention and R&D funding. CO2 capture by gas–liquid absorption is
a relatively well understood process but cost reduction is required in order to apply
CO2 capture on a large scale. In recent years many demonstration and pilot plant
projects have been funded all over the world. Research efforts have led to
remarkable development of this process by improving its fundamental under-
standing, optimising performance under conditions seen in the energy sector and
beyond, and creating new innovative processes.
Historical experience gained from applying CO2 capture by gas–liquid absorp-
tion processes in various natural and industrial processes can be employed to create
new innovative processes that meet present-day challenges. The scale of CO2
capture required for applications in the energy sector and beyond is much greater
than ever before in the industrial history of this process and it is therefore essential
that the gas–liquid absorption process is much more efficient than it was previously.
In some applications absorption is physical thus if not accompanied by
high-pressure it is characterised by relatively small CO2 fluxes. In other applications
absorption is chemical meaning that CO2 dissolved in the liquid phase undergoes
one or more chemical reactions. For example, living plants utilise diluted CO2 from
air. Since physical absorption from air would be very slow the plants evolved
complex CO2 fixation mechanisms relying on chemical absorption that enables
concentration of CO2 in the stroma present in living cells at atmospheric pressure
and very low CO2 concentration in air. This natural process is driven by solar
energy with relatively low efficiency (3–6%). In addition, it requires a lot of land
per unit mass of fixed CO2.
Industry today requires relatively intensive CO2 capture processes that with
minimal energy penalty and without excessive land requirement would be able to
absorb CO2 from a range of industrial gases, including flue gases from the energy
sector. Conventionally, industrial CO2 capture by gas–liquid absorption has been
Introduction to Carbon Dioxide Capture by Gas–Liquid Absorption … 3

achieved by means of scrubbing/stripping columns using single solvents (physical


or chemical) or blended solvents but these processes consume energy and de-rate
power plants. Therefore, more energy efficient advanced solvent based capture
processes are urgently required [4, 5].
This chapter is structured as follows. Sect. 2 discusses CO2 capture by gas-liquid
absorption applied in nature while Sect. 3 in industry. Sect. 4 analyses requirements
for process innovations to meet present-day challenges in the energy sector and
beyond. Finally, Sect. 5 provides conclusions and outlook for the future.

2 CO2 Capture by Gas–Liquid Absorption in Nature

CO2 capture achieved by gas–liquid absorption is a very old natural process. In


nature it has been present at least since the creation of ocean and CO2 comprising
atmosphere. Since then CO2 has been constantly captured from air by oceanic
water. Dissolved CO2 is shifted to carbonate ions, mainly bicarbonate ions
(HCO3 ). The absorbed CO2 mainly accumulates in the surface layer because it
very slowly propagates through the oceanic water due to limited convective
motions in deep ocean. CO2 utilisation by living photosynthesising organisms is to
some extent possible at the ocean surface where sunlight is available but miner-
alisation processes (e.g. chitin and calcium carbonate formation) are not very much
affected due to shallow penetration of dissolved CO2. Currently, together with the
rise of CO2 content in the atmosphere above the pre-industrial average the rate of
CO2 absorption by the ocean increased. One unwanted consequence is surface
ocean acidification [6] that may eventually affect the stability of fragile marine
ecosystems.
The ocean CO2 capture process plays a significant role in removing CO2 from
the atmosphere. It was estimated that oceans/hydrosphere captured 8.4 Pg CO2/yr in
2009, i.e. 29% of anthropogenic CO2 emissions [1, 2]. It also contributes to rock
formation by precipitating minerals in ocean which naturally store CO2. In addition,
CO2 is the main carbon source for marine organisms.
Another very old natural process involving CO2 capture by gas–liquid absorp-
tion is the first step of photosynthesis. Terrestrial plants absorb CO2 from air
through permeable cell membranes to stroma (aqueous fluid) of leaf cells. After
permeating the cell wall CO2 dissolves physically in the stroma. Further, dissolved
CO2 reacts with H+ ions and electrons. CO2 also reacts directly with abundant water
via a reaction catalysed by an carbonic anhydrase enzyme [7, 8]. These reactions
concentrate CO2 within the cells in the reacted forms. Further, the products of these
reactions are shifted to organic compounds again by enzymatic catalysis completely
removing them from the stroma. More specifically, plants using the C3 carbon
fixation mechanism shift CO2-derived products by combining them with the
enzyme ribulose 1,5-bisphosphate (RuBisCo) yielding two molecules of glycerate
3-phosphate. CO2-derived carbon is then incorporated into the plant’s Calvin cycle.
Plants using the C4 carbon fixation approach employ an additional CO2
4 W.M. Budzianowski

concentration mechanism relying on reacting CO2 and its derivatives first with
phosphoenolpyruvate (PEP) in a reaction catalysed by an enzyme PEP carboxylase
yielding oxaloacetic acid. Oxaloacetic acid is subsequently translocated within the
cell and decarboxylated releasing CO2 products which then immediately react with
RuBisCo and as previously are incorporated into the plant’s Calvin cycle.
Interestingly, although C4 plants represent only approximately 5% of global plant
biomass, their contribution to CO2 uptake by all terrestrial plants is about 30% [9]
meaning that C4 fixation is globally more effective in capturing CO2. C4 plants
employed e.g. as energy crops could therefore enhance biological sequestration of
atmospheric CO2 and simultaneously deliver bioenergy. The current rate of biomass
production (about 130 TW [10]) is several times greater than global energy con-
sumption (15.1 TW in 2012 [11]) and could be further increased by using C4 plant
based energy cropping rotations.
CO2 capture for use in photosynthesis is also important for removing CO2 from
the atmosphere. It was estimated that soil and biosphere captured 10.3 Pg CO2/yr in
2009, i.e. 35% of anthropogenic CO2 emissions [1, 2]. Biological CO2 capture is a
main source of carbon for terrestrial plants thus contributing to biomass production
and biological solar energy harvesting. Two step-change developments include the
evolution of (a) very efficient carbonic anhydrase (CA) enzyme that catalyses CO2
dissolution in stroma and (b) mechanisms for biochemical CO2 concentration in the
stroma of living cells by C4 plants. CA has already been applied in carbon capture
demonstration projects for CO2 removal from industrial flue gases [12]. Several
other concepts of nature-inspired carbon fixation are explored for industrial appli-
cation but most are fairly complex and are at an early stage of development.
One example is biomimetic solar energy harvesting which requires some further
fundamental developments before achieving commercialisation potential [13].
Biomimetic materials that attempt to implement biological mechanisms into CO2
capture and utilisation also attract attention. The benefits of biomimetic materials are
associated with highly ordered architectures, lightweight, and capability to combine
strength and toughness. They can also be chemically and thermally inert, non-toxic,
multifunctional, capable to adapt biological self-repair and regeneration mecha-
nisms, energy efficient, low cost, and characterised by rapid mass transfer [14].

3 CO2 Capture by Gas–Liquid Absorption in Industry

Gas-liquid absorption is the most mature CO2 capture process in industry due to its
high efficiency and lower cost [15]. The first application of CO2 capture in industry
was CO2 removal from natural gas employed in the 1920s. The aim was to improve
the quality of natural gas by raising its calorific value. In addition, by separating
ballast CO2 less energy was required for gas compression and transportation via gas
grids. The separated CO2, possibly after some minor treatment, was released to the
atmosphere as it had no commercial value. The process incorporated gas–liquid
absorption using early physical solvents. The first known CO2 capture process
Introduction to Carbon Dioxide Capture by Gas–Liquid Absorption … 5

patent was filed in 1927, and it was essentially a series of absorption towers [16]. In
the 1930s physical solvents were used to separate CO2 and deliver it for food and
chemicals productions from gases comprising up to 25% CO2. In addition, CO2 was
separated from ambient air in cryogenic air separation plants producing N2, O2 and
Ar in order to prevent the equipment fouling by dry ice formation [17]. In the early
1940s, physical solvents were used for CO2 separation from CO2 rich streams (up
to 80%) at higher pressures (up to 10 MPa). In the 1950s CO2 capture was used to
separate CO2 in H2 production in refineries. Most of the CO2 captured during these
times was released to the atmosphere. First chemical solvents were also used during
this time.
In the early 1970s, CO2 captured from natural gas processing was for the first
time utilised in Enhanced Oil Recovery (EOR). The first CO2 utilisation facility was
erected in Texas (USA) where CO2 was transmitted to an oil field and injected thus
boosting oil recovery. EOR has since proven very successful and added value to
captured CO2. Following the success of EOR, Enhanced Gas Recovery (EGR) was
developed a few years later. Millions of tonnes of CO2 have been sequestered by
EOR and EGR since its creation. CO2 is now sourced not only from natural gas but
occasionally also from various industrial processes. From the very beginning
EOR/EGR included CO2 underground sequestration but it was not linked with CO2
emissions mitigation objectives during these early applications. However, the idea
of capturing CO2 to prevent it from being released into the atmosphere to mitigate
climate change was suggested a few years later, in 1977 [18].
In natural gas processing plants CO2 needs to be separated in order to raise the
energy content of the natural gas and minimise its transportation costs. The sepa-
rated CO2 often needs to be cleaned in order to obtain a marketable product. Natural
gas processing facilities have to therefore capture and purify the CO2 before they
produce natural gas and CO2 as useable commodities. The CO2 can be used in
EOR/EGR projects achieving a value in monetary terms. In EOR/EGR the CO2 is
partly extracted along with the oil/natural gas, separated and reused thus benefi-
cially creating a CO2 loop [19].
Other applications include the separation of CO2 from hydrogen for the use in
various hydrogenation processes and in producing beverage-grade CO2.
Industrial technological developments relied on using more efficient solvents and
their blends such as physical and later chemical solvents. Some process develop-
ments led to reduced energy penalty using physical solvents such as Selexol,
Rectisol, Purisol, Jeffsol, and Morphysorb. The physical solvent processes are
effective at high pressure where for example Purisol has much greater CO2 loading
capacity than MEA. They all employ chemicals that are non-reactive towards CO2
but with high physical solubility. One process was commercialised that employs a
combination of physical and chemical solvents (Sulfinol by Shell). After that
reactive processes were developed and commercialised with particular focus on low
pressure gases with lower concentrations of CO2. Among them several MEA and
MDEA based processes were deployed (e.g. Fluor’s Econamine, Lummus MEA).
A KM-CDR process employed hindered secondary amine KS-1. Besides, hot
potassium carbonate processes were developed and used in industrial practice (e.g.
6 W.M. Budzianowski

Benfield, Catacarb, Flexsorb, Vetrocoke). In addition ammonia based processes


were proposed on the market (Alstom’s Chilled Ammonia, ECO2). These processes
employed ammonia solvent and due to ammonia volatility [20] absorption was
realised at low temperatures (typically below 10 °C).
In further developments the gas–liquid absorption technology of the single
amines were gradually shifted to solvent blends [21, 22] such as MEA/MDEA and
others (Aker Clean Carbon, Alstom Advanced Amine, Cansolv, HTC PureEnergy,
Advanced PCCC). Additionally various reaction promoters such as piperazine were
added to amine solvents [23]. Subsequently more and more sophisticated solvents
were researched and demonstrated. For example Siemens proposed the POSTCAP
process involving an amino-acid salt solvent. The process benefited from low
environmental impact while technical parameters were comparable with Best
Available Technologies (BATs) at that time. Further, a few companies proposed
catalysed CO2 capture by carbonic anhydrase however the process has little com-
mercial success due quick denaturation of CA under realistic industrial conditions.
Other advanced solvent based capture processes (ASBCPs) included aminosili-
cones, chilled ammonia or potassium carbonate with phase change. More sophis-
ticated solvent systems using new approaches such as phase change solvents have
recently been researched or piloted [24] but as of 2015 were not commercially
available [25].

4 Perspectives for the Energy Sector and Beyond

The essential challenge that all commercial CO2 capture gas–liquid absorption
processes have to overcome is insufficient energy efficiency. State-of-the-art CO2
capture processes are capable of separating CO2 from flue gases with a total
equivalent work requirement of about 1 MJ/kgCO2 meaning that thermodynamic
efficiency of separating CO2 is about 16% (see Chap. 4 of this book for further
explanations on this topic). Therefore, the main challenge of CO2 separation is how
to reduce these energy requirements and raise efficiency.
Decarbonisation of the energy sector may be alternatively achieved by a shift
from fossil fuel to renewable energy sources. Harvesting renewables to meet all
global energy demands may however be challenging mainly due to high production
cost, limited generation potential and high energy demands. In addition, renewable
industries require very expensive infrastructures that are CO2 intensive by them-
selves (in the life cycle context). The increasing penetration of renewables will
likely rely on harvesting lower quality sources than at the beginning which will
have higher associated life cycle CO2 emissions. This might be a problem for some
renewable industries that is today often overlooked. Regarding the potential of
renewable energy, according to International Renewable Energy Agency (IRENA)
with policies in place and under consideration in 2014, the global penetration of
modern renewable energy (excluding traditional biomass) will reach 14% of total
final energy consumption (TFEC) by 2030 [26]. If an extended policy package
Introduction to Carbon Dioxide Capture by Gas–Liquid Absorption … 7

proposed in IRENA’s REmap 2030 [26] is adopted and successfully implemented,


the penetration of modern renewables (excluding traditional biomass) in the global
energy mix may theoretically reach about 27% by 2030. This will be insufficient to
decarbonise the whole energy sector. The further expansion of renewable energy
industries may put pressure on growing life cycle CO2 emissions thus truly effective
decarbonisation will be challenging. It also emphasises that decarbonisation of the
energy sector cannot increase emissions in cooperating industries and decarboni-
sation beyond the energy sector is equally important.
In recent long term projections it was suggested that under the scenario of
limiting CO2 concentration in the atmosphere to 450 ppm by 2100 the share of
fossil fuels in total primary energy supply (TPES) would decrease to about 44%
while the share of renewables would increase to about 50% in 2100 [19]. However,
to maintain this high share of fossil fuels coal based power plants might need up to
55% carbon capture utilisation and storage (CCUS) while natural gas fired power
plants up to 25% CCUS in 2100. CO2 separations would need to be implemented
throughout various niche industries that use fossil fuels or biomass and significant
potential lies in highly concentrated CO2 streams for which CO2 capture is feasible
at relatively low costs.

5 Technological Developments in Gas–Liquid Absorption

The needs identified in the energy sector and other modern industries imply that
there will be high potential for energy efficient CO2 capture technologies including
energy efficient options relying on gas–liquid absorption. In this context advanced
solvent based capture processes (ASBCPs) will need to be developed and
employed. Energy efficient ASBCPs will require simultaneous developments in
materials science, nanotechnology, catalysis, process engineering, systems research
and environmental science. Figure 1 presents the multiple scales that a particular
ASBCP for CO2 separation may entail. Most process innovations start with material
development that needs to have optimal properties for mass transfer and stability

Fig. 1 Steps across advanced solvent based capture process design


8 W.M. Budzianowski

over multiple scrubbing/stripping cycles. The material may be a phase changing


solvent, have catalytic properties or benefit from microencapsulation or polarity
swing. The specific physico-chemical properties of materials may increase CO2
loading capacity by simultaneously binding CO2 chemically and dissolving phys-
ically (e.g. ionic liquids). Further, mass and heat transfers enable linkage of these
favourable basic properties of the developed material with the requirement of the
separation process. In general, CO2 transfer needs to be sufficiently fast while
thermal effects should be limited to move the entire ASBCP closer to the reversible
process. Finally, the ASBCP needs to integrate with the energy system to make use
of synergies that are feasible due to the fact that energy is conserved. The ASBCP
also needs to integrate with the environment to minimise adverse environmental
impacts of the technology, in particular in the whole life cycle. All of these inno-
vation steps need to be achieved simultaneously in order to fully account for
interactions between all parts of the low-carbon energy system.
Historically, developments in processes for CO2 capture by gas–liquid absorp-
tion have been driven by some specific natural and industrial drivers. Developments
in nature were primarily driven by the need of sourcing CO2 from diluted air.
Therefore, plants evolved catalysed absorption (using CA) allowing for process
intensification by several orders making it effective despite the very low separation
driving force. The second driver for innovation was associated with the need to
colonise tropical arid regions by plants. In this way C4 plants capable of enriching
CO2 in stroma by enzymatically catalysing CO2 reaction with water (first
employing CA followed by PEP carboxylase) were much more effective in CO2
fixation. Due to temporal carbon fixation C4 plants were more active in capturing
CO2 during colder nights and CO2 conversion during hot days. C4 plants achieved
improved water efficiency. For example, at 30 °C C3 grasses lost approximately
833 molecules of H2O per one fixed molecule of CO2, whereas C4 grasses lost only
277 [27]. Due to the combination of intensified CO2 fixation and decreased water
losts C4 plants conserved soil moisture and easily colonised tropical arid environ-
ments being more efficient than C3 plants. However, C3 plants colonised colder
regions without water deficiency because the shortcoming of C4 plants relied on the
need to fix every CO2 molecule twice (first by PEP and second by RuBisCO)
implying that they used more energy than C3 plants (30 vs. 18 molecules of ATP
per one molecule of glucose). Therefore, C3 plants were efficient in regions with
lower solar illuminance while C4 plants were superior under high solar radiation
intensity and associated water shortages.
Historical industrial developments in CO2 capture by gas–liquid absorption were
driven by techno-economic factors associated with the need to improve quality and
hence value of natural gas. Later, along with EOR/EGR implementation, CO2
capture allowed an increase in natural gas production meaning that economic dri-
vers appeared. Developments led to physical solvents being replaced with chemical
solvents and solvent blends especially when low CO2 partial pressure were present.
Current drivers in the energy sector are similar to those seen in industry in the
past with the difference being that the goal of CO2 capture relates mainly to mit-
igating CO2 emissions and protecting climate. Capturing CO2 does not directly
Introduction to Carbon Dioxide Capture by Gas–Liquid Absorption … 9

contribute to the quality of energy products such as electricity from power plants or
liquid fuels from refineries. Therefore, external economic drivers implemented by
energy policy instruments are essential to make CO2 capture from industrial flue
gases an attractive business line.
CO2 capture by gas-liquid absorption is also applicable in the renewable energy
sector, e.g. for biomethane production by CO2 separation from biogas [28–30]. It
also has potential application in relation to various industrial CO2 rich sources, e.g.
in the cement industry. Since renewable energy harvesting usually requires large
backing industries (to provide construction materials and feedstocks) CO2 capture
in industry will be very important in the near future in order to minimise life cycle
CO2 footprint of renewables [31]. Drivers in these industries may be to some extent
economical without external subsidies especially if CO2 can be valorised. But in
industries cooperating with the renewable energy sector additional drivers must be
implemented by policymaking.

6 Conclusions and Outlook

CO2 capture by gas–liquid absorption has been used in nature for a very long time
and in industry for almost a century. In nature CO2 is absorbed by oceans and by
living plants. In 2009 the scale of these two natural absorption processes was as
high as 788 PgCO2/year while the contribution to mitigating anthropogenic CO2
emissions was also high amounting to 64% (18.7 PgCO2/yr). Natural process
developments for CO2 capture include catalysed absorption using CA in C3 plants
and combined CA/PEP carboxylase in C4 plants. The C3 mechanism is more energy
efficient in CO2 capture because it is a one step process while in the C4 mechanism
the second step consumes additional energy. Industrial developments started from
physical solvents and were followed by chemical solvents and solvent blends.
Advanced solvent based capture processes are required to improve energy effi-
ciency of the CO2 capture process and give promise for the implementation of CO2
capture to reduce emissions in the energy sector and beyond. Drivers for imple-
mentation of CO2 capture in industry are mainly economical. In the energy sector
CO2 mitigation strategies need to carefully consider the whole life cycle and a
balance between renewable, biomass and fossil fuel sources need to be found. Since
the expanding renewable energy sector has large associated industries, it is
important that CO2 emissions in these industries are minimised, potentially by
involving CO2 capture. Technological developments are essential in achieving true
decarbonisation throughout all economic sectors and therefore all chapters of this
book provide insights into energy efficient processes for CO2 capture by gas–liquid
absorption.

Acknowledgments This study has been supported by the members of the Renewable Energy and
Sustainable Development (RESD) Group (Poland) under the project RESD-RDG03/2016 which is
gratefully acknowledged.
10 W.M. Budzianowski

References

1. GCP (Global Carbon Project) (2011) Carbon budget and trends 2010. www.
globalcarbonproject.org/carbonbudget
2. Budzianowski WM (2013) Modelling of CO2 content in the atmosphere until 2300: Influence
of energy intensity of gross domestic product and carbon intensity of energy. Int J Glob
Warming 5(1):1–17
3. IEA (International Energy Agency) (2011) Key World Energy Statistics, Paris
4. Budzianowski WM (2016) Explorative analysis of advanced solvent processes for energy
efficient carbon dioxide capture by gas-liquid absorption. Int J Greenhouse Gas Control
49:108-120. http://dx.doi.org/10.1016/j.ijggc.2016.02.028
5. Budzianowski WM (2015) Single solvents, solvent blends, and advanced solvent systems in
CO2 capture by absorption: a review. Int J Glob Warming 7(2):184–225
6. Shanableh A, Merabtene T, Omar M, Imteaz M (2011) Impact of surface ocean acidification
on the CO2 absorption rate. Int J Glob Warming 3(1–2):163–172
7. Li L, Fu ML, Zhao YH, Zhu YT (2012) Characterization of carbonic anhydrase II from
Chlorella vulgaris in bio-CO2 capture. Environ Sci Pollut Res 19(9):4227–4232
8. Pendersen-van Elk NJMC, Derks PWJ, Fradette S, Veersteg GF (2012) Kinetics of absorption
of carbon dioxide in aqueous MDEA solutions with carbonic anhydrase at 298 K. Int J
Greenhouse Gas Control 9:385–392
9. Osborne CP, Beerling DJ (2006) Nature’s green revolution: The remarkable evolutionary rise
of C4 plants. Philos Trans of the Royal Society B: Biol Sci 361(1465):173–194
10. Steger U, Achterberg W, Blok K, Bode H, Frenz W, Gather C, Hanekamp G, Imboden D,
Jahnke M, Kost M, Kurz R, Nutzinger HG, Ziesemer T (2005) Sustainable development and
innovation in the energy sector. Springer, Berlin
11. IEA (International Energy Agency) (2014) Key World Energy Statistics, Paris
12. Akermin Inc. (2013) Final Scientific/Technical Report−Advanced Low Energy Enzyme−
Catalyzed Solvent for CO2 Capture. http://www.netl.doe.gov/File%20Library/Research/Coal/
carbon%20capture/post-combustion/FE0004228-Final-Report-01-07-2014.pdf
13. Boghossian AA, Ham MH, Choi JH, Strano MS (2011) Biomimetic strategies for solar energy
conversion: a technical perspective. Energy Environ Sci 4(10):3834–3843
14. Kumar P, Kim K-H (2016) Recent progress and innovation in carbon capture and storage
using bioinspired materials. Appl Energy 172:383–397
15. Leung DYC, Caramanna G, Maroto-Valer MM (2014) An overview of current status of
carbon dioxide capture and storage technologies. Renew Sustain Energy Rev 39:426–443
16. Gaus W, Hochschwender K, Schunck W (1927) Process for extracting carbon dioxide from
gaseous mixtures and forming alkaline carbonates. U.S. Patent 1897725
17. Greenwood K, Pearce M (1953) The removal of carbon dioxide from atmospheric air by
scrubbing with caustic soda in packed towers. Trans Inst Chem Eng 31:201–207
18. IEA GHG. A brief history of CCS and current status. http://www.ieaghg.org/docs/General_
Docs/Publications/Information_Sheets_for_CCS_2.pdf. 2015-12
19. Budzianowski WM (2017) Implementing carbon capture, utilisation and storage in the
circular economy. Int J Glob Warming
20. Budzianowski WM (2011) CO2 reactive absorption from flue gases into aqueous ammonia
solutions: the NH3 slippage effect. Environ Prot Eng 37(4):5–19
21. Yu YS, Wang GX, Lu HF, Zhang ZX, Rudolph V (2013) Characterizing the transport
properties of multi-amine solutions for CO2 capture by molecular dynamics simulation.
J Chem Eng Data 58(6):1429–1439
22. Dash SK, Samanta AN, Bandyopadhyay SS (2014) Simulation and parametric study of post
combustion CO2 capture process using (AMP + PZ) blended solvent. Int J Greenhouse Gas
Control 21:130–139
Introduction to Carbon Dioxide Capture by Gas–Liquid Absorption … 11

23. Li H, Le Moullec Y, Lu J, Chen J, Valle Marcos JC, Chen G (2014) Solubility and energy
analysis for CO2 absorption in piperazine derivatives and their mixtures. Int J Greenhouse Gas
Control 31:25–32
24. Smith KH, Lee A, Mumford K, Li S, Indrawan I, Thanumurthy N, Temple N, Anderson C,
Hooper B, Kentish S, Stevens G (2015) Pilot plant results for a precipitating potassium
carbonate solvent absorption process promoted with glycine for enhanced CO2 capture. Fuel
Process Technol 135:60–65
25. Cowan RM, Jensen MD, Pei P, Steadman EN, Harju JA (2011) Current status of CO2 capture
technology development and application. National Energy Technology Laboratory U.S.
Department of Energy
26. IRENA (International Renewable Energy Agency) (2014) REmap 2030: a renewable energy
roadmap. IRENA, Abu Dhabi
27. Sage R, Russell M (1999) C4 Plant Biology
28. Budzianowski WM (2012) Benefits of biogas upgrading to biomethane by high-pressure
reactive solvent scrubbing. Biofuels Bioprod Biorefin 6(1):12–20
29. Budzianowski WM (2012) Negative carbon intensity of renewable energy technologies
involving biomass or carbon dioxide as inputs. Renew Sustain Energy Rev 16(9):6507–6521
30. Budzianowski WM (2012) Value-added carbon management technologies for low CO2
intensive carbon-based energy vectors. Energy 41(1):280–297
31. Budzianowski WM, Postawa K (2017) Renewable energy from biogas with reduced carbon
dioxide footprint: implications of applying different plant configurations and operating
pressures. Renew Sustain Energy Rev. Doi:10.1016/j.rser.2016.05.076
Assessment of Thermodynamic Efficiency
of Carbon Dioxide Separation in Capture
Plants by Using Gas–Liquid Absorption

Wojciech M. Budzianowski

Abstract Typical carbon capture plants include CO2 separation and compression
steps. CO2 separation from diluted flue gases may be achieved by using gas-liquid
absorption. This process requires work input for e.g. separating CO2 from flue gases
and regenerating the CO2 loaded solvent. Hence, CO2 capture plants involving
gas-liquid absorption consume a remarkable part of power and thermal energy
generated by power plants. By increasing the thermodynamic efficiency of capture
plants one can increase the produced power and save fossil fuels. Therefore, this
study provides a quantitative assessment of the thermodynamic efficiency of CO2
separation in capture plants. To this aim the minimum work required for CO2
separation and actual work input in realistic carbon capture plants are estimated.
The results reveal that for the state-of-the-art MEA solvent the thermodynamic
efficiency of the capture plant is about 16%, for state-of-the-art advanced solvent
based capture process (ASBCP) is about 25%, while given the progress in devel-
oping ASBCPs in near future it may reach about 30%. Additional measures to
reduce the energy requirement of the capture plant such as heat pumps are also
discussed. This all means that CO2 separation by gas-liquid absorption is still a
relatively inefficient process and remarkable potential for further improvements
with step change innovations in gas-liquid absorption exist and may be beneficially
used for optimising CO2 capture plants.

Nomenclature
ASBCP Advanced solvent based capture process
E Energy, J
G Gibbs free energy, J
MEA Monoethanolamine
n Molar flow rate, kmol/s

W.M. Budzianowski
Consulting Services, Wrocław, Poland
W.M. Budzianowski (&)
Renewable Energy and Sustainable Development (RESD) Group, Wrocław, Poland
e-mail: wojciech.budzianowski@gmail.com

© Springer International Publishing AG 2017 13


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_2
14 W.M. Budzianowski

NGCC Natural gas combined cycle


pi Partial pressure of the i-th gas, Pa
p Total pressure, Pa
PC Pulverised coal
PCC Postcombustion capture
Q Heat, J
R Ideal gas constant = 8.314 J/(mol K)
T Absolute temperature, K
TREF Reference temperature = 293.15 K
TSOURCE Source temperature = 393.15 K
W Work, J
Wact Actual work, J
Wmin Minimum work of separation, J
yCO
i
2 Mole fraction of CO2 in the gas mixture i, –
iCO2
yi Mole fraction of non-CO2 remainder, −
η Thermodynamic efficiency, –
ηturbine Turbine efficiency = 90%

Subscripts and superscripts


A Stream A
B Stream B
C Stream C
i Index
sep Separation
TOTAL Total

1 Introduction

Energy efficient decarbonisation of the energy sector and major industries is


essential for achieving sustainable development. Recent research efforts in CO2
capture led to the development of many new processes that may facilitate reduction
of CO2 emissions. However, technological readiness levels (TRLs) of these pro-
cesses are different and they use different necessary energetic inputs (e.g. waste
heat, extracted stream, compressor work). It is thus difficult to compare the per-
formance of such diverse processes [1]. Mature CO2 capture technologies may have
slightly improved energy efficiencies compared to emerging technologies but at the
same time they may offer much less potential for further improvements. Due to
technology learning the future energy efficiency performance of various techno-
logical options currently at different stages of development is often unclear [2].
Assessment of Thermodynamic Efficiency of Carbon Dioxide … 15

CO2 capture by gas-liquid absorption undergoes rapid development and several


new processes aimed at reducing its energy requirements have recently been pro-
posed. Examples include solvent blends and advanced solvent based capture pro-
cesses applying facilitated CO2 capture by employing phase changing solvents,
catalysed absorption or microencapsulated solvents [3, 4]. These new techniques
should be explored in order to assess their potential for getting closer to the ther-
modynamic limits of CO2 capture.
Two laws of thermodynamics are essential tools for assessing the energy effi-
ciency of CO2 capture plants. The first law reflects the principle of energy con-
servation and the second law accounts for process irreversibility. The second law
efficiency shows how far a given capture process is from its thermodynamic limit.
The methodology employed in this study uses the minimum work required for CO2
separation under idealised operating conditions which is calculated from the
combined first and second law of thermodynamics. Further, actual work of CO2
separation is estimated through literature surveys and confirmed by calculations.
The provided assessment of thermodynamic efficiencies offer some qualitative
insights into the energy efficiency of CO2 capture by gas-liquid absorption and may
quantify potentials for future improvements. The study begins with calculation of
the minimum work of CO2 separation (Sect. 2) which is followed by the deter-
mination of actual work requirements from actual capture plant evidence and by
calculation (Sect. 3). The thermodynamic efficiency of CO2 separation in capture
plants involving gas-liquid absorption is calculated in Sect. 4. Section 5 discusses
additional related measures capable of reducing the work requirements of CO2
separation. Finally, Sect. 6 summarises conclusions drawn from the study.

2 Minimum Thermodynamic Work of CO2 Separation

The minimum work required for CO2 separation under idealised operating condi-
tions is calculated from the combined first and second law of thermodynamics. It
uses the flow rates and compositions of inlet and outlet streams and operating
temperature. Figure 1 shows a CO2 capture plant along with a CO2 emitting plant
and corresponding gas streams. Stream A represents a flue gas comprising CO2
while stream B is a CO2 rich stream, and stream C is the remainder of flue gases.
The capture plant operates with a certain fixed capture rate and the purity level of
CO2 rich stream is imposed.

Fig. 1 Schematic of the CO2 capture plant involving gas-liquid absorption


16 W.M. Budzianowski

The minimum work required for separating CO2 from a flue gas mixture for an
isothermal and isobaric process equals to the negative of the difference in Gibbs free
energy of the separated final states (streams B and C in Fig. 1) from the mixed
initial state (stream A in Fig. 1). This is the negative of Gibbs free energy of
mixing. For an ideal gas the Gibbs free energy change between stream A to streams
B and C is:

Wmin ¼ DGsep ¼ DGB þ DGC  DGA ð1Þ

For an ideal mixture, the partial molar Gibbs free energy for each gas is [5, 6]:
 
@G pi
¼ Gi þ RTln
0
ð2Þ
@ni p

Therefore, the total Gibbs free energy of an ideal gas mixture can be expressed as:
X @G
GTOTAL ¼ ni ð3Þ
i
@ni

The minimum work required to shift from state A to states B and C is associated
with the free energy difference between the product and reactant states, which can
be calculated by inserting Eqs. (2) to (3) resulting in:
 2  CO2   
GA ¼ nCO
A GCO2 þ nA
2 0
GACO2 þ RT nCO
ACO2 0
A ln yA þ nACO
A
2
ln yACO
A
2

ð4AÞ
 2  CO2   
GB ¼ nCO
B GCO2 þ nB
2 0
GBCO2 þ RT nCO
BCO2 0
B ln yB þ nBCO
B
2
ln yBCO
B
2

ð4BÞ
 2  CO2   
GC ¼ nCO
C GCO2 þ nC
2 0
GCCO2 þ RT nCO
CCO2 0
C ln yC þ nCCO
C
2
ln yCCO
C
2

ð4CÞ

This ideal mixing takes place at constant temperature and pressure. By substi-
tuting Eqs. (4A–C) to Eq. (1) the minimum work of separation is obtained:
2  2  CO2      3
nB yCO
B ln yB þ 1  yCO
B
2
ln 1  yCO
B
2

6  CO2  CO2      7
Wmin ¼ RT 6
4 þ nC yC ln yC þ 1  yC ln 1  yC
CO2 CO2 7
5 ð5Þ
 CO2  CO2     
 nA yA ln yA þ 1  yACO2
ln 1  yACO2

Equation (5) is subsequently used to calculate the minimum work of separation.


Figure 2 presents the minimum work Wmin for CO2 separation as a function of the
molar concentration of CO2 in the gas mixture (stream A in Fig. 1). Varied
Assessment of Thermodynamic Efficiency of Carbon Dioxide … 17

Fig. 2 Impact of
temperature, CO2 capture rate
and CO2 stream purity on the
minimum thermodynamic
work of CO2 separation for
gases with different CO2
contents. Parameters varied:
T—(300–350 K), capture
rate—(50–90%), CO2 stream
purity—(90–98%)

parameters include T (300–350 K), capture rate (50–90%), and CO2 stream purity
(90–98%). As it can be observed the minimum work required for CO2 separation
increases with decreasing CO2 concentration. In contrast, Wmin increases with
increasing temperature, capture rate and CO2 stream purity.
More specifically, the minimum work per kg of CO2 captured in fixed tem-
perature of 318 K, capture rate of 90% and CO2 stream purity of 98% at inlet gas
comprising 10% CO2 is 174 kJ/kgCO2. When CO2 is separated from air (inlet gas
contains 0.04% CO2) the minimum work is 509 kJ/kgCO2, which is three times
greater. When CO2 is separated from biogas [7–9] combustion flue gases (inlet gas
has 50% CO2) Wmin is 54 kJ/kgCO2, which is three times smaller. Effect of tem-
perature is less pronounced: at 300 K Wmin is 164 kJ/kgCO2 (6% less) and at 350 K
it is 192 kJ/kgCO2 (10% more). When capture rate is decreased to 50% Wmin is
149 kJ/kgCO2 (14% less). When CO2 purity is set at 90% Wmin is 159 kJ/kgCO2
(11% less).

3 Actual Work Requirement of CO2 Separation


in Realistic Carbon Capture Plants

The minimum work of separation reflects a theoretical idealised isothermal and


isobaric CO2 capture process that cannot be achieved in the practice. In an actual
CO2 capture plant employing gas-liquid absorption significantly more work will be
18 W.M. Budzianowski

required. For example, pumps are required to deliver the solvent to the scrubber and
stripper which takes work. Similarly electrical blowers drive a flue gas through the
scrubber. The solvent regeneration requires thermal energy to obtain the CO2 rich
and regenerated solvent streams. Each of these processes have associated effi-
ciencies based upon irreversibilities, such as friction, heat transfer, gas expansion,
gas mixing, etc. Therefore, actual work requirement depends on the unit operations
of the capture process and their individual efficiencies.
Consequently, in addition to the minimum work which is insufficient to operate a
CO2 capture plant some extra work is needed. This lost work or lost exergy reflects
inefficiencies of a given process realisation and cannot be avoided in the practice.
The actual work of CO2 separation can be estimated by using evidence from
realistic carbon capture plants or by calculation of a model carbon capture plant.
These two methods are employed below.

3.1 Evidence from Realistic Carbon Capture Plants

The actual work requirement of CO2 capture plants involving gas-liquid absorption
operated under realistic conditions can be found in literature. Table 1 summarises
such evidence taken from literature relevant to state-of-the-art CO2 capture solvents
(especially MEA). It gives a breakdown of work requirement involving contribu-
tions from the flue gas blower, rich/lean solvent pumps, reboiler, and miscellaneous
minor components.
As seen from Table 1 work requirement of plant components varies across
various CO2 capture plants reported in literature. It is associated with differences in

Table 1 Literature overview of breakdowns of actual work requirements of realistic carbon


capture plants involving state-of-the-art gas-liquid absorption
Capture plant item Work requirement [kJ/kgCO2] References
Flue gas blower 50 [10]
70 [11]
Rich/lean solvent pumps 37 [12]
2 [13]
3.2 [14]
1.2 [15]
Work equivalent of reboiler duty 716–1110 [10]
(actual thermal energy) (2867–4443) [14]
497 (3956) [11]
491 (4063) [12]
964 (2892)
Miscellaneous (baseload—control valves, 25 [10]
minor units)
Assessment of Thermodynamic Efficiency of Carbon Dioxide … 19

Table 2 Representative actual work requirements of CO2 separation in capture plants involving
state-of-the-art gas-liquid absorption
Capture plant item Work requirement [kJ/kgCO2]
Flue gas blower 70
Rich/lean solvent pumps 10
Work equivalent of reboiler duty 900
Miscellaneous (baseload—control valves, minor units) 25
Total 1005

plant design, flue gas CO2 content, flue gas contaminants, capture rate, plant
capacity, employed solvent etc.
Table 2 summarises representative actual work requirements of CO2 separation
by state-of-the-art gas-liquid absorption.

3.2 Calculations for a Model Carbon Capture Plant

Figure 3 displays a schematic of a model CO2 capture plant that is used to calculate
work requirements and thus to improve the consistency of literature evidence
provided in Sect. 3.1. It illustrates that a model CO2 capture plant consists of two
major components: (i) CO2 separation and (ii) CO2 compression. In order to

Fig. 3 Schematic of a model CO2 capture plant involving state-of-the-art gas-liquid absorption
used in the calculation of actual work requirements of CO2 separation
20 W.M. Budzianowski

calculate work required for separating CO2 only the former component needs to be
accounted for.
The methodology used in current calculations includes blowing work, pumping
work, reboiler equivalent work and miscellaneous contributions. It has recently
been applied for evaluating power requirements of CO2 separation from biogas by
using gas-liquid absorption involving pressurised water scrubbing [16]. Here, heat
released in the condenser is neglected and integration measures are not accounted
for, except the cross heat exchanger.

3.2.1 Blowing Work

The work requirement for blowing flue gas is calculated from the gas flow rate (qG)
and total pressure rise (Dp). The shaft work requirement (WB) is obtained by using
the mechanical efficiency of the blower (ηB).

WB ¼ qG Dp=gB ð6Þ

The mechanical efficiency values of 0.78 and 0.75 are used for the blower and
pump, respectively [17].

3.2.2 Pumping Work

The work requirements for pumping the rich and lean solvents as well as cooling
water are calculated from the water density (qL), gravitational acceleration (g) and
liquid flow rate (qL). In order to obtain shaft power (WP), the mechanical efficiency
of a pump (ηP) is also accounted for.

Wp ¼ qL gqL HT =gp ð7Þ

The total pumping work includes contributions from pumping the rich and lean
solvents as well as cooling water (WP-RS, WP-LS, WP-COOL).
The total pressure head (HT) is needed to calculate work requirement for
pumping and blowing. It is calculated as the sum of the pressure difference
(HSCR − HATM), the static (HS) and dynamic (HD) heads as follows:

HT ¼ ðHSCR  HATM Þ þ HS þ HD ð8Þ

The pressure difference HSCR − HATM is obtained by subtracting the pressure in


the scrubber and the atmospheric pressure.
The static head is taken equal to the height of columns (scrubber/stripper). The
dynamic head is calculated from the Darcy-Weisbach equation assuming optimum
gas velocity 20 m/s and liquid velocity 2.5 m/s:
Assessment of Thermodynamic Efficiency of Carbon Dioxide … 21

L w2
HD ¼ f ð9Þ
D 2g

Friction factor f is obtained explicitly from a relationship approximating the


Colebrook-White equation [18].

6:4
f ¼   pffiffiffi2:4 ð10Þ
lnðReÞ  ln 1 þ 0:01Re De 1 þ 10 De

CO2 scrubbing is more efficient at lower temperatures characterised by increased


CO2 solubility. Therefore, lean solvent is cooled in a heat exchanger using water as
a coolant. Cooling is designed to reduce lean solvent temperature to 5 K above
ambient temperature. The flow rate of cooling water (qCOOL) required to cool the
lean solvent is calculated from energy balance and subsequently inserted into
Eq. (7) to calculate pumping work associated with lean solvent cooling.

qC qC cPC ðTCout  TCin Þ


qCOOL ¼ ð11Þ
qL cPL ðTLout  TLin Þ

3.2.3 Reboiler Equivalent Work

Qsens is estimated as heat that needs to be added to heat the solvent from temper-
ature TINLET to TSTRIPPER:

Qsens ¼ mS Cp ðTSTRIPPER  TINLET Þ ð12Þ

Qvap ¼ mS Hvap ð13Þ

Qact ¼ Qsens þ Qvap ð14Þ

Thermal energy is subsequently converted to equivalent work by using a con-


version factor obtained from Carnot efficiency [19, 20]. In addition, turbine effi-
ciency is used to convert the work into electricity.
 
TSOURCE  TREF
Wequivalent ¼ Qact gturbine ð15Þ
TSOURCE

3.2.4 Miscellaneous Contributions to Total Work

Miscellaneous contributions to total work include control valves and other minor
units associated with CO2 separation.
22 W.M. Budzianowski

Table 3 Calculated work requirements of the model CO2 capture plant involving CO2 separation
by state-of-the-art gas-liquid absorption
Capture plant item Work requirement [kJ/kgCO2]
Flue gas blower 40
Rich/lean solvent pumps 20
Work equivalent of reboiler duty 890
Miscellaneous (baseload—control valves, minor units) 25
Total 975

3.2.5 Total Work Requirement of the Model CO2 Capture Plant

The calculations takes into account CO2 capture plant technical parameters from
[21]. The results are presented in Table 3. As seen from comparison of Tables 2
and 3 total work requirement and breakdown are similar. Therefore these calculated
actual work requirements will be used in the next Section.

4 Thermodynamic Efficiency of CO2 Separation

Based on the minimum work of separation and actual work requirements the
thermodynamic efficiency may be calculated as the ratio of minimum and actual
work.

Wmin
g¼ ð16Þ
Wact

Wmin is the useful effect of separating flue gases and it is equal to the Gibbs free
energy of CO2 separation. Wact is the work input required to run the realistic CO2
capture plant. Wact includes work consumed by pumps, blowers and compressors as
well as work equivalent of thermal energy consumed by the stripper. This ther-
modynamic efficiency uses only work or work equivalent inputs/outputs and is
therefore consistent with the second law of thermodynamics.
Thermodynamic efficiencies of separation processes have been investigated for a
variety of pollutant reduction techniques in a wide range of concentrations. In
Fig. 4 the second law efficiency is plotted as a function of pollutants concentration
(at separation rate 90%, PC power plant 500 MWel). Data are adopted from various
real separation processes such as CO2 capture in coal PCC plants (by amine
scrubbing) and NGCC plants (by amine scrubbing), as well as for separations of
NOX (by selective catalytic reduction), SOX (by wet flue gas desulfurization), and
Hg (by activated carbon injection). As it can be observed the second law efficiency
increases with increasing pollutant concentration [22]. In view of these results the
specific nature of CO2 is such that it has very high concentration compared to other
Assessment of Thermodynamic Efficiency of Carbon Dioxide … 23

Fig. 4 Influence of pollutant


concentration in flue gases on
the second law efficiency of a
flue gas cleaning process, data
adapted from [23]

typical pollutants contained in flue gases, e.g. SOX, NOX or Hg. It means that CO2
separation may require different approaches than other highly diluted pollutants.
Due to higher concentrations and flow rates the thermodynamic efficiency is more
pronounced for separating CO2 than other minor contaminants.
The CO2 capture plant separates relatively concentrated gases and therefore the
second low efficiency of CO2 separation is higher than that of diluted pollutants,
typically it is higher than 10%.
The minimum work of separation (for 90% CO2 capture rate, 98% CO2 stream
purity, temperature 318 K, and 13% CO2 content in flue gases) amounts to
158 kJ/kgCO2. Taking into account the actual work requirement of MEA based
CO2 capture of about 975 kJ/kgCO2 the second law efficiency calculated from
Eq. 16 is 16.2%. The iCAP project obtained a two immiscible liquid phases based
ASBCP with the experimental energy requirement of 2400 kJ/kgCO2 for solvent
regeneration. It translates to total work equivalent of about 630 kJ/kgCO2 and
yields the second law efficiency of 25.1%. Given this progress and several other
ongoing projects the second law efficiency of about 30% (527 kJ/kgCO2) may be
achieved by CO2 capture systems employing innovative ASBCPs in the future.
Figure 5 compares these calculations and expectations.

Fig. 5 Potential for the


reduction of work
requirement of CO2 capture
by ASBCPs with technology
development. Parameters:
CO2 capture rate—90%, CO2
stream purity—98%,
temperature—318 K, CO2
content in flue gases—13%
24 W.M. Budzianowski

It is worth to mention that a typical PC power plant having a CO2 emission


intensity of 1000 kgCO2/MWhel (0.000278 kgCO2/kJel) produces 3600 kJel/kgCO2
emitted. Consequently, the state-of-the art MEA CO2 capture process that consumes
975 kJ/kgCO2 would use 27% of electricity delivered by such a power plant.
It emphasises that the progress in reducing work requirement of CO2 separation
is essential.

5 Examples of Additional Measures Capable of Reducing


Work Requirement

The actual work can be decreased by optimising one or all of the separation steps.
The optimisation may rely on tuning existing CO2 capture processes or applying
new ones bringing a step change in technology [e.g. advanced solvent based capture
processes (ASBCPs)].
As shown in breakdown given in Table 3 the most significant contribution is
associated with reboiler duty. Therefore, techniques that deliver thermal energy to
the reboiler may greatly improve the performance of CO2 capture plants. For
example, heat pumps are able deliver thermal energy taken from the environment or
waste heat sources available in power plants [24]. Namely, using an
absorption-driven heat pump for waste heat recovery the energy requirement of a
conventional solvent based CO2 separation may be potentially reduced to
1670 kJ/kgCO2 [25]. If a heat pump uses renewable energy taken from the envi-
ronment the CO2 capture plant may overcome thermodynamic limits of closed
systems. The application of heat pumps may rely on cooling the scrubber which
enhances absorption and heating the stripper which enhances solvent regeneration.
The scrubber is fed with flue gases having relatively high temperature which
ensures that this type of heat source is stable in capture plants.
Another measure is associated with improved allocation of thermal energy
within capture plants. It may also be achieved by a distributed cross heat exchanger.
This device functions by delivering more of the heat to the bottom section of the
stripper than to the top section. The stripper with lower temperature at the top has a
lower solvent vapour fraction, which reduces useful heat lost to the condenser
[19, 25].
Further, process intensification techniques for dedicated for post-combustion
CO2 capture [26], if could be suitably integrated with ASBCPs can add additional
energy efficiency benefits to the whole capture plant.
Assessment of Thermodynamic Efficiency of Carbon Dioxide … 25

6 Conclusions

This study analyses thermodynamic efficiency of CO2 separation by gas-liquid


absorption. The minimum work required for CO2 separation under idealised
operating conditions is provided. Insights in blowing work, pumping work, reboiler
equivalent work and miscellaneous contributions are given based on realistic
capture plants found in literature and calculations made for the model CO2 capture
plant. Taking into account actual work required to separate CO2 the second law
efficiency is calculated. For the MEA solvent it is about 16%, for the state-of-the-art
ASBCP about 25%, and for future ASBCPs it may achieve about 30% or more.
The second law efficiency is limited to closed systems. However, when a CO2
capture plant operates as an open system the thermodynamic limits may be over-
come. This can be implemented by using renewable energy, e.g. through heat
pumps utilisation. Heat pumps may also facilitate the distribution of thermal energy
within the system thus leading to limited irreversibilities and exergy losses and
consequently to improved thermodynamic efficiency. Heat management is partic-
ularly relevant since it is clearly the major contributor to total work requirement of
CO2 separation, at least when thermal regeneration of solvent is employed. Due to
the scale of CO2 capture plants all feasible measures need to be applied simulta-
neously in order to achieve the meaningful reduction of work requirement of CO2
separation.

Acknowledgments This study has been supported by the members of the Renewable Energy and
Sustainable Development (RESD) Group (Poland) under the project RESD-RDG03/2016 which is
gratefully acknowledged.

References

1. Calbry-Muzyka S, Edwards CF (2014) Thermodynamic benchmarking of CO2 capture


systems: exergy analysis methodology for adsorption processes. Energy Procedia 63:1–17
2. Rochedo PRR, Szklo A (2013) Designing learning curves for carbon capture based on
chemical absorption according to the minimum work of separation. Appl Energy 108:383–391
3. Budzianowski WM (2016) Explorative analysis of advanced solvent processes for energy
efficient carbon dioxide capture by gas-liquid absorption. Int J Greenhouse Gas Control
49:108–120
4. Budzianowski WM (2015) Single solvents, solvent blends, and advanced solvent systems in
CO2 capture by absorption: a review. Int J Glob Warming 7(2):184–225
5. Gaskell D. (1995) Introduction to the thermodynamics of materials. Taylor & Francis.
Washington D.C
6. House KZ, Harvey CF, Aziz MJ, Schrag DP (2009) The energy penalty of post-combustion
CO2 capture & storage and its implications for retrofitting the U.S. installed base. Energy
Environ Sci 2:193–205
7. Budzianowski WM (2012) Negative carbon intensity of renewable energy technologies
involving biomass or carbon dioxide as inputs. Renew Sustain Energy Rev 16(9):6507–6521
8. Budzianowski WM (2012) Value-added carbon management technologies for low CO2
intensive carbon-based energy vectors. Energy 41(1):280–297
26 W.M. Budzianowski

9. Budzianowski WM (2012) Benefits of biogas upgrading to biomethane by high-pressure


reactive solvent scrubbing. Biofuels, Bioprod Biorefin 6(1):12–20
10. Svendsen HF, Hessen ET, Mejdell T (2011) Carbon dioxide capture by absorption, challenges
and possibilities. Chem Eng J 171(3):718–724
11. Zhang G, Yang Y, Xu G, Zhang K, Zhang D (2015) CO2 capture by chemical absorption in
coal-fired power plants: energy-saving mechanism, proposed methods, and performance
analysis. Int J Greenhouse Gas Control 39:449–462
12. Yu J, Wang S (2015) Modeling analysis of energy requirement in aqueous ammonia based
CO2 capture process. Int J Greenhouse Gas Control 43:33–45
13. Geuzebroek FH, Schneiders LHJM, Kraaijveld GJC, Feron PHM (2004) Exergy analysis of
alkanolamine-based CO2 removal unit with AspenPlus. Energy 29(9–10):1241–1248
14. Skorek-Osikowska A, Kotowicz J, Janusz-Szymańska K (2012) Comparison of the energy
intensity of the selected CO2-capture methods applied in the ultra-supercritical coal power
plants. Energy Fuels 26(11):6509–6517
15. Arias AM, Mores PL, Scenna NJ, Mussati SF (2016) Optimal design and sensitivity analysis
of post-combustion CO2 capture process by chemical absorption with amines. J Clean Prod
115:315–331
16. Budzianowski WM, Wylock CE, Marciniak PA (2017) Power requirements of biogas
upgrading by water scrubbing and biomethane compression: comparative analysis of various
plant configurations. Energy Convers Manag. doi:10.1016/j.enconman.2016.03.018
17. Razi N, Svendsen HF, Bolland O (2013) Cost and energy sensitivity analysis of absorber
design in CO2 capture with MEA. Int J Greenhouse Gas Control 19:331–339
18. Avci A, Karagoz I (2009) A novel explicit equation for friction factor in smooth and rough
pipes. ASME J Fluids Eng 131:061203
19. Lin Y, Rochelle GT (2016) Approaching a reversible stripping process for CO2 capture.
Chem Eng J 283:1033–1043
20. Kim H, Lee KS (2016) Design guidance for an energy-thrift absorption process for carbon
capture: analysis of thermal energy consumption for a conventional process configuration.
Int J Greenhouse Gas Control 47:291–302
21. Alhajaj A, Mac Dowell N, Shah N (2016) A techno-economic analysis of post-combustion
CO2 capture and compression applied to a combined cycle gas turbine: Part I. A parametric
study of the key technical performance indicators. Int J Greenhouse Gas Control 44:26–41
22. House KZ, Baclig AC, Ranjan M, van Nierop EA, Wilcox J, Herzog HJ (2011) Economic and
energetic analysis of capturing CO2 from ambient air. Proc Natl Acad Sci USA 108
(51):20428–20433
23. Wilcox J. Carbon capture. 2012. Springer, New York
24. Wołowicz M, Milewski J, Futyma K, Bujalski W (2014) Boosting the efficiency of an
800 MW-class power plant through utilization of low temperature heat of flue gases. Appl
Mech Mater 483:315–321
25. Higgins SJ, Liu YA (2015) CO2 capture modeling, energy savings, and heat pump
integration. Ind Eng Chem Res 54(9):2526–2553
26. Wang M, Joel AS, Ramshaw C, Eimer D, Musa NM (2015) Process intensification for
post-combustion CO2 capture with chemical absorption: a critical review. Appl Energy
158:275–291
Process Implications of CO2 Capture
Solvent Selection

Leigh T. Wardhaugh and Ashleigh Cousins

Abstract In the development of energy efficient solvents for advanced CO2 cap-
ture processes, it has often been the case that potentially excellent solvents are
overlooked or set aside because of perceived process difficulties or extra costs due
to the physical, chemical or thermodynamic properties of the solvent. This chapter
considers the process implications of solvent selection with the objective of
selecting and designing the process to suit the solvent rather than forcing the solvent
into an existing process. The chapter discusses the design, modelling and costing of
conventional amine processes insofar as advanced energy efficient solvents still rely
on this process. Individual solvent properties, including reaction kinetics, thermo-
dynamics and physical properties are outlined and their impact on design are dis-
cussed. Aspects such as degradation products, corrosivity and environmental
considerations will impact the selection of process equipment and materials of
construction. The impact of these properties of energy efficient solvents on indi-
vidual unit operations are also discussed in some detail.

Nomenclature and Acronyms


Acronyms
FDG Flue gas desulphurization
HPLC High pressure liquid chromatography
HSS Heat stable salts
HTU Height of a transfer unit
IC Ion chromatography
LP Low pressure
MDEA Methyldiethanolamine
MEA Monoethanolamine

L.T. Wardhaugh (&)


CSIRO Energy, Mayfield West, NSW 2304, Australia
e-mail: leigh.wardhaugh@csiro.au
A. Cousins
CSIRO Energy, Pullenvale, QLD 4069, Australia

© Springer International Publishing AG 2017 27


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_3
28 L.T. Wardhaugh and A. Cousins

MNPZ Mono-nitrosopiperazine
MP Medium pressure
PCC Post-combustion capture
PZ Piperazine
SCR Selective catalytic reduction
VLE Vapour-liquid-equilibrium

Symbols
Cp Heat capacity (constant pressure)
E Efficiency; capture efficiency; Energy
G Mass flowrate of gas
H Height (of packing)
vap
DHCO 2
Desorption enthalpy for CO2
I Interest
L Mass flowrate of liquid
m Mass flowrate
MW Molecular weight
P Pressure, Plant capacity
Q Heat, Energy
T Temperature
v Volumetric fraction of a component in a gas
w Weight fraction
a CO2 loading (moles CO2/moles active solvent component)
D Difference

Subscripts
0 Unloaded solvent
am Amine or other active component of the capture solvent
cap Capital (cost)
CO2 Value for CO2 or its fraction of a stream
cond Condenser, Condensate
cw Cooling water
DES Desorber (stripper)
FG Flue Gas
in Inlet flow
L Lean (stripped) solvent
out Outlet flow
R Rich (containing captured CO2) solvent
sens Sensible heat
tot Total
Process Implications of CO2 Capture Solvent Selection 29

1 Introduction

Processes to capture CO2 from combustion flue gas streams have developed from
processes used to remove impurities (principally H2S, CO2) from natural gas
streams that are usually at high pressure. Physical absorbents such as methanol or
propylene carbonate, used in pressure swing processes are economically viable
when the partial pressure of CO2 (or the target impurity) is above approximately
500 kPaabs [1], whereas the CO2 partial pressure in flue gas streams is in the range
5–20 kPaabs. The development of thermal swing absorption processes to capture
CO2 using alkanolamines, first patented in 1930, made it economically feasible to
remove impurities from gas streams with low partial pressures of CO2. The
Gas/Spec FT-1 process developed by Dow Chemical and subsequently marketed by
Fluor Daniel as the Economine FG process used monoethanolamine
(MEA) together with proprietary additives and corrosion inhibitors to effectively
remove CO2 from low pressure oxygen containing streams in the absence of
reducing agents and with SO2 levels below 10 ppmv [2]. In the search for more
efficient CO2 capture solvents, the absorption/desorption process using 30 wt%
MEA has become the baseline solvent against which the new solvent developments
are often compared. MEA has the advantage of having a high reaction rate but the
disadvantages of high energy requirement (for the reverse reaction—desorption),
high corrosivity, toxicity and susceptibility to degradation through reaction with
other chemical species. Higher solvent concentrations would be preferred but are
limited by considerations of corrosivity, degradation rate and viscosity.
With the current world-wide focus on the need to reduce the emission of CO2
into the atmosphere, a wide range of absorption solvents (absorbents) have been
investigated [3] to achieve the goal of low cost capture of CO2 from flue gases,
hence the topic of this book. With the proposal of alternative absorbents, however,
what is often forgotten is the process design, operability and capital cost implica-
tions of the proposed absorbent choice. This is the topic of this chapter.
The following sections will put the choice of absorbent in the context of total
plant design and costing; consider the thermodynamic and physical properties that
impact design, operation and hence the total cost of the project, the design options
and how they are impacted by absorbent choice, and a final word on operability,
safety and other issues relating to absorbent choice. The overall message is that the
process must be designed for the solvent, rather than trying to force fit an advanced
solvent into a conventional process design. A new absorbent can only truly be
evaluated in a process designed specifically and optimized for the absorbent in
question.
30 L.T. Wardhaugh and A. Cousins

2 The Design and Costing of Post-Combustion Capture


Plants

2.1 Design Process Overview

The familiar layout of a typical Post-Combustion Capture (PCC) plant is shown in


Fig. 1. Figure 1 is unusual in that the size of the icon for each unit operation is
scaled to the capital cost of the item (costs based on Ramezan et al. [4], pro-rated on
power or size). A similar figure could be generated illustrating the energy cost of
each unit operation, with electrical energy pro-rated according to the power station
efficiency. Such figures could in principle be created for each solvent option to
provide a graphic illustration of the impact of solvent choice on cost.
The scale of the process is determined by the flue gas inlet mass flowrate (Gin)
and the required capture efficiency (ECO2: —usually set at 90%) while the flue gas
source determines the flue gas composition, flue gas molecular weight (MWFG) and
specifically the CO2 volume fraction ðvCO2 : Þ. MWCO2 is the molecular weight of
CO2.

ECO2  vCO2 :  Gin  MWCO2


CO2 captured ½kg=hr ¼ GCO2 ¼ ð1Þ
MW FG

Given a basic process using MEA with an (unloaded) alkanolamine concen-


tration (wam ) of 30 wt%; a cyclic loading [the difference between lean and rich

Fig. 1 Typical PCC plant using 30 wt% MEA. The size of the icon for each unit operation is
scaled to the capital cost of the item (costs based on Ramezan et al. [4], pro-rated on power or size;
adapted from Cottrell et al. [5])
Process Implications of CO2 Capture Solvent Selection 31

loadings (aR–aL)] of approximately 0.3 mol CO2/mol MEA, then for each tonne of
CO2 captured, up to 16.4 tonnes of solvent must be circulated through the absorber
and stripper setting the size of pumps, piping and heat exchangers.
In general for an alkanolamine of molecular weight MWam, the required
unloaded solvent circulation rate (L0) is given by:
 
ECO2  vCO2  Gin  MWam
Unloaded solvent rate ½kg=hr ¼ Lo ¼ ð2Þ
ð/R  /L Þ  MWFG  wam
 
wam  /L MWCO2
Lean solvent rate ½kg=hr ¼ LL ¼ Lo 1 þ ð3Þ
MWam
 
wam  /R MWCO2
Rich solvent rate ½kg=hr ¼ LR ¼ Lo 1 þ ð4Þ
MWam

The rich and lean loadings cannot be independently set. It is desirable to have a
high cyclic loading, but a high rich loading may be the consequence of an
overdesigned absorber (increased capital cost) while a low value of lean loading
may indicate a desorber column that is being run with too high a steam rate
(increased operating cost). The minimum energy requirement (Qtot/CO2 captured—
MJ/kg or GJ/t) is determined in a series of experiments (real or modelled) over a
range of values of Liquid to Gas mass flowrate ratios (L/G) with the heat input
adjusted to give a fixed value of Capture Efficiency ðECO2 Þ. It is presumed that the
values of L/G chosen are not too far from the contactor optimum to avoid biasing
the outcome of the experiment. This aspect is discussed in more detail in Sect. 4.2.
Due to the formation of bicarbonate, tertiary amines tend to have higher
absorption capacities than primary or secondary amines. CO2 solubility in an
absorbent is often determined via various vapour-liquid-equilibrium measurements
discussed in Sect. 3.2. These methods can be used to determine the equilibrium
CO2 loading of the absorbent under absorber and stripping column conditions. The
difference gives the loading capacity of the absorbent. Absorbents with a high
cyclic capacity are often preferred, as less absorbent is required for a given CO2
removal. However, a high cyclic capacity is of little benefit if the mass transfer of
CO2, and the rate of reaction is slow (taller absorber column required). Knowledge
of the gas and liquid flowrates allows the design of the contactor column diameters
to be determined (discussed further in Sect. 4.2) and a knowledge of the perfor-
mance of the solvent (particularly the rate of reaction—Sect. 3.1; and diffusivity—
Sect. 3.3) allows the column heights (or key parameters for other types of con-
tactors—Sect. 5) to be determined.
Inlet gas and liquid streams to CO2 absorber columns are typically maintained at
temperatures in the range 40–60 °C. This temperature range typically provides a
balance between the system thermodynamics, which favour lower temperatures,
and reaction kinetics, which favour higher temperatures [6]. Higher temperatures
will also improve viscosities and diffusion coefficients, enhancing overall mass
32 L.T. Wardhaugh and A. Cousins

transfer rates. CO2 absorption into aqueous amines is an exothermic process, which
leads to temperature increases in the column. This will lead to a temperature bulge
in the absorber column, the location of which will be affected by the operating
conditions of the plant (especially L/G). Peak temperatures of up to 70 °C have
been observed in PCC pilot plants operating on 30 wt% MEA [7].
The total heat requirement for the PCC process can be determined directly from
the steam consumption in the reboiler usually carried out by measuring the steam
inlet pressure and the exiting condensate flowrate. This total heat measurement
must be corrected for heat losses which in small scale equipment can be significant
and variable leading to significant errors. Alternatively, the total heat requirement
ðQtot Þ can be determined, using appropriate measurements from its components,
namely:

Qtot ðkWÞ ¼ Qsens þ QCO2 þ Qcond ð5Þ

The components of the total heat requirement are as follows:


• The sensible heat requirement (Qsens)

Qsens ðkWÞ ¼ LDESin  CpL  ðTDESout  TDESin Þ ð6Þ

where LDESin is the mass flow of absorption liquid entering the stripping column
(kg/s), CpL is the specific heat capacity of the absorption liquid (kJ/kgK), and
ðTDESout  TDESin Þ is the temperature difference between the hot lean absorption
liquid leaving the reboiler and the incoming (preheated) rich absorption liquid (K).
• The energy required to reverse the CO2 absorption reactions QCO2

vap
QCO2 ðkWÞ ¼ GCO2 DHCO 2
ð7Þ
vap
where GCO2 GCO2 is the mass flow rate of CO2 produced/captured (kg/s), and DHCO 2
is the desorption enthalpy for CO2 (kJ/kg).
• The energy in the excess steam leaving the stripping column (Qcond)—ap-
proximated by the condenser duty, (kW)

Qcond ¼ mcw :CPcw :DTcw ð8Þ

where mcw is the mass flow rate of cooling water to the stripping column condenser
(kg/s), CPcw is the specific heat of the cooling water (kJ/kgK), and DTcw is the
temperature difference of the cooling water into and out of the condenser (K).
Process Implications of CO2 Capture Solvent Selection 33

The value obtained from Eq. (8) may have to be corrected for any sub-cooling of
the condensate and, if significant, the effect of the pressure drop across the stripper.
The total heat requirement for the PCC process ðQtot Þ can then be determined from
the summation of the components as given in Eq. (5).
Typically a less energy intensive solvent is also less reactive requiring taller
absorber vessels. Therefore an energy cost is replaced with a capital cost. Much of
the research into alternate solvents seeks to break this nexus between reactivity and
desorption energy requirements [8].

2.2 Process Modelling for Different Solvent Choices

Considerable work has gone into the development of rigorous design procedures for
PCC plants which in turn can be used to develop complete cost studies (summa-
rized briefly in Sect. 2.3). The development of process models continues even for
the system considered the baseline (30 wt% MEA) with work done by Plaza et al.
[9], Zhang et al. [10], Freguia and Rochelle [11] and more recent contributions by
von Harbou et al. [12] and Razi et al. [13]. Thorough validation has also been made
against pilot scale results [14, 15].
Process models rely for their precision on 2 criteria:
• A thorough and precise representation of the thermodynamics, chemical reaction
kinetics and physical properties of the chemical species present and their various
interactions. This can be readily tested prior to the development of the process
model itself. For new solvents, their intermediate reaction products, products of
degradation etc. may not be adequately covered in existing property databases
so that appropriate property model parameters must be incorporated into the
program. This may require the close liaison with the process software developer.
Researchers such as Yu et al. [16] have also considered modelling the properties
of solvent blends.
• Correct allowance for operational matters such as foaming, back-mixing,
inadequate distribution of gas and liquid streams in the contactors; heat loss or
fouling in heat exchangers; multiphase flow in pipelines, valves and equipment
etc. Appropriately designed experiments, in which the interchange of informa-
tion between model and experiment can help elucidate each of these operational
issues and thus strengthen the applicability of the models.
Clearly from the above, process modelling has a key role to play in capture
solvent evaluation but the particular pathway and outcomes will be strongly
dependent on the choice of solvent.
34 L.T. Wardhaugh and A. Cousins

2.3 Overview of Process Costing Relating to Solvent Choice

The costing of a particular solvent selection is the most important step but also the
most contentious with a huge variation in the values quoted ($/t CO2 captured) for a
completed PCC plant project. This is mainly due to differing costing assumptions.
Efforts have been made to standardize costing assumptions and processes [17]. Key
cost parameters are the assumptions concerning energy cost (CostEnergy —$/GJ) and
interest on capital (Icap) taken as an average annual figure. As shown by Dave et al.
[18] the relative cost of capital and energy prevailing in different locations will
mean that the choice of solvent must be determined on a case by case basis with no
clear winner based on energy cost alone.
The purpose here is to illustrate the effect of solvent choice on design and cost
and for this a simplified cost structure can illustrate the change in total cost
(DCosttot—$/yr) of a solvent choice with some relatively simple calculations
(Eq. 9) for a plant of capacity P ðtCO2 =yrÞ. Of course a full design and detailed cost
analysis would be required to fully validate the choices made.

DCosttot ¼ DCostcap  Icap þ DEtot  P  CostEnergy ð9Þ



where the change in capital cost DCostcap can be estimated by pro-rating the
change in equipment size (e.g. column height) based on the change in the average
value of parameters such as mass and heat transfer coefficients, pressure drop, etc.
[19]; while the change in energy requirement ðDEtot Þ derives directly from the
experimental determination of the optimum L/G ratio and/or the changing
requirements of, for example, total pressure drop. The cost of the energy (steam and
electrical) is based on the locally prevailing opportunity cost of electricity and the
efficiency of the power plant in question.

3 Solvent Properties and Their Impact on Design

In this section a focus is given to the solvent properties that most effect process
selection, design and operation and discusses how the design may be modified to
overcome potential problems and take advantage of the unique properties of par-
ticular solvents.

3.1 Reaction Kinetics

The most important property of a CO2 capture solvent formulation is its ability to
reversibly react with CO2, the rate at which and extent to which this occurs is a
function of solvent composition, temperature and CO2 loading. The reaction is
Process Implications of CO2 Capture Solvent Selection 35

reversed by heating the CO2 loaded solution and by reducing the CO2 partial
pressure (e.g. by generating excess steam). Measurements of the kinetics (discussed
below) and phase equilibria (discussed in Sect. 3.2) dominate the literature con-
cerning solvent choices. The rate of reaction determines the required liquid resi-
dence time and hence the size of the contactor equipment. Experiments and/or
modelling determine the optimum conditions of temperature, pressure and
liquid/gas (L/G) mass flowrate ratios to complete the design process.
Slow reacting absorbents may require large contact areas and residence times to
achieve the target level of CO2 removal, leading to large and expensive columns.
Primary and secondary amines tend to react the fastest with CO2 as they form a
stable carbamate. Tertiary and hindered amines catalyse the hydrolysis of CO2 to
form a bicarbonate ion and protonated amine. The formation of bicarbonate is
slower compared to carbamate formation, thus CO2 removal via tertiary and hin-
dered amines tends to be slower than from primary or secondary amines [20].
CO2 mass transfer rates and reaction kinetics can be measured in the laboratory
via a number of experimental techniques:
• Wetted wall column—A wetted wall column is a simple gas/liquid contacting
device that is used to determine the mass transfer of CO2 from the gas phase to a
liquid absorbent under carefully controlled conditions. A thin liquid sheet of
known surface area is bought into counter-current contact with the gas phase.
Gas phase concentration measurements are used to determine the flux of CO2
taken up by the absorbent. Plots of the CO2 absorption flux against the applied
driving force (difference between actual and equilibrium partial pressure) are
then used to determine the overall mass transfer coefficient [21]. This is a useful
measurement for initial screening of absorbents, and also for determining fun-
damental data required for modelling purposes. Absorbents with a high CO2
mass transfer rate are typically preferred.
• Spray column—spray columns are sometimes used for absorbents with slow
reaction rates that require larger surface areas for meaningful mass transfer
determinations [22]. The method is similar to that used with the wetted wall
experiments, except the surface area is provided by liquid droplets.
• Stopped flow—Stopped flow spectrophotometers are rapid mixing devices used
to study the kinetics of fast reactions in solution. For PCC applications, these are
used to measure the reaction kinetics of CO2 with various absorbents. Syringes
are used to force solutions into a mixing chamber where the flow is ‘stopped’ by
an opposing piston. Measurements can be made on the level of a few mil-
liseconds [21].
36 L.T. Wardhaugh and A. Cousins

3.2 Phase Equilibria

The ultimate capacity of the absorbent for capturing CO2 is determined by the phase
equilibria relationships between the absorbent and CO2 as a function of tempera-
ture, solvent composition, and CO2 partial pressure. Due to the formation of
bicarbonate, tertiary amines tend to have higher absorption capacities than primary
or secondary amines. CO2 solubility in an absorbent is often determined via
vapour-liquid-equilibrium (VLE) measurements. There are various methods avail-
able for completing these measurements which are used to determine the distri-
bution of CO2 between the gas and liquid phases for absorbents under various
conditions. VLE of alkanolamine systems has been investigated by many
researchers including Bishnoi and Rochelle [23], Dang and Rochelle [24], Jou et al.
[25], Kumar et al. [26], Fan et al. [27], and Park et al. [28].
Most conventional capture solvents remain as a single phase liquid throughout
the absorption and desorption processes. Measurements of phase equilibria are
mostly concerned with the equilibrium content of CO2 that has reacted with the
absorbent forming a mix of dissolved CO2, carbamates, bicarbonates and carbon-
ates as a function of temperature, gas phase CO2 partial pressure and solvent
composition. An overall CO2 content of the liquid is sufficient to carry out pre-
liminary design, while advanced rate-based design techniques require a detailed
knowledge of the speciation which can be determined by Stopped Flow reactor
techniques [21] or NMR measurements [27].
The condition of perfect phase equilibria is seldom reached in any process
equipment and the terms “approach to equilibrium” or “stage efficiency” are used to
express this feature. A more rigorous approach is to consider the transfer of material
through successive layers at the gas-liquid interface and the reactions that occur in
those layers.
Blended amines often incorporate a fast reacting species (termed a “promotor”)
together with a higher capacity (but slow reacting) tertiary or hindered amine, thus
gaining advantage of each specie’s best properties. The modelling of the reaction
pathways in such systems can become quite complex [29, 30].
The vapour pressure of the absorbent species is often over-looked in early
screening. A highly volatile absorbent, such as ammonia, can suffer from significant
absorbent loss. This has implications not only on the environmental impact of the
process, but also the cost. Care needs to be taken when assessing the possible
vapour emissions of absorbent species. Early work with piperazine (PZ) suggested
it would have a higher volatility than the standard MEA absorbent due to its lower
boiling point (146 °C for PZ compared to 170 °C for MEA). However, CO2 loaded
piperazine solutions were in fact found to have a lower volatility than the standard
MEA [31]. Operation at a pilot plant treating coal power station flue gas saw no
appreciable loss of piperazine through the vapour phase during operation [32].
A high vapour pressure does not necessarily rule out an absorbent for CO2 capture
applications. Instead, additional engineering controls may be required. An example
is the ammonia process where absorption is completed at low temperatures to
Process Implications of CO2 Capture Solvent Selection 37

minimise evaporative losses. Careful design of downstream cleaning equipment


(e.g. the absorber wash section) can also minimise absorbent and other emissions to
the environment [33].
Increasingly there is an interest in solvents that undergo a phase change as a
result of the reaction with CO2 to form 2 liquid phases [34] or a slurry phase.
Processes can be developed to take advantage of this phase separation to reduce the
energy demand by sending only the phase rich in reacted solvent to the stripper and
recycling the unreacted (lean) phase back to the top of the absorber. In each case it
is necessary to know the equilibrium between all the possible phases—VL1L2E
(Vapour-Liquid1-Liquid2 Equilibria) or VLSE (Vapour-Liquid-Solid Equilibria). It
is also necessary to know the rate and conditions of temperature and pressure at
which the additional phases form and separate to determine equipment operating
conditions while the physical nature of the phases (droplet and particle size dis-
tribution, surface tension, density difference) will determine the design and cost of
the additional separation equipment.
An aqueous solution of NH3 can be considered, in one respect, as being the
ultimate amine, with an hydrogen atom attached to the NH2—functional group
rather than an organic tail. Ammonia is cheap to produce (relative to other amines),
does not degrade, reacts more rapidly and has lower energy demands (at compa-
rable concentrations) and is environmentally benign (relative to other amines).
Although ammonia is normally a vapour, the very high solubility in water makes
aqueous ammonia a potential capture solvent.
The process issues with aqueous ammonia include the loss of ammonia in the
vapour streams (slip) due to its high vapour pressure, the formation of solid
ammonium bicarbonate deposits in vapour streams leading to line blockages and
the low reaction rates and high energy requirements which are a direct consequence
of the very low concentrations and temperatures that must be applied to reduce the
slip problem.

3.3 Solvent Viscosity and Diffusivity

The viscosity of the solvent typically increases with the increasing molecular
weight of the amine, with increasing amine concentration and with decreasing
temperature. Viscosity also typically increases with loading, and this should be
determined experimentally as a function of the loading. The increase in viscosity
has two effects on the performance of the solvent as a capture agent. Firstly, the
diffusion coefficient is inversely proportional to viscosity, so that the mass transfer
coefficient will decrease accordingly. Secondly, in a packed column, surface is
generated by the flow of the liquid on the inclined surfaces of the packing under
gravity. Increasing viscosity will lead to thicker fluid layers on the packing (lower
specific surface area per unit liquid volume), greater hold-up and longer residence
times. It is usually not recommended to use a packed bed with capture solvents with
a viscosity greater than 5 mPas [35], although more recent work suggests that this is
38 L.T. Wardhaugh and A. Cousins

not as serious a limit as originally thought. Viscosity also affects the liquid-side
mass transfer coefficient through affecting turbulence in the liquid phase.
Absorbents with a viscosity that is too low will not form suitable liquid films over
column packing. Thus, for absorbents whose viscosity falls outside the suitable
range for packed column applications, alternative mass transfer devices should be
considered.
Tray columns use gas pressure to generate surface area through the bubbling and
spray formed as the gas passes through the cross-flowing liquid. Increasing vis-
cosity severely reduces the formation of spray and increases the energy required to
form bubbles. For a similar reason, spray columns will not form fine sprays above
approximately 50 mPas. Rotating contactor equipment creates surface through the
addition of centrifugal energy (at a cost) and is therefore applicable to a much wider
range of solvent viscosities. The very high specific surface areas (per unit liquid
volume) that are possible in rotating contactors and consequent thin liquid layers
can compensate for the reduced diffusion rates by providing a shorter path length to
the solvent reactant. This is discussed further in Sect. 5.
Viscosity also plays a key role in the performance of heat exchanger equipment
through its effect on Reynolds Number. This is of particular concern in the reboiler
where laminar flow or poor flow distribution could lead to hot spots and more rapid
thermal degradation of the amine. The performance of the cross heat exchanger will
be determined by the difference in viscosities between the rich and lean solvents as
discussed above.
The viscosity of various loaded amine systems has been investigated by a
number of researchers including Freeman and Rochelle [36], Amundsen et al. [37],
Kumar et al. [38], Shokouhi et al. [39], Weiland et al. [40]. Diffusion and viscosity
are closely related via an inverse relationship. Diffusion coefficients for CO2 into
absorbents are often calculated via the N2O analogy. This states that the ratio of
diffusion coefficients between N2O and CO2 in water is equal to that in an aqueous
amine solution [29].

3.4 Surface Tension

The surface tension of a liquid seeks to minimise the surface area of the fluid, e.g. by
forming droplets. This is a result of the unbalanced forces acting on liquid molecules
at the surface. Molecules within a liquid are pulled in all directions by intermolecular
forces. However, molecules at the surface have no intermolecular forces pulling them
upward away from the surface. These intermolecular attractions thus tend to pull the
surface molecules into the liquid and cause the surface to tighten like an elastic film
[41]. Liquids that have strong intermolecular forces also have high surface tensions.
The surface tension of a fluid is an important consideration in gas/liquid contacting
devices as it affects the distribution of liquid over packing (or other contactor) sur-
faces (greater spreading associated with lower surface tension), liquid hold-up and
Process Implications of CO2 Capture Solvent Selection 39

the formation and breakup of foams. Work completed on packed columns under CO2
absorption conditions determined that for high surface area packing materials,
reduced surface tensions could increase the packing effective area, thought to be due
to removal of capillary phenomena [42]. Lowering the surface tension of fluids via
surfactant addition, however, can have deleterious effects. Sedelies et al. [43]
investigated the effect of surface tension on packing effective area, however sur-
factant addition to achieve low surface tensions resulted in significant foam forma-
tion. Tsai et al. [42] similarly had to add an anti-foaming agent to their solution in
order to perform low surface tension experiments.
Surface tension is a measure of the amount of energy required to stretch or
increase the surface area of a liquid by a unit area [41] and is difficult to measure
precisely. As surface tension manifests itself in a number of phenomena, there are a
number of different techniques to measure it (e.g. du Noüy ring, du Noüy-Padday,
Wilhelmy plate, spinning drop, pendant drop, bubble pressure, drop volume, cap-
illary rise etc.) with many methods using tensiometers or microbalances to measure
the weight of the liquid meniscus.
The surface tension of a fluid will decrease with an increase in temperature,
becoming zero at its boiling point and disappearing at the critical temperature. The
surface tension of a liquid will also be affected by impurities and degradation
products. An example is the lowering of surface tension through the addition of
surfactants.

3.5 Chemical Stability and Solvent Degradation

Amines undergo the following irreversible reactions that will consume the costly
capture solvent, create deposition layers especially on the reboiler heat transfer
surfaces that must be periodically cleaned, form by-products that accumulate in the
solvent [e.g. heat stable salts (HSS)] and must be removed, or they may sponta-
neously leave with the gas streams increasing the emission of toxic materials. The
nature and extent of chemical stability is determined by the nature of the solvent
and the operating parameters and must be studied carefully. The impact of chemical
stability and the possible emission of by-products may not be fully understood until
the solvent in question is run for extended periods with full recycle including
regeneration at pilot scale (e.g. Radgen et al. [44]) or at demonstration scale such as
the SaskPower PCC demonstration plant at the Boundary Dam Power Station,
Saskatchewan [45].
There are 4 main processes leading to the degradation of the active capture
species all of which lead to a loss in cyclic CO2 loading and increased costs.
• Thermal degradation—occurring mainly in the reboiler leading to more volatile
by-products or free radicals that can subsequently polymerize forming the
deposition layers on heat transfer surfaces. The presence of ammonia gas in the
exit flue gas stream is the clearest indicator of thermal degradation. The process
40 L.T. Wardhaugh and A. Cousins

consequence is the limitation placed on the reboiler tube surface temperature


leading to the 120 °C limit for MEA.
• Oxidative degradation—as oxygen is always present in flue gases (furnaces
must be run with some degree of excess air to ensure complete combustion), a
small amount will always be dissolved in the solvent and this is sufficient to
form by-products such as formic acid, formaldehyde, etc.
• Reaction with other acid gases. All amines react with the acid gases (SO2; SO3;
NO2; N2O3) that are always present to some extent in flue gases by similar
reaction pathways as CO2, however the products of reaction will not decompose
back to the reactants in the stripper and are hence termed ‘heat stable salts’
(HSS). These are typically removed in a small heat exchanger, termed a
Reclaimer, operated at sufficient temperature to boil off the amine returning it to
the process. The residue from this vessel is bled off and should be considered as
a hazardous waste.
• NOx (nitrogen oxides) will readily react with secondary amines (R-NH-R’) to
form nitrosamines which have been of particular concern until engineering
solutions were found to mitigate the impact of these reactions [46–48].

3.6 Corrosivity

Unloaded aqueous amines are mildly corrosive to carbon steel, however, the cor-
rosivity of solvents typically increases as they degrade and absorb acid gases from
the flue gas [49]. The experience gained in natural gas processing plant provides
some insight into likely corrosion under PCC conditions. One critical difference
between natural gas processing conditions and PCC applications is the oxygen
content of the gas. This is typically minimal under natural gas processing condi-
tions, but can be in excess of 10 vol.% for power station flue gases being the
feed-stock to PCC plants. Most amines will degrade in the presence of oxygen, and
a number of the oxidative degradation products are corrosive. Thus corrosion under
PCC conditions will likely be higher than that witnessed to date in natural gas
processing plant. In addition, a number of the techniques used to minimise corro-
sion in natural gas plant, such as using an inert gas blanket to minimise oxygen
ingress, will not be applicable under the higher oxygen conditions in PCC. Severe
corrosion has been linked to amine degradation [1]. As a result, amines that are
susceptible to oxidative degradation, such as diethanolamine (DEA), are less suit-
able for PCC applications.
Degradation products formed through reactions with oxygen are noted to be
particularly acidic [50]. These include oxalic, glycolic, formic and acetic acid salts.
These are stronger acids than the carbonic acid formed through the absorption of
CO2, and will build-up in amine solutions over time. Unlike the absorbed CO2,
these salts cannot be thermally regenerated, and hence are known as Heat Stable
Process Implications of CO2 Capture Solvent Selection 41

Salts (HSS). There are recommended upper limits for these salts in most commonly
used amines [1]. When the salt content gets too high, a slip stream of the lean amine
solution will generally be sent for reclamation via caustic addition and batch dis-
tillation. For DEA and methyl-diethanolamine (MDEA) solutions, caustic addition
is recommended once HSS level reaches 0.5 wt% [1].
Different amines will have different levels of corrosivity due to the species
formed when they react with CO2 and their chemical stability. Typically, primary
amines tend to form more corrosive solutions, followed by secondary amines,
whilst tertiary amines the least. The reason for this ranking is not well understood.
One suggestion relates to the amines ability to form a stable carbamate [51, 52]. For
this reason, tertiary amine absorption liquid solutions based around MDEA have
become popular for natural gas processing. However, due to the low CO2 con-
centration and low flue gas pressure, PCC applications will likely require
fast-reacting amines, such as primary amines or to use tertiary amines in combi-
nation with a ‘promoter’.
As amines with high concentration are likely to form higher concentrations of
acidic degradation products, there are maximum concentration recommendations
for most common absorbents in order to limit corrosion. MEA for example has a
recommended maximum concentration of 20 wt% under natural gas processing
conditions [1]. However there are benefits in using higher concentration absorbents,
such as reduced absorbent and energy requirements. As a result, many proprietary
solutions use higher absorbent concentrations, but will also include corrosion
inhibitors. An example is the Fluor Econamine FG Plus utilizing a higher con-
centration of MEA coupled with proprietary corrosion inhibitors [53].
As corrosion rates tend to increase with an increase in CO2 loading, highly
loaded absorbents will suffer from greater corrosion affects. As a result, there are
recommended upper limits on acid gas loading for most commonly used amines [1].
Under PCC applications, absorbent loading is often altered in order to minimise the
regeneration energy requirement of the process. Kittel, Gonzalez [54] suggest
optimum operating conditions for a 30 wt% MEA solution occur at a lean loading
of 0.25 mol CO2/mol MEA. This is above the lean loading limit typically recom-
mended. Thus operation under optimum energy conditions for PCC applications
may result in higher levels of plant corrosion.
Corrosion effects can be minimised through engineering design. Minimising acid
gas flashing and turbulence through avoiding restrictions, tight corners, pump
cavitation and minimising absorbent flow rates, are recommended [51]. Where
engineering controls are not sufficient, corrosion inhibitors are often used. Inhibitor
choice needs to be made carefully as inhibitors can affect plant operation (e.g.
through increased foaming). Incorrect inhibitor choice can also increase oxidative
amine degradation, potentially increasing corrosion rates [55]. Use of inhibitors can
allow the use of lower grade, cheaper metals for construction. Where this is not
sufficient, higher grade metals may be required. In natural gas processing plant,
known corrosion prone areas include the base of the absorber column and high
temperature areas such as the base of the stripping column [1]. As such, stainless
steel is recommended for these sections, whilst other areas of the plant can be
42 L.T. Wardhaugh and A. Cousins

constructed from lower grade carbon steel. Corrosion rates under PCC conditions
are expected to be higher than those typically seen in natural gas processing. As a
result, the use of cheaper carbon steels for plant components might not be possible,
thus other low cost construction methods are being evaluated. The absorber col-
umns at both the Test Centre Mongstad (TCM) [56] and the Boundary Dam CO2
capture plant [45] use ceramic lined concrete towers for their absorber columns.
Novel absorbents being developed for CO2 capture applications often have
limited information about their long term stability and corrosiveness under flue gas
conditions. Significant research work is being conducted into corrosion rates and
the formation of degradation products under PCC conditions. In addition, novel,
low-cost construction materials are also being evaluated [57–59]. Advanced sol-
vents such as ionic liquids or amino acids display little or no corrosivity giving
these solvents a distinct advantage.

3.7 Material Compatibility

While corrosivity addresses the interaction between solvents and metal compo-
nents, material compatibility deals with the interaction between solvents and
non-metallic components such as seals, gaskets etc. The development of new sol-
vents will inevitably raise questions of compatibility with the materials of plant
construction and the components of equipment such as pumps, blowers, com-
pressors and instrumentation. These complex chemical interactions must be
established through laboratory and long term pilot or demonstration scale testing.
Apart from the contactor vessels themselves, points of particular concern include
the rich absorption liquid lines at the base of the absorber (containing the highest
quantities of contaminant materials), rich absorption liquid exit from the lean/rich
cross heat exchanger (due to the increase in temperature at this point), the hot lean
absorption liquid exiting the stripping column [1] and the absorber wash
(post-treatment) section [58].
While suitable metals (e.g. various grades of stainless steel) can be found to
achieve corrosion resistance, this adds significantly to the cost of these enormous
contactor columns. To reduce the cost of the absorber and pre-treatment columns,
the SaskPower Boundary Dam project adopted ceramic tile lined concrete vessels
[60]. Extended operation of this facility will provide a wealth of information on not
only the chemical and physical resistance of the tiles, but also the adhesive and
grouting materials. The Niederaussem pilot plant in Germany has been evaluating a
plastic-lined concrete module [57], to achieve further reductions in construction
cost.
Process Implications of CO2 Capture Solvent Selection 43

3.8 Other Properties

When selecting an energy efficient absorbent for CO2 capture applications, items
such as the absorbent’s reaction rate with CO2, its cyclic loading, and energy
requirements are often used for initial screening. However, before an absorbent is
used at pilot-scale, other factors need to be considered. The full evaluation of the
performance of a new solvent formulation requires additional physical properties to
be experimentally determined, as a function of temperature, pressure, solvent
concentration, and loading, including the following:
• Heat capacity. Necessary for heat balance and heat transfer calculations and
hence the evaluation of heat exchanger performance and total heat requirement.
Note that heat capacity can be a strong function of loading [61].
• Thermal conductivity. Necessary for heat transfer calculations and hence the
evaluation of heat exchanger performance.
• Density. A fundamental property affecting fluid flow, drainage rates and key to
any multi-phase separation processes [37, 62].
• pH. The net outcome of the reactions occurring in the process. This is more
likely to be measured in the process directly as an indication of the extent of
reaction (loading). Refer to Sect. 6.5.

4 Process Options and Design Relating to Solvent Choice

This section considers the impact of solvent choice on the selection and design of
individual unit operations.

4.1 Gas-Liquid Contactors

The central elements of a PCC plant are the absorption column (absorber) and the
desorption column (stripper). Conventional design of the PCC gas-liquid contactors
at commercial scale foresees the use of multi-train packed columns with random
packing or (increasingly) structured packing. Alternate contactor types are con-
sidered in Sect. 5. There are numerous design procedures in the literature with some
degree of discrepancy between them, making pilot scale validation essential. It is in
the validation and scale-up process that pit-falls relating to solvent choice occur as
discussed below.
44 L.T. Wardhaugh and A. Cousins

4.2 Packed Bed Contactors

The packed bed operates by the intimate contact of the liquid which flows under
gravity and (ideally) spreads over the packing material surface with maximum
coverage, while the gas travels upwards by pressure differential in the space left
between the liquid coated surfaces (i.e. counter-current and co-continuous). CO2
gas (and other residual acid and inert gas impurities) diffuse into the liquid at the
interface and react with the capture solvent. The reaction depletes the interface of
CO2 encouraging more CO2 to pass through the interface. Physical absorption of
the gas components is also occurring, limited by the diffusion rate and equilibrium
capacity of the capture solution for each gas component. The ‘lean’ capture solution
is distributed evenly across the top of each of the packing sections then trickles
down under gravity over the packing surface which is designed to maintain as even
and complete a gas-liquid interface as possible, becoming ‘enriched’ in CO2 as it
travels the length of the column and collects in the column base.
As there is a considerable variation in the flow stream compositions, tempera-
tures and (to a lesser extent) pressure and flowrate along the column, a differential
approach is usually taken and a rate based calculation procedure, utilizing one or
more film layers, avoids unrealistic assumptions about equilibrium between phases.
Computerized methods are commercially available which allow a choice of design
procedures. General advice is to not mix design procedures and to check that the
validation of the selected procedure covers the physical property range of the
selected solvent (especially viscosity).
The gas velocity in the contactor column is set to just beyond the ‘load point’
(below which the gas and liquid flows do not affect each other) and where there is a
slight improvement in the mass transfer rate. At higher flowrates, of either liquid or
gas, the pressure drop increases dramatically and/or becomes unstable and the mass
transfer rate falls at which point the column is said to have ‘flooded’. The optimum
operating point is therefore defined as a certain % flood (gas velocity ratio to the
‘flooding velocity’) or as a specific pressure drop per unit height of packing (DP/H)
and this defines the diameter of the contactor columns and, at larger scales, the
number of columns. At lower liquid rates, especially if liquid distribution across the
packing is poor (distributer design or surface tension effects), gas may by-pass part
or all of the liquid reducing the amount of CO2 that can be captured.

4.3 Absorption Column

The choice of capture solution determines the height of the absorber through the
rate of reaction and the diffusion rate of CO2 into the solution. Detailed calculation
procedures and automated design processes are available based on laboratory
measurements of vapour-liquid equilibria (Sect. 3.2); reaction kinetics (Sect. 3.1)
and physical properties, especially viscosity and diffusivity (Sect. 3.3).
Process Implications of CO2 Capture Solvent Selection 45

The amount of CO2 that can be captured in the absorber is fixed by the difference
between the rich loading (determined for the most part by the absorber design) and
the lean loading (determined by the stripper design and operation). A taller absorber
column will provide the residence time and interfacial area but this may be inef-
ficient. Running the stripper at higher reboiler rates (more excess steam) to remove
more CO2 (delivering a leaner return capture solution to the top of the absorber)
may also be inefficient in terms of energy consumption. To determine the optimum
operating point for a given capture solvent a series of experiments is carried out on
pilot facilities in which the Liquid to Gas mass flowrate ratio (L/G) is varied while a
fixed value of %capture is maintained to determine the minimum total energy input.
Higher liquid rates (at fixed gas rate) will capture more CO2 but require more
energy (per unit CO2 captured) due to the higher sensible heat requirement (Eq. 6).
Lower liquid rates will require that the returning lean solvent be stripped to a greater
degree in order to achieve the same % capture again requiring more energy per unit
CO2 captured.
It is clear from the previous discussion, however, that if the flows are varied,
then the column has moved away from its optimum operating point (% flood or
DP/H) which will also affect the results. Clearly it is impractical to run each L/G
point for each solvent in a different optimally designed column, but this is one of
the reasons that there is little agreement as to optimum L/G and minimum energy
requirements for a given solvent. Some piloting facilities address this issue in part
by having a variable liquid entry point thus varying the effective height of packing
for the different capture solutions and L/G ratios being tested. Tests would be
carried out, for example, by choosing liquid entry points that give equivalent rich
loadings. A more definitive approach would be to use the pilot data to thoroughly
validate process models to a high level of confidence and to use the models to
determine operating points for a range of L/G ratios with each point representing an
optimum column design. Such models would then be used to determine the energy
minimum and (more cogently) the economic optimum and this used to compare the
choice of capture solutions.
Absorption is accompanied by a release of heat (the heat of reaction, equivalent
to the heat of vaporization of CO2) some of which is transferred to the gas stream
but most of which leads to an increase in temperature within the column (reducing
the equilibrium loading) and increasing the rich solvent exit temperature. Control of
this temperature bulge is desirable as this leads to an increase in the cyclic loading.
As a result, a number of cooling methods have been proposed to minimise the effect
of the temperature bulge, allowing the column to operate in a more isothermal
manner [63, 64]. The most suitable location for inter-cooling on the absorber
column is not necessarily the location of the temperature bulge. The CO2 con-
centration in the vapour phase will typically decrease as the flue gas moves up an
absorber column. It is the difference between this gas phase CO2 partial pressure
and the equilibrium CO2 partial pressure that provides the driving force for mass
transfer to occur. When these two pressures approach the same value, the driving
force for mass transfer approaches zero, and the column is said to be ‘pinched’.
Inter-cooling can have a significant effect on the equilibrium CO2 partial pressure,
46 L.T. Wardhaugh and A. Cousins

and the location of inter-cooling is best placed where it can maximise the difference
between the equilibrium and operating CO2 partial pressures in the column.
Under PCC conditions, this is typically found to be in the lower portion of the
absorber column, where equilibrium CO2 partial pressures are higher [65], how-
ever, there are situations where intercooling towards the top of the absorber column
is advantageous (e.g. to reduce the evaporation of a more volatile solvent com-
ponent). Multiple inter-cooling steps may also be economically desirable.

4.4 Desorption Column

The desorption column consists of:


• A large stripping section (most of the column—the reason that it is often
referred to as the ‘stripper’) in which CO2 is stripped out of the capture solution
by the action of heat and reduced partial pressure provided by the steam, gen-
erated in the reboiler, that is travelling up through the column packing and
counter-current to the downflowing liquid; and
• A small rectification section in which the returning reflux (all of the condensed
excess steam) re-captures as much of the vaporized amine as possible. The
amount of amine that enters the vapour stream is determined by the equilibrium
vapour-liquid relationship which is a function of composition, temperature and
loading as discussed in Sect. 3.2 and can be significant for certain choices of
capture solution. The rectification section may also comprise a small section in
which rich solvent, split off prior to entering the cross exchanger, enters the
column above the normal feed point to retrieve some of the energy in the exiting
steam (‘cold rich split’ process modification). The merits of this process mod-
ification also depend on the choice of capture solvent.
The desorption column is usually smaller than the absorber in both height and
diameter and may consist of trays or packing or a combination of the two in
different sections. The column may also run at elevated pressures or different
pressures in different sections [66] to minimize the column size (though not nec-
essarily the cost); to minimize the loss of amine to the vapour stream; and to reduce
the cost of the CO2 product compression by providing this product stream partially
pressurized. Once again, the choices here are economic and to a large degree
determined by the choice of capture solution.
The major energy consumer of a PCC plant is the reboiler, however, energy
savings can be achieved throughout the system and particularly in the desorber
through more efficient design and energy saving modifications [6] the relative effect
of each of which are also determined by the choice of capture solution.
Process Implications of CO2 Capture Solvent Selection 47

4.5 Heat Exchangers

In terms of choice of solvent, the major impact on heat exchanger design and
operation is through:
• The steam temperature and hence pressure that must be provided to achieve the
optimum degree of solvent stripping,
• The sensitivity of the solvent to degradation in the reboiler, and
• The change in solvent properties (especially density, viscosity, thermal con-
ductivity and heat capacity) with the change in CO2 loading.
In a power station retrofit design the steam supply rate and pressure will impact
the operation of the power station steam circuit requiring modifications such as
replacement power turbines [4]. The use of 30 wt% MEA as capture solvent
requires a supply of low pressure (LP) steam from the power station steam circuit
due to the reboiler limiting temperature of approximately 120 °C. At this temper-
ature, an adequate degree of stripping can be achieved. For concentrated piperazine
(PZ) as capture solvent, higher stripper temperatures are possible due to the lower
degradation rates which may require that the steam supply to the reboiler be sourced
from the Medium Pressure (MP) section of the power station steam circuit thus
presenting a different design task [32, 67, 68].
It may be desirable to operate the stripper column at higher pressures to mini-
mize solvent losses and reduce the CO2 product compression costs. Higher stripper
operating pressures also require higher reboiler steam pressures (to maintain the
temperature driving force) or possibly larger heat transfer areas (larger equipment).
Regeneration at higher pressures will require the stripper column to be designed as
a pressure vessel, leading to higher costs.
The degree and type of degradation of the solvent increases with temperature
and it is on the heat exchange surfaces of the reboiler that the highest temperatures
will be experienced. The partial vaporization of solvent that occurs in the reboiler
could lead to local hot-spots that in turn lead to higher rates of degradation than
expected. Higher rates of circulation through the reboiler can alleviate this effect.
The common types of reboiler systems include:
• Forced circulation reboiler, in which a pump supplies lean solvent from the base
of the stripper column to the reboiler. This type is more expensive but easier to
control and vary as operating conditions change.
• Thermosyphon reboiler, in which the boiling solvent in the reboiler presents a
lower density than the solvent in the base of the stripper column with this
density difference driving a natural circulation of solvent through the reboiler.
For a thermosyphon reboiler to work effectively the solvent properties, espe-
cially density, viscosity, heat capacity and thermal conductivity must be well
understood as a function of temperature, solvent concentration and loading.
For some solvents the physical properties, especially viscosity and heat capacity
are a strong function of the loading. This will affect primarily the operation of the
48 L.T. Wardhaugh and A. Cousins

cross heat exchanger and the resulting outlet temperatures and overall system heat
balance. Adequate attention in design will provide allowances for this effect
especially during unsteady conditions.

4.6 Pumps

The pressures and temperatures of the conventional PCC plant are relatively
moderate and the selection of suitable pumping equipment is quite straight-forward.
Choice of capture solution will affect the selection of seals and pump components to
handle the corrosivity and material compatibility of the capture solvent. Particular
care should be taken of rubber components that could swell or degrade in the
presence of capture solvent, impurities or degradation products. Compatibility tests
must be carried out if specific compatibility information is not available, which is
particularly applicable to newly developed capture solvents.

4.7 Flue Gas Blower

There is not sufficient pressure in the typical power station flue duct to drive the flue
gas through the additional piping, columns and packing of the capture plant gas
stream. Therefore additional flue gas pumping capacity must be provided. This is an
axial fan or rotary blower usually placed at the outlet of the pre-treatment column
where the presence of corrosive impurities (e.g. SO2), residual particulates and gas
temperature are at a minimum. The flue gas is therefore drawn through the
pre-treatment column then pumped through the absorber and post-treatment column
sections in a single step. It may be more economical to place a second fan or blower
at the inlet or outlet of the post-treatment column to provide additional capacity.
The placement of the fan(s) or blower(s) affects the column pressure and hence the
absorption driving force, so that upstream placement is preferred, while on the other
hand, the total volume of gas is decreasing through the gas train, favouring
downstream placement. The water vapour content of the gas stream also affects the
design by changing the gas density.
The capture solvent does not normally contact the flue gas blower but impacts the
design and selection by its influence on the overall gas-side pressure drop. For
example, a less reactive capture solution will require either a taller column with a
more open packing (to maintain total pressure drop) and/or demand a higher pressure
to be delivered by the fan or blower. Note that each type of fan or blower will deliver
specific capacity and pressure limits with significant step changes in these values and
in the cost, so that the design task is not a simple linear relationship [19].
Process Implications of CO2 Capture Solvent Selection 49

5 Alternate Contactor Designs

Although packed columns with structured packing are most likely to be the con-
tactors in first generation PCC plants, alternate contactor types should be considered
especially in light of the wide range of capture solvents under development.

5.1 Spray Columns

The simplest and oldest gas-liquid contactor is the spray column although this
belies the vast research that has gone into perfecting the spray nozzle that is at the
heart of the device to deliver the droplet size distribution that is best suited to the
process. The spray column delivers the lowest gas phase pressure drop but has a
very short residence time that is not suited to the reaction times typical of amines
with CO2. The spray column does find application as the pre-treatment column
which is required if flue gas desulphurization (FDG) has not been implemented or
does not meet the specifications required by the selected CO2 capture solvent. Spray
columns have been investigated for PCC applications by [69–71] each of whom
show promising results. A different system has been proposed [72] and successfully
tested [73] in which the gas passes between sheets of liquid adsorbent sprayed
horizontally through the contactor chamber.
Liquids with viscosities above 50 mPas do not readily form droplets (rotating
contactors, discussed below, are better suited to such capture solutions). At lower
solution viscosities, the average drop size is determined by both the viscosity and
the energy input. Intuitively, a very fine droplet size would be ideal as this dra-
matically increases the interfacial area per unit volume of liquid, however this
proves to be counterproductive as the finer droplets are easily entrained and require
lower gas velocities and hence larger vessel diameters [74].

5.2 Tray Columns

Tray columns are the workhorse of distillation systems and processes. Liquid
travels across a tray from a ‘downcomer’ in a thin continuous layer through which
gas passes by means of holes (sieve tray designs); valve caps or bubble caps
mounted in the tray, then flowing over a weir that maintains the liquid layer and into
the downcomer of the next tray. Because of the backmixing that occurs due to
(desired) foaming and spray formation, each tray is considered to be single equi-
librium stage and (perhaps) many stages are required in a vertical column to achieve
the desired separation. Despite extremely high interfacial areas generated by the
formation of foam and spray on each tray, tray columns are less favoured for gas
absorption processes because of higher gas phase pressure drops (the cumulative
50 L.T. Wardhaugh and A. Cousins

pressure for the gas to force its way through multiple liquid layers); the extra height
on each tray needed for spray disengagement (to avoid liquid being carried back up
the column) thus adding to the total column height; and the extra diameter to allow
for the liquid downcomer and to restrict spray formation by reducing gas velocities
at the point of tray entry (holes or cap openings).
Tray columns may find application with specific capture solution choices whose
physical properties (especially viscosity and surface tension) may readily form foams
but not readily form sprays and therefore should not be completely discounted.

5.3 Rotating Contactors

With the introduction of the ‘Higee’ rotating packed bed by ICI in the 1970s [75],
the possibility of operating with a wider range of liquids in a more compact device
has opened up. Since then a wide range of rotating devices has been explored. The
application of rotating packed beds to CO2 capture from flue gases has been
explored by Cheng and Tan [76] who showed that the height of a transfer unit
(HTU) is significantly reduced, and Yu et al. [77] who showed that capture per-
formance in the rotating packed bed could be maintained despite a 7-fold increase
in the viscosity of a capture solvent chosen for its non-precipitating operational
advantages. Jassim et al. [78] investigated a range of MEA solvent concentrations
up to 100 wt% and showed that concentrations higher than the conventional 30 wt
% could give effective operation in physically smaller devices. The liquid side
volumetric mass transfer coefficient was increased by an order of magnitude in the
application of a rotating packed bed contactor to ionic liquids [79]. Ionic liquids
show promise as stable non-volatile capture solvents but are typically of higher
viscosity and lower reaction rate. This is a classic example of a promising capture
solvent looking for an appropriate process.
Rotating internals other than packing have been proposed by Makarytchev et al.
[80] who have investigated the performance of the inverted cone and Wang et al. [81]
who have shown the improved performance of a modified rotating disc contactor for a
range of applications including CO2 capture.
A recent development [82] that has a centrally located rotating tube as the only
internal with the liquid emitted as rotating continuous sheets taking the form of an
auger or blades, is able to provide surface area matching that of a packed bed and is
also able to pump the gas through the column, reducing or possibly even elimi-
nating the need for a flue gas blower.
The centrifugal force provided by rotation causes the liquid to spread more
evenly and more thinly on the available surfaces increasing overall mass transfer
rates and allowing capture solvents with higher viscosities (higher concentrations
and/or higher molecular weights) to be effectively utilized whereas application of
these solvents in a conventional packed bed would give inadequate performance
due to the lack of a driving force for fluid flow. While the mechanical issues may
for the moment preclude the use of rotating devices at the massive scale required
Process Implications of CO2 Capture Solvent Selection 51

by, for example, power station capture facilities, the process intensification afforded
by these devices may find a niche in smaller scale, space limited applications such
as off-shore or ship-board capture processes. Such applications would require the
specific development of a class of capture solvents to optimize the performance of
such devices.

6 Operation and Safety Considerations

6.1 Environmental Considerations

Post combustion capture using aqueous amines is currently the most technologi-
cally advanced method for removing CO2 from power station flue streams.
Under PCC conditions, release of amine and amine by-products to the environment
could occur through carryover with the exiting flue gas, or through spills from the
process. Thus consideration of the toxicity and potential environmental impact of
new and existing amines and other potential CO2 capture absorbents warrants
consideration.
Most amines are considered hazardous substances and are classified as dan-
gerous goods. They are corrosive, irritants, and can cause respiratory problems if
inhaled. As such, various levels of protective equipment may be required for
maintenance personnel working on CO2 capture plant where there is potential for
exposure to the absorbent. The choice of absorbent can have a significant effect on
the process requirements, and hence the level of personnel exposure. Precipitating
absorbents for example may have increased cleaning requirements to deal with
blockages. Highly corrosive absorbents may have a higher propensity for leaks. An
absorbent’s rate of degradation, and the type of degradation products formed can
also have a significant effect on its toxicity and safe handling requirements. Amino
acids have received attention as a potential CO2 capture absorbent as they are
considered environmentally friendly [83]. Eide-Haugmo et al. [84] evaluated the
eco-toxicity and biodegradability in the marine environment for a number of
absorbents being considered for PCC applications. Tertiary amines were found to
have a low biodegradability, whilst the amino acids evaluated had low toxicity and
high biodegradation potential.
In addition to emissions of the absorbent, formation and emission of degradation
and by-products should also be considered. Secondary amines are often preferred
for CO2 capture applications due to their fast reaction rates. Piperazine in particular
is already widely used as a rate promoter in blends with other amines (e.g. MDEA).
When applied to CO2 removal from combustion flue gases however, other issues
arise. All amines can form nitrosamines through reaction with nitrosating com-
pounds (e.g. NO absorbed from combustion flue gases), however, only secondary
amines will form a stable nitrosamine directly [85]. This is a concern as many
nitrosamines are carcinogenic. The potential for nitrosamine formation requires
52 L.T. Wardhaugh and A. Cousins

consideration for all amine systems as most amines will contain trace levels of other
amines, and can form a number of amine products as they degrade, however, this
will be a particular concern for systems using secondary amines. Currently, the
Norwegian Climate and Pollution Agency has set limits for the sum of all
N-nitrosamine and N-nitramine from exceeding 0.3 ng/m3 in air and 4 ng/L in
waters downwind of the Mongstad CO2 capture facility [85, 86]. Piperazine, a
secondary amine, was recently evaluated on a coal combustion flue gas [32].
During long term evaluation, the formation of mono-nitrosopiperazine (MNPZ) in
the absorbent was monitored. MNPZ concentration was noted to increase with
operation, reaching a steady-state level after approximately 200 h. Operating the
stripping column at higher temperatures (up to 155 °C) was noted to reduce the
level of MNPZ in the absorbent, thought to be due to thermal break-down of the
nitrosamine. A recent study [87], quantified emissions from a CO2 capture pilot
plant using an aged MEA solution to treat a coal combustion flue gas. Small
concentrations of Nitrosodiethanolamine were found in the solvent liquor, but was
not detected in the wash water or gaseous emissions from the plant.
As mentioned in Sect. 3.2 above, one of the items to consider when selecting an
absorbent is it’s volatility under CO2 capture conditions. This will not only affect
the economics of the process, due to absorbent make-up requirements, but also the
environmental implications of the process. Absorbents with a high volatility may
still be applicable to CO2 capture applications, but may require more extensive
downstream cleaning requirements. In addition to loss of absorbent, loss of volatile
degradation products (such as NH3) and the potential for further reaction in the
atmosphere requires consideration [88].
The capacity of most CO2 capture absorbents will likely decrease with continued
operation. This will be a result of the degradation of the absorbent and build-up of
by-products such as HSS. As such, commercial PCC processes will likely include
some form of amine reclamation. A common reclamation method is caustic addition
followed by batch distillation. This boils off the amine, which is then returned to the
process, whilst the degradation products and other unwanted material remains
behind as a waste sludge. Disposal of this waste requires careful consideration [89]
as it will be hazardous in nature. The extent and frequency of reclamation will
depend on the absorbent choice and operating conditions, with absorbents experi-
encing high rates of degradation producing higher rates of waste. One method for
dealing with this reclaimer waste is to return it to the power station furnace where it
can be burnt alongside the fuel. This requires careful consideration however as this
could require the power station to be re-classified as a hazardous waste combustor
in some jurisdictions [90].

6.2 Phase Separation—Planned and Unplanned

A number of promising absorbents for acid gas removal are often overlooked, or
their use limited, due to operating difficulties such as phase separation when used in
Process Implications of CO2 Capture Solvent Selection 53

a conventional absorption/desorption process. An absorbent that forms unwanted


precipitates during operation, for example, can clog pipework or column internals,
and increase corrosion rates through enhanced erosion.
An example of one such absorbent is piperazine (PZ). Piperazine is a secondary
amine that shows promise for post combustion CO2 removal applications due to its
fast reaction rates, low corrosivity and good stability (compared to the standard
30 wt% aqueous MEA) [67, 91]. Its concentration, however, is typically restricted
to below 10 wt% due to solubility issues. Aqueous piperazine will form precipitates
under ambient conditions at concentrations above this. As such, use of piperazine is
typically restricted to being a rate promoter (e.g. when blended with the tertiary
amine N-methyldiethanolamine, MDEA). Recent work however has shown that
concentrated piperazine solutions can be used for CO2 removal applications pro-
vided the CO2 loading of the absorbent is maintained between certain limits. Under
typical operating conditions (i.e. temperatures above 40 °C) the CO2 loading of an
8 molal (40 wt%) solution of piperazine should be maintained between 0.1 and
0.4 mol CO2/mol alkalinity [67] to remain soluble. If the plant is shut down
however, and lower temperatures are experienced, then a narrower range of CO2
loading is required to avoid solubility issues. When operating a plant, unplanned or
emergency shut down of the process must be considered. At the Stanwell
Corporation Ltd. owned Tarong CO2 capture pilot plant, heat tracing was applied to
solvent lines to minimise precipitation issues resulting from unplanned plant
outages when operating with an 8 molal piperazine solution [68]. This is an
example of trying to force fit the absorbent to a standard absorption/desorption
process. Precipitation from solution in unplanned shut-downs also affects advanced
energy efficient solvents such as amino acids.
Ammonia is another absorbent considered for CO2 capture applications where
precipitation can be an issue. Precipitation can occur in the absorber, where low
temperatures are used to minimise vapour losses, and also in the stripping column
condenser [92]. Precipitation however is not necessarily a problem. Alstom have
developed the Chilled Ammonia Process (CAP) which operates at absorber tem-
peratures of 5–10 °C. Under these conditions precipitation of ammonium carbonate
and bicarbonate will occur [93]. Alstom are currently demonstrating the chilled
ammonia process at a number of pilot plants worldwide [94]. Precipitation expe-
rienced in the stripping column condenser can also be avoided or reduced through
modifications to the standard process. Redirecting a portion of the cold rich solvent
to the stripping column condenser section could remove precipitation issues and
lower condenser cooling duties [95].
Phase separation can be beneficial to the CO2 removal process, particularly when
suitable process choices are made. Phase separation can be used to separate CO2
loaded absorbent, lowering the amount of absorbent sent to regeneration, poten-
tially lowering energy requirements. IFP Energies Nouvelles in France have
developed the DMX absorbents [96, 97]. These absorbents undergo liquid-liquid
phase separation as CO2 is absorbed (temperature dependant). The captured CO2
concentrates in one of the two liquid phases. This can be decanted and thus only a
54 L.T. Wardhaugh and A. Cousins

CO2 to
CO2 lean
Lean amine compression
flue gas

Cooler
Decanter
Lean amine Cooler
Absorber
Stripper
Rich amine

Blower Lean/rich heat


exchanger
Reboiler
Flue gas

Rich amine

Lean amine

Fig. 2 Simplified flow diagram of the DMX process based on Raynal et al. [97]

portion of the total flow is sent to the stripper for regeneration. A simplified flow
diagram of the DMX process is provided in Fig. 2.
Another instance of useful phase separation is in the combined CO2 and SO2
capture process (CASPER) being developed by TNO in the Netherlands [98] and
CSIRO in Australia [99]. Here, both SO2 and CO2 from a combustion flue gas are
captured by an amino acid solution. The CO2 loaded absorbent is regenerated via
standard thermal stripping. A slip stream of the lean absorbent is separated for
removal of the captured SO2. This is achieved by allowing K2SO4 to crystallise,
separating the sulphur compounds from the absorbent. The K2SO4 crystals are then
removed via filtration, and clean, regenerated absorbent is recycled to the absorber.
A flow diagram of the CASPER process is provided in Fig. 3.
An absorbent’s propensity to form precipitates is often identified during routine
laboratory analysis. Despite this, unexpected precipitation can still occur once an
absorbent has been transferred to an operating plant. Slightly different operating
conditions and the provision of nucleation sites often means precipitation can be
more of an issue in an operating plant. When operating with NH3 at the Munmorah
power station, dilute NH3 concentrations were used to minimise vapour losses. In
addition, absorber column temperatures closer to ambient were used to minimise
the likelihood of forming precipitates in this column. Despite these precautions,
precipitation of solid ammonium bicarbonate still occurred in the stripping column
condenser, causing shut-down of the facility [100]. CO2 lean flue gas emitted to the
atmosphere will likely contain trace absorbent vapour. The conditions at the flue
gas exit can sometimes be sufficient for localised precipitation to occur.
Process Implications of CO2 Capture Solvent Selection 55

Recycle stream
K2CO3

Crystallisation Filtration
unit Solvent + unit
solid K 2SO4
Wash
liquid K2SO4 Solids
To stack Filtration
Slurry unit K2SO4
precipitation
Wash Split
stream CO2 to
Filtrate
Cooler compression

Combined
SO2/CO2 Cooler
Cooler
absorption Absorber Condenser

Stripper
Condenser
pump

Blower Lean/rich heat


exchanger
Reboiler
Flue gas

Rich solvent CO2 removal


pump

Lean solvent
pump

Fig. 3 Process flow diagram of the CASPER process (based on Misiak et al. [98] and Cousins
et al. [99])

6.3 Pretreatment Requirements

When standard CO2 removal technologies are applied to combustion flue gases,
significant pre-treatment of the flue gas may be required. This is because trace
constituents in the flue gas (e.g. sulphur and nitrogen oxides) will react with most
alkanolamines forming heat stable salts. These salts bind strongly with the absor-
bent, lowering its ability to capture CO2, and can also increase the corrosiveness of
the absorbent solution. This removal of sulphur and nitrogen oxides from the flue
gas may be required even if upstream cleaning technologies such as selective
catalytic reduction (SCR) and flue gas desulphurisation (FGD) are in place. Current
flue gas cleaning technologies are designed for existing regulations. The European
directive for large combustion plants restricts SO2 emissions from power stations to
below 200 mg/Nm3 for new plants above 100 MWth. This limit however is still
above that recommended for most conventional amine technologies used for CO2
capture. Thus further treatment is often required before the flue gas enters the CO2
capture plant, either through an upgrade of the existing flue gas cleaning systems, or
56 L.T. Wardhaugh and A. Cousins

through installing new systems that could be equivalent in size and cost to the CO2
absorber column.
If upstream FGD and SCR are not in place, then the temperature of the flue gas
may also require adjustment. Flue gas temperatures of up to 180 °C are possible,
well above the typical operating temperature of most amine based absorbers (nearer
40 °C). In addition, depending on the flue gas source, removal of excess water or
saturation of the flue gas may be required. This is because most conventional CO2
capture technologies use aqueous solutions of alkanolamines. Water entering the
process with the flue gas, or leaving via evaporation, can significantly affect the
concentration of the absorbent. Thus monitoring and maintenance of the water
balance of the CO2 capture system is important.
Pre-treatment systems can be relatively simple, such as a caustic wash installed
upstream of the CO2 capture plant. However, due to the low pressure and large
volume of flue gas to be treated, pre-treatment systems will be large equipment in a
commercial CO2 capture plant. Rather than apply expensive pre-treatment to
combustion flue gases, new research is looking into the combined capture of
multiple pollutants, often removing the need for pre-treatment systems. This has the
potential to reduce the number of process stages required for the overall separation,
reducing cost. Examples include the Cansolv process currently in use at the
Boundary Dam coal-fired power station, the CASPER process [98, 99], the CS-Cap
process being developed by CSIRO in Australia [101], and the various processes
using ammonia (chilled ammonia process, Alstom [94]; ECO2, Powerspan [102]).
Again, fitting conventional technologies to combustion flue gases might not prove
to be the most economic strategy in the future. This is particularly likely in
countries such as Australia, where FGD and SCR are not currently used.

6.4 Foaming

An absorbent’s propensity to foam can have a significant impact on the effective-


ness and operability of the CO2 removal process. Mechanical foaming can be
caused by excessive gas velocities through the absorber [103]. Chemical foaming
can be caused by contaminants such as suspended solids, organic acids, corrosion
inhibitors, condensed hydrocarbons, grease, and degradation products [1, 50]. For
post combustion capture at coal-fired power stations, the potential for trace fly ash
to persist in the flue gas entering the CO2 capture plant will likely exacerbate
foaming issues. Sudden changes in column pressure drop or liquid level can
indicate foaming. Foaming can reduce the contact area between the gas and liquid
phases, reducing plant efficiency [50], and can also lead to excessive absorbent loss
[1]. Pressure and liquid level fluctuations can lead to unstable plant operation. The
best method for dealing with foaming is to remove the cause, for example through
avoiding or removing contaminants. Where this is not possible or sufficient,
anti-foaming agents can be added to the absorbent. An indication of an absorbent’s
foaming propensity, and the effectiveness of various anti-foaming agents, can be
Process Implications of CO2 Capture Solvent Selection 57

determined via simple laboratory analysis [50]. Different amines will exhibit
varying levels of foaming [104]. Whilst some absorbent characteristics can be
indicative of foaming (such as hydrophilic head with a hydrophobic tail), there is
currently no clear method for predicting quantitatively an amine’s propensity to
foam based on absorbent properties.
Foaming in PZ-MEA systems has been studied by Chen et al. [105] who noted
the complex effects and interactions of degradation products; metal ions; hydro-
carbons; additives such as oxidation inhibitors. Conventional antifoam agents will
generally assist in reducing foaming tendency though the complex relationships and
the impact of the antifoam agent on the mass transfer coefficient should be thor-
oughly studied in the laboratory and validated in pilot scale studies.

6.5 On-Line Solvent Property Determination

Once a plant is operational, the online measurement of various solvent properties


can provide valuable information concerning the performance of the plant. Key
performance indicators are the rich and lean loadings, the stability of the solvent
concentration (water mass balance), solvent loss via exiting gas streams and solvent
degradation. On-line analytical techniques are available or under development
including online automatic titrations or spectroscopic methods [106] that are cap-
able of measuring these parameters directly, but are usually expensive and complex
to operate, maintain and interpret. Simpler and less expensive measurement tech-
niques, as listed below, can provide at least qualitative information about changes in
plant performance, for example, a change in density and viscosity, but not in pH,
will indicate a change in the solvent concentration and indicate the action required
to correct the water imbalance. The relationship between the physical properties and
the key performance indicators is complex and requires detailed knowledge of these
properties as a function of solvent concentration, CO2 loading and temperature.
Online measurement techniques for gas streams include FT-IR spectroscopy
[107], non-dispersive infrared spectroscopy, chemiluminescence (e.g. for NO),
paramagnetic (e.g. for O2), and micro-gas chromatography [108]. For more detailed
analysis of absorbent trace contaminants or degradation products, more sophisti-
cated analysis measurements may be required [109]. Where manual sampling is
required for subsequent laboratory analysis, special techniques must be developed
when liquid droplets or particulates are present in the gas streams flow.
• On-line gas analysis. This is the routine determination of the capture rate using
IR or similar devices on the incoming and outgoing gas streams. These streams
must be completely dry to avoid damaging the instruments but can be com-
plemented by in-line humidity measurements as discussed below. Condensables
are either removed prior to the instrument, or the inlet lines and the measuring
device is maintained at high temperature to avoid condensation. More sophis-
ticated analyses to detect solvent loss, contaminants (SOx, NOx, O2),
58 L.T. Wardhaugh and A. Cousins

degradation products (especially ammonia, formic acid) are also possible and
increasingly utilized.
• On-line liquid analysis. The ideal situation is a direct measure of the CO2 loading,
solvent composition (to determine solvent loss, water balance) and the presence of
degradation products. In practice, these techniques are still under development
and require considerable calibration effort for each solvent composition and
loading. The National Carbon Capture Centre in the U.S.A. uses online automatic
titration for analysis of absorbent concentration and CO2 loading.
The following physical property measurements can be readily made on-line:
• pH—provides an indication of CO2 loading as the solution will become less
alkaline as CO2 is taken up by the capture solvent. pH is not a strong function of
temperature or solvent concentration.
• Density—online density measurements are possible using Coriolis meters
(usually in conjunction with flow measurements). Correlations between density
and liquid CO2 loading have been developed by Bui et al. [110] and Freeman
[111].
• Humidity can be readily determined in gas inlet and outlet lines but is usually
close to 100%. In this case it is possible that more water is present than detected,
having condensed on the cooler pipe and vessel walls.
• Viscosity can be determined directly using variants of the Coriolis meter,
although a pressure drop and flow measurement in any section of line or across a
restriction is an adequate measure of a change in viscosity and will provide a
strong indication of solvent concentration change (water balance issues) if
loading (capture rate) and temperature are relatively constant.
• Conductivity is a direct measure of the ionic species present and can be tied to
other measurements to indicate performance trends.

6.6 Analytical Requirements

Offline measurements to determine solvent and CO2 concentration include:


• Total alkalinity via standard acid/base titration [107],
• Specific absorbent concentration via gas or liquid chromatography [112].
• CO2 concentration is commonly determined by pH titration, or via the barium
chloride method [112].
In addition to the online measurements recorded at post combustion capture
plant, a number of off-line measurements will likely be required. Whilst work is
progressing towards online determination of absorbent concentration and acid gas
loading, as mentioned above, these techniques are complex and still under devel-
opment. As such, absorbent samples may still need to be collected for analysis
Process Implications of CO2 Capture Solvent Selection 59

offline of parameters such as absorbent concentration, acid gas loading, and extent
of degradation.
• Absorbent concentration. One of the key measurement requirements will be the
absorbent concentration. Absorbent concentration can vary through losses
(aerosol and evaporative), degradation, and lack of a water balance in the
process. It is desirable to maintain absorbent concentration at its target level.
Viscosity and corrosion will often increase with an increase in absorbent con-
centration, whilst energy requirements will increase with a decrease in con-
centration. A rough approximation of absorbent concentration can be made via
standard acid/base titration [107]. However, as number of degradation products
are also alkaline in nature, this method can over-estimate the actual absorbent
concentration. More accurate concentration measurements can be made for the
specific absorbent via gas or liquid chromatography [112]. Chromatographic
measurements however usually require sophisticated analytical equipment.
• Acid gas loading. Knowledge of the rich and lean acid gas loading of the
absorbent can be used to verify absorption rates, identify the operating region of
the process, and ensure high corrosion conditions are avoided. Absorbent CO2
concentration is commonly determined by pH titration, or via the barium
chloride method [112].
• Extent of degradation. As mentioned previously, absorbents used for CO2
removal will degrade over time. Some of these degradation products are cor-
rosive, and so there is generally a recommended maximum concentration for
heat stable salts in aqueous amines used for CO2 removal. In addition, an
absorbent that suffers from high degradation rates may prove a costly absorbent
choice due to high absorbent make-up rate requirements.
The formation of degradation products can be estimated via the loss in absorbent
concentration. This will not be completely accurate as absorbent concentration can
also be lost through other mechanisms (e.g. evaporative losses). Total alkalinity is
not a good method to use for determining absorbent concentration for this purpose
as some degradation products will also be alkaline in nature. Reduction in specific
absorbent concentration (e.g. via high pressure liquid chromatographic (HPLC)
determination) however can be used to provide this estimate. Where degradation
products are known, individual analysis can also be done e.g. HSS determination
via anion Ion Chromatography (IC) [113].

7 Conclusions

This chapter has discussed the process implications of selecting a particular energy
efficient CO2 capture solvent system and has explained how the unique physical
and chemical properties of the selected solvent impact on design choices, costs and
operability.
60 L.T. Wardhaugh and A. Cousins

Individual solvent properties, including reaction kinetics, thermodynamics and


physical properties have been outlined and their impact on design discussed.
Aspects such as degradation products, corrosivity and environmental considerations
impact the selection of process equipment and materials of construction which has a
further impact on cost. The impact of these properties of energy efficient solvents on
individual unit operations have also been discussed in some detail.
Regardless of the nature of the advanced energy efficient solvent system, these
fundamental aspects of design must always be taken into consideration. As all
capture solvent systems have advantages and challenges, the goal of process design
is to make the best use of the advantages and mitigate the impact of the challenges.
The thorough evaluation of a particular solvent system should be carried out in a
process design that has been optimized for that solvent. While this is not always
possible, process modelling tools allow the process design impacts to be taken into
account and scaled to a commercial scale design and costing.

References

1. Kohl AL, Nielsen R (1997) Gas purification, 5th edn. Gulf Publishing
2. Sander MT, Mariz CL (1992) The Fluor Daniel Econamine FG process—past experience
and present-day focus. In: 1st International Conference on Carbon Dioxide Removal
(ICCDR), Amsterdam, Netherlands, 04–06 Mar 1992, pp 341–348
3. Budzianowski WM (2015) Single solvents, solvent blends, and advanced solvent systems in
CO2 capture by absorption: a review. Int J Glob Warm 7(2):184–225. Doi:10.1504/ijgw.
2015.067749
4. Ramezan M, Skone TJ, Nsakala NY, Liljedahl GN (2007) Carbon dioxide capture from
existing coal-fired power plants. National Energy Technology Laboratory
5. Cottrell AJ, McGregor JM, Jansen J, Artanto Y, Dave N, Morgan S, Pearson P, Attalla MI,
Wardhaugh L, Yu H, Allport A, Feron PHM (2009) Post-combustion capture R&D and pilot
plant operation in Australia. In: Gale J, Herzog H, Braitsch J (eds) Greenhouse Gas Control
Technologies 9, vol 1. Energy Procedia. pp 1003–1010. Doi:10.1016/j.egypro.2009.01.133
6. Cousins A, Wardhaugh LT, Feron PHM (2011) A survey of process flow sheet modifications
for energy efficient CO2 capture from flue gases using chemical absorption. Int J Greenhouse
Gas Control 5:605–619. Doi:10.1016/j.ijggc.2011.01.002
7. Cousins A, Cottrell A, Lawson A, Huang S, Feron PHM (2012) Model verification and
evaluation of the rich-split process modification at an Australian-based post combustion CO2
capture pilot plant. Greenh Gases 2(5):329–345
8. Puxty G, Rowland R, Allport A, Yang Q, Bown M, Burns R, Maeder M, Attalla M (2009)
Carbon dioxide postcombustion capture: a novel screening study of the carbon dioxide
absorption performance of 76 amines. Environ Sci Technol 43(16):6427–6433
9. Plaza JM, Van Wagener D, Rochelle GT (2009) Modeling CO2 capture with aqueous
monoethanolamine. In: Gale J, Herzog H, Braitsch J (eds) Greenhouse Gas Control
Technologies 9, vol 1. Energy Procedia. pp 1171–1178. Doi:10.1016/j.egypro.2009.01.154
10. Zhang Y, Chen H, Chen CC, Plaza JM, Dugas R, Rochelle GT (2009) Rate-based process
modeling study of CO2 capture with aqueous monoethanolamine solution. Ind Eng Chem
Res 48(20):9233–9246. Doi:10.1021/ie900068k
11. Freguia S, Rochelle GT (2003) Modeling of CO2 capture by aqueous monoethanolamine.
AIChE J 49(7):1676–1686
Process Implications of CO2 Capture Solvent Selection 61

12. von Harbou I, Imle M, Hasse H (2014) Modeling and simulation of reactive absorption of
CO2 with MEA: results for four different packings on two different scales. Chem Eng Sci
105:179–190. Doi:10.1016/j.ces.2013.11.005
13. Razi N, Svendsen HF, Bolland O (2014) Assessment of mass transfer correlations in
rate-based modeling of a large-scale CO2 capture with MEA. Int J Greenhouse Gas Control
26:93–108. Doi:10.1016/j.figgc.2014.04.019
14. Dugas R, Alix P, Lemaire E, Broutin P, Rochelle G (2009) Absorber model for CO2 capture
by monoethanolamine; application to CASTOR pilot results. Energy Procedia 1(1):103–107
15. Saimpert M, Puxty G, Qureshi S, Wardhaugh L, Cousins A (2013) A new rate based absorber
and desorber modelling tool. Chem Eng Sci 96:10–25. Doi:10.1016/j.ces.2013.03.013
16. Yu YS, Lu HF, Wang GX, Zhang ZX, Rudolph V (2013) Characterizing the transport
properties of multiamine solutions for CO2 capture by molecular dynamics simulation.
J Chem Eng Data 58(6):1429–1439. Doi:10.1021/je3005547
17. Rubin ES, Short C, Booras G, Davison J, Ekstrom C, Matuszewski M, McCoy S (2013) A
proposed methodology for CO2 capture and storage cost estimates. Int J Greenhouse Gas
Control 17:488–503. Doi:10.1016/j.ijggc.2013.06.004
18. Dave N, Do T, Palfreyman D, Feron PHM, Xu S, Gao S, Liu L (2011) Post-combustion
capture of CO2 from coal-fired power plants in China and Australia: an experience based
cost comparison. In: Gale J, Hendriks C, Turkenberg W (eds) 10th international conference
on Greenhouse Gas Control Technologies, vol 4. Energy Procedia, pp 1869–1877. Doi:10.
1016/j.egypro.2011.02.065
19. Peters MS, Timmerhaus KD, West RE (2003) Plant design and economics for chemical
engineers, 5th edn. McGraw-Hill, Boston
20. Kierzkowska-Pawlak H, Chacuk A, Siemieniec M (2014) Reaction kinetics of CO2 in
aqueous 2-(2-aminoethylamino)ethanol solutions using a stirred cell reactor. Int J
Greenhouse Gas Control 24:106–114. Doi:10.1016/j.ijggc.2014.03.004
21. Conway W, Beyad Y, Maeder M, Burns R, Feron P, Puxty G (2014) CO2 absorption into
aqueous solutions containing 3-piperidinemethanol: CO2 mass transfer, stopped-flow
kinetics, H-1/C-13 NMR, and vapor-liquid equilibrium investigations. Ind Eng Chem Res
53(43):16715–16724. Doi:10.1021/ie503195x
22. Kunze A-K, Dojchinov G, Haritos VS, Lutze P (2015) Reactive absorption of CO2 into
enzyme accelerated solvents: from laboratory to pilot scale. Appl Energy 156:676–685.
Doi:10.1016/j.apenergy.2015.07.033
23. Bishnoi S, Rochelle GT (2000) Absorption of carbon dioxide into aqueous piperazine:
reaction kinetics, mass transfer and solubility. Chem Eng Sci 55(22):5531–5543
24. Dang HY, Rochelle GT (2003) CO2 absorption rate and solubility in
monoethanolamine/piperazine/water. Sep Sci Technol 38(2):337–357. Doi:10.1081/ss-
12016678
25. Jou FY, Otto FD, Mather AE (1994) Vapor-liquid-equilibrium of carbon-dioxide in aqueous
mixtures of monoethanolamine and methyldiethanolamine. Ind Eng Chem Res 33(8):2002–
2005
26. Kumar PS, Hogendoorn JA, Feron PHM, Versteeg GF (2003) Equilibrium solubility of CO2
in aqueous potassium taurate solutions: Part 1. Crystallization in carbon dioxide loaded
aqueous salt solutions of amino acids. Ind Eng Chem Res 42(12):2832–2840. Doi:10.1021/
ie0206002
27. Fan GJ, Wee AGH, Idem R, Tontiwachwuthikul P (2009) NMR studies of amine species in
MEA-CO2-H2O system: modification of the model of vapor-liquid equilibrium (VLE). Ind
Eng Chem Res 48(5):2717–2720. Doi:10.1021/ie8015895
28. Park SJ, Shin HY, Min BM, Cho A, Lee JS (2009) Vapor-liquid equilibria of water plus
monoethanolamine system. Korean J Chem Eng 26(1):189–192
29. Puxty G, Rowland R (2011) Modeling CO2 mass transfer in amine mixtures: PZ-AMP and
PZ-MDEA. Environ Sci Technol 45(6):2398–2405. Doi:10.1021/es1022784
62 L.T. Wardhaugh and A. Cousins

30. Li H, Le Moullec Y, Lu JH, Chen J, Marcos JCV, Chen GF, Chopin F (2015) CO2 solubility
measurement and thermodynamic modeling for 1-methylpiperazine/water/CO2. Fluid Phase
Equilib 394:118–128. Doi:10.1016/j.fluid.2015.03.021
31. Freeman SA, Dugas R, Van Wagener D, Nguyen T, Rochelle GT (2009) Carbon dioxide
capture with concentrated, aqueous piperazine. 10.1016/j.egypro.2009.01.195. In: Gale J,
Herzog H, Braitsch J (eds) Greenhouse Gas Control Technologies 9, vol 1. Energy Procedia,
pp 1489–1496
32. Cousins A, Nielsen P, Huang S, Cottrell A, Chen E, Rochelle GT, Feron PHM (2015)
Pilot-scale evaluation of concentrated piperazine for CO2 capture at an Australian coal-fired
power station: duration experiments. Greenhouse Gases 5(4):363–373. Doi:10.1002/ghg.1507
33. Moser P, Schmidt S, Stahl K, Vorberg G, Lozano GA, Stoffregen T, Roesler F (2014)
Demonstrating emission reduction—results from the post-combustion capture pilot plant at
Niederaussem. In: Dixon T, Herzog H, Twinning S (eds) 12th international Conference on
Greenhouse Gas Control Technologies, Ghgt-12, vol 63. Energy Procedia. pp 902–910.
Doi:10.1016/j.egypro.2014.11.100
34. Pinto DDD, Knuutila H, Fytianos G, Haugen G, Mejdell T, Svendsen HF (2014) CO2 post
combustion capture with a phase change solvent. Pilot plant campaign. Int J Greenhouse Gas
Control 31:153–164. Doi:10.1016/j.ijggc.2014.10.007
35. Song D, Seibert AF, Rochelle GT (2014) Effect of liquid viscosity on the liquid phase mass
transfer coefficient of packing. In: Dixon T, Herzog H, Twinning S (eds) 12th international
conference on Greenhouse Gas Control Technologies, Ghgt-12, vol 63. Energy Procedia.
Elsevier Science Bv, Amsterdam, pp 1268–1286. Doi:10.1016/j.egypro.2014.11.136
36. Freeman SA, Rochelle GT (2011) Density and viscosity of aqueous (piperazine plus carbon
dioxide) solutions. J Chem Eng Data 56(3):574–581. Doi:10.1021/je1012263
37. Amundsen TG, Oi LE, Eimer DA (2009) Density and viscosity of monoethanolamine plus
water plus carbon dioxide from (25 to 80) degrees C. J Chem Eng Data 54(11):3096–3100.
Doi:10.1021/je900188m
38. Kumar PS, Hogendoorn JA, Feron PHM, Versteeg GF (2001) Density, viscosity, solubility,
and diffusivity of N2O in aqueous amino acid salt solutions. J Chem Eng Data 46(6):1357–
1361
39. Shokouhi M, Jalili AH, Samani F, Hosseini-Jenab M (2015) Experimental investigation of
the density and viscosity of CO2-loaded aqueous alkanolamine solutions. Fluid Phase
Equilib 404:96–108. Doi:10.1016/j.fluid.2015.06.034
40. Weiland RH, Dingman JC, Cronin DB, Browning GJ (1998) Density and viscosity of some
partially carbonated aqueous alkanolamine solutions and their blends. J Chem Eng Data 43
(3):378–382
41. Chang R (1998) Chemistry, 6th edn. McGraw Hill, New York
42. Tsai RE, Schultheiss P, Kettner A, Lewis JC, Seibert AF, Eldridge RB, Rochelle GT (2008)
Influence of surface tension on effective packing area. Ind Eng Chem Res 47(4):1253–1260.
Doi:10.1021/ie0707801
43. Sedelies R, Steiff A, Weinspach P-M (1987) Mass transfer area in different gas-liquid
reactors as a function of liquid properties. Chem Eng Technol 10:1–15
44. Radgen P, Rode H, Reddy S, Yonkoski J (2014) Lessons learned from the operation of a 70
tonne per day post combustion pilot plant at the coal fired power plant in Wilhelmshaven,
Germany. In: Dixon T, Herzog H, Twinning S (eds) 12th international conference on
Greenhouse Gas Control Technologies, Ghgt-12, vol 63. Energy Procedia. Elsevier Science
Bv, Amsterdam, pp 1585–1594. Doi:10.1016/j.egypro.2014.11.168
45. Reitenbach G (2015) SaskPower’s Boundary Dam Carbon Capture Project Wins POWER’s
Highest Award. Power 159(8):24
46. Jackson P, Attalla MI (2010) N-Nitrosopiperazines form at high pH in post-combustion
capture solutions containing piperazine: a low-energy collisional behaviour study. Rapid
Commun Mass Spectrom 24(24):3567–3577. Doi:10.1002/rcm.4815
Process Implications of CO2 Capture Solvent Selection 63

47. Knuutila H, Svendsen HF, Asif N (2014) Decomposition of nitrosamines in aqueous


monoethanolamine (MEA) and diethanolamine (DEA) solutions with UV-radiation. Int J
Greenhouse Gas Control 31:182–191. Doi:10.1016/j.ijggc.2014.10.002
48. Fine NA, Rochelle GT (2013) Thermal decomposition of n-nitrosopiperazine. In: Dixon T,
Yamaji K (eds) Ghgt-11, vol 37. Energy Procedia. Elsevier Science Bv, Amsterdam,
pp 1678–1686. Doi:10.1016/j.egypro.2013.06.043
49. Pearson P, Cousins A (2015) Assessment of corrosion in amine-based post-combustion
capture of carbon dioxide systems. In: Feron P (ed) Absorption-based post-combustion
capture of Carbon Dioxide, in press edn. Elsevier (in press)
50. GPSA (2004) GPSA engineering data book. Tulsa, Oklahoma
51. Dupart MS, Bacon TR, Edwards DJ (1993) Understanding corrosion in alkanolamine gas
treating plants. 1. Proper mechanism diagnosis optimizes amine operations. Hydrocarbon
Process 72(4):75–80
52. Dupart MS, Bacon TR, Edwards DJ (1993) Understanding corrosion in alkanolamine gas
treating plants. Hydrocarbon Process 72(5):89–94
53. Johnson DW, Reddy S, Brown JH (2009) Commercially available CO2 capture technology.
Power 153(8):58–60
54. Kittel J, Gonzalez S (2014) Corrosion in CO2 post-combustion capture with alkanolamines—
a review. Oil Gas Sci Technol 69(5):915–929. Doi:10.2516/ogst/2013161
55. Bello A, Idem RO (2006) Comprehensive study of the kinetics of the oxidative degradation
of CO2 loaded and concentrated aqueous monoethanolamine (MEA) with and without
sodium metavanadate during CO2 absorption from flue gases. Ind Eng Chem Res 45
(8):2569–2579. Doi:10.1021/ie050562x
56. de Koeijer G, Enge Y, Sanden K, Graff OF, Falk-Pedersen O, Amundsen T, Overa S (2011)
CO2 Technology Centre Mongstad—design, functionality and emissions of the amine plant.
In: Gale J, Hendriks C, Turkenberg W (eds) 10th international conference on Greenhouse
Gas Control Technologies, vol 4. Energy Procedia, pp 1207–1213. Doi:10.1016/j.egypro.
2011.01.175
57. Moser P, Schmidt S, Uerlings R, Sieder G, Titz J-T, Hahn A, Stoffregen T (2011) Material
testing for future commercial post-combustion capture plants—results of the testing
programme conducted at the Niederaussem pilot plant. In: Gale J, Hendriks C,
Turkenberg W (eds) 10th international conference on Greenhouse Gas Control
Technologies, vol 4. Energy Procedia, pp 1317–1322. Doi:10.1016/j.egypro.2011.01.189
58. Cousins A, Ilyushechkin A, Pearson P, Cottrell A, Huang S, Feron PHM (2013) Corrosion
coupon evaluation under pilot-scale CO2 capture conditions at an Australian coal-fired
power station. Greenh Gases 3(3):169–184. Doi:10.1002/ghg.1341
59. Kittel J, Idem R, Gelowitz D, Tontiwachwuthikul P, Parrain G, Bonneau A (2009) Corrosion
in MEA units for CO(2) capture: pilot plant studies. In: Gale J, Herzog H, Braitsch J
(eds) Greenhouse Gas Control Technologies 9, vol 1. Energy Procedia, vol 1, pp 791–797.
Doi:10.1016/j.egypro.2009.01.105
60. Monea M (2013) Boundary Dam—progress and status. Paper presented at the 7th
Trondheim CCS conference, Trondheim 4–6 June 2013
61. Weiland RH, Dingman JC, Cronin DB (1997) Heat capacity of aqueous monoethanolamine,
diethanolamine, N-methyldiethanolamine, and N-methyldiethanolamine-based blends with
carbon dioxide. J Chem Eng Data 42(5):1004–1006
62. Shaikh MS, Shariff AM, Bustam MA, Murshid G (2014) Physicochemical properties of
aqueous solutions of sodium l-prolinate as an absorbent for CO2 removal. J Chem Eng Data
59(2):362–368. Doi:10.1021/je400830w
63. Aroonwilas A, Veawab A (2007) Heat recovery gas absorption process
64. Leites IL, Sama DA, Lior N (2003) The theory and practice of energy saving in the chemical
industry: some methods for reducing thermodynamic irreversibility in chemical technology
processes. Energy 28(1):55–97
64 L.T. Wardhaugh and A. Cousins

65. Cousins A, Wardhaugh LT, Feron PHM (2011) Preliminary analysis of process flow sheet
modifications for energy efficient CO2 capture from flue gases using chemical absorption.
Chem Eng Res Des 89(8A):1237–1251. Doi:10.1016/j.cherd.2011.02.008
66. Oyenekan BA, Rochelle GT (2007) Alternative stripper configurations for CO2 capture by
aqueous amines. AIChE J 53(12):3144–3154. Doi:10.1002/aic.11316
67. Freeman SA, Dugas R, Van Wagener DH, Nguyen T, Rochelle GT (2010) Carbon dioxide
capture with concentrated, aqueous piperazine. Int J Greenhouse Gas Control 4(2):119–124.
Doi:10.1016/j.ijggc.2009.10.008
68. Cousins A, Huang S, Cottrell A, Feron PHM, Chen E, Rochelle GT (2015) Pilot-scale
parametric evaluation of concentrated piperazine for CO2 capture at an Australian coal-fired
power station. Greenhouse Gases 5(1):7–16. Doi:10.1002/ghg.1462
69. Kuntz J, Aroonwilas A (2008) Performance of spray column for CO2 capture application.
Ind Eng Chem Res 47(1):145–153. Doi:10.1021/ie0617021
70. Seyboth O, Zimmermann S, Heidel B, Scheffknecht G (2014) Development of a spray
scrubbing process for post combustion CO2 capture with amine based solvents. In: Dixon T,
Herzog H, Twinning S (eds) 12th international conference on Greenhouse Gas Control
Technologies, Ghgt-12, vol 63. Energy Procedia. Elsevier Science Bv, Amsterdam,
pp 1667–1677. Doi:10.1016/j.egypro.2014.11.176
71. Kavoshi L, Rahimi A, Hatamipour MS (2015) CFD modeling and experimental study of
carbon dioxide removal in a lab-scale spray dryer. Chem Eng Res Des 98:157–167. Doi:10.
1016/j.cherd.2015.04.023
72. Awtry AR, Brasseur JK, Henshaw TL, Hobbs KR, McDermott WE, Miller NJ,
Neumann DK, Nizamov BR, Tobias JA, Anderson JL, Courtright JL (2009) Gas liquid
contactor module, useful for e.g. removing gas phase molecules, comprises liquid inlet, gas
inlet, gas outlet, array of nozzles in communication with the liquid inlet and gas inlet, gas
liquid separator and liquid outlet. CN102217152-B
73. Awtry A (2013) Status of the Carbon Dioxide Absorber Retrofit Equipment (CARE)
program. Paper presented at the 2013 CO2 Capture Technical Meeting
74. Wardhaugh L (2010) Toward significant capital cost reduction in post combustion capture
plants. Paper presented at the Chemeca 2010, Adelaide, South Australia, 26–29 Sept 2010
75. Ramshaw C (1983) HIGEE Distillation—an example of process intensification. Chem Eng
Lond 389:13–14
76. Cheng HH, Tan CS (2009) Carbon dioxide capture by blended alkanolamines in rotating
packed bed. In: Gale J, Herzog H, Braitsch J (eds) Greenhouse Gas Control Technologies 9,
vol 1. Energy Procedia, pp 925–932. Doi:10.1016/j.egypro.2009.01.123
77. Yu CH, Wu TW, Tan CS (2013) CO2 capture by piperazine mixed with non-aqueous solvent
diethylene glycol in a rotating packed bed. Int J Greenhouse Gas Control 19:503–509.
Doi:10.1016/j.ijggc.2013.10.014
78. Jassim MS, Rochelle G, Eimer D, Ramshaw C (2007) Carbon dioxide absorption and
desorption in aqueous monoethanolamine solutions in a rotating packed bed. Ind Eng Chem
Res 46(9):2823–2833. Doi:10.1021/ie051104r
79. Zhang L-L, Wang J-X, Liu Z-P, Lu Y, Chu G-W, Wang W-C, Chen J-F (2013) Efficient
capture of carbon dioxide with novel mass-transfer intensification device using ionic liquids.
AIChE J 59(8):2957–2965. Doi:10.1002/aic.14072
80. Makarytchev SV, Langrish TAG, Fletcher DF (2005) Exploration of spinning cone column
capacity and mass transfer performance using CFD. Chem Eng Res Des 83(A12):1372–
1380. Doi:10.1205/cherd.03413
81. Wang YD, Fei WY, Sun JH, Wan YK (2002) Hydrodynamics and mass transfer
performance of a modified rotating disc contactor (MRDC). Chem Eng Res Des 80(4):392–
400
82. Wardhaugh L, Allport A, Garland E, Solnordal C (2013) A novel gas-liquid contactor for
PCC based on a rotating liquid sheet. Paper presented at the The 7th Trondheim CCS
Conference (TCCS-7): CO2 Capture, Transport and Storage, Trondheim, Norway, 4–6 June
2013
Process Implications of CO2 Capture Solvent Selection 65

83. Knuutila H, Aronu UE, Kvamsdal HM, Chikukwa A (2011) Post combustion CO2 capture
with an amino acid salt. In: Gale J, Hendriks C, Turkenberg W (eds) 10th international
conference on Greenhouse Gas Control Technologies, vol 4. Energy Procedia, pp 1550–
1557. Doi:10.1016/j.egypro.2011.02.024
84. Eide-Haugmo I, Brakstad OG, Hoff KA, Sorheim KR, da Silva EF, Svendsen HF (2009)
Environmental impact of amines. In: Gale J, Herzog H, Braitsch J (eds) Greenhouse Gas
Control Technologies 9, vol 1. Energy Procedia, vol 1, pp 1297–1304. Doi:10.1016/j.
egypro.2009.01.170
85. Cousins A, Nielsen PT, Huang S, Rowland R, Edwards B, Cottrell A, Chen E, Rochelle GT,
Feron PHM (2015) Pilot-scale evaluation of concentrated piperazine for CO2 capture at an
Australian coal-fired power station: Nitrosamine measurements. Int J Greenhouse Gas
Control 37:256–263. Doi:10.1016/j.ijggc.2015.03.007
86. Dai N, Shah AD, Hu L, Plewa MJ, McKague B, Mitch WA (2012) Measurement of
nitrosamine and nitramine formation from NO reactions with amines during amine-based
carbon dioxide capture for postcombustion carbon sequestration. Environ Sci Technol 46
(17):9793–9801. Doi:10.1021/es301867b
87. Azzi M, Tibbett A, Halliburton B, Element A, Artanto Y, Meuleman E, Feron P (2014)
Assessing atmospheric emissions from amine-based CO2 post-combustion capture processes
and their impacts on the environment—a case study, vol 1. Measurement of emissions from
a monoethanolamine-based post-combustion CO2 capture pilot plant. 1
88. Azzi M, Angove D, Campbell I, Cope M, Emmerson K, Feron P, Patterson M, Tibbett A,
White S (2014) Assessing atmospheric emissions from an amine-based CO2
post-combustion capture processes and their impacts on the environment—a case study,
vol 2. Atmospheric chemistry of MEA and 3D air quality modelling of emissions from the
Loy Yang PCC plant, Final report. Global Carbon Capture and Storage Institute
89. Nurrokhmah L, Mezher T, Abu-Zahra MRM (2013) Evaluation of handling and reuse
approaches for the waste generated from MEA-based CO2 capture with the consideration of
regulations in the UAE. Environ Sci Technol 47(23):13644–13651. Doi:10.1021/es4027198
90. Sexton A, Dombrowski K, Nielsen P, Rochelle G, Fisher K, Youngerman J, Chen E,
Singh P, Davison J (2014) Evaluation of reclaimer sludge disposal from post-combustion
CO2 capture. In: Dixon T, Herzog H, Twinning S (eds) 12th international conference on
Greenhouse Gas Control Technologies, Ghgt-12, vol 63. Energy Procedia, pp 926–939.
Doi:10.1016/j.egypro.2014.11.102
91. Rochelle G, Chen E, Freeman S, Van Wagener D, Xu Q, Voice A (2011) Aqueous
piperazine as the new standard for CO2 capture technology. Chem Eng J 171(3):725–733.
Doi:10.1016/j.cej.2011.02.011
92. Yang N, Yu H, Li LC, Xu DY, Han WF, Feron P (2014) Aqueous ammonia (NH3) based
post combustion CO2 capture: a review. Oil Gas Sci Technol 69(5):931–945. Doi:10.2516/
ogst/2013160
93. Darde V, Thomsen K, van Well WJM, Stenby EH (2009) Chilled ammonia process for CO2
capture. In: Gale J, Herzog H, Braitsch J (eds) Greenhouse Gas Control Technologies 9, vol
1. Energy Procedia, pp 1035–1042. Doi:10.1016/j.egypro.2009.01.137
94. Telikapalli V, Kozak F, Francois J, Sherrick B, Black J, Muraskin D, Cage M, Hammond M,
Spitznogle G (2011) CCS with the Alstom chilled ammonia process development program—
field pilot results. In: Gale J, Hendriks C, Turkenberg W (eds) 10th international conference
on Greenhouse Gas Control Technologies, vol 4. Energy Procedia, Elsevier Science Bv,
Amsterdam, pp 273–281. Doi:10.1016/j.egypro.2011.01.052
95. Yu H, Qi G, Xiang Q, Wang S, Fang M, Yang Q, Wardhaugh L, Feron P (2013) Aqueous
ammonia based post combustion capture: results from pilot plant operation, challenges and
further opportunities. Paper presented at the the 11th Greenhouse Gas Control Technologies
(GHGT) conference, Kyoto International Conference Center, Japan, 18–22 Nov 2012
96. Aleixo M, Prigent M, Gibert A, Porcheron F, Mokbel I, Jose J, Jacquin M (2011) Physical
and chemical properties of DMX (TM) solvents. In: Gale J, Hendriks C, Turkenberg W
66 L.T. Wardhaugh and A. Cousins

(eds) 10th international conference on Greenhouse Gas Control Technologies, vol 4. Energy
Procedia. Elsevier Science Bv, Amsterdam, pp 148–155. Doi:10.1016/j.egypro.2011.01.035
97. Raynal L, Alix P, Bouillon PA, Gomez A, de Nailly ML, Jacquin M, Kittel J, di Lella A,
Mougin P, Trapy J (2011) The DMX (TM) process: an original solution for lowering the cost
of post-combustion carbon capture. In: Gale J, Hendriks C, Turkenberg W (eds) 10th
international conference on Greenhouse Gas Control Technologies, vol 4. Energy Procedia.
Elsevier Science Bv, Amsterdam, pp 779–786. Doi:10.1016/j.egypro.2011.01.119
98. Misiak K, Sanchez CS, van Os P, Goetheer E (2013) Next generation post-combustion
capture: Combined CO2 and SO2 removal. In: Dixon T, Yamaji K (eds) Ghgt-11, vol 37.
Energy Procedia. Elsevier Science Bv, Amsterdam, pp 1150–1159. Doi:10.1016/j.egypro.
2013.05.212
99. Cousins A, Artanto Y, Meuleman E, Pearson P, Puxty G, Jansen J, Conway W, Slater N,
Curtis E, Monch A, Feron P, Verheyen V, Misiak K, Huizinga A, Sanchez Sanchez C, van
Os P, Goetheer E, Castelow P (2014) Combined low-cost pre-treatment of flue gas and
capture of CO2 from brown coal-fired power stations. Paper presented at the Low Rank Coal,
Third International Industry Symposium, Melbourne, Australia, 28 Apr–1 May 2014
100. Yu H, Qi G, Wang S, Morgan S, Allport A, Cottrell A, Do T, McGregor J, Wardhaugh L,
Feron P (2012) Results from trialling aqueous ammonia-based post-combustion capture in a
pilot plant at Munmorah Power Station: gas purity and solid precipitation in the stripper. Int J
Greenhouse Gas Control 10:15–25. Doi:10.1016/j.ijggc.2012.04.014
101. Puxty G, Wei SC-C, Feron P, Meuleman E, Beyad Y, Burns R, Maeder M (2014) A novel
process concept for the capture of CO2 and SO2 using a single solvent and column. In:
Dixon T, Herzog H, Twinning S (eds) 12th international conference on Greenhouse Gas
Control Technologies, Ghgt-12, vol 63. Energy Procedia, pp 703–714. Doi:10.1016/j.
egypro.2014.11.078
102. McLarnon CR, Duncan JL (2009) Testing of ammonia based CO(2) capture with
multi-pollutant control technology. In: Gale J, Herzog H, Braitsch J (eds) Greenhouse
Gas Control Technologies 9, vol 1. Energy Procedia, vol 1, pp 1027–1034. Doi:10.1016/j.
egypro.2009.01.136
103. Manning FS, Thompson RE (1991) Oilfield processing of petroleum, vol 1: Natural gas, vol
1. Penwell Publishing Company, Tulsa
104. Campbell KLS, Lapidot T, Williams DR (2015) Foaming of CO2-loaded amine solvents
degraded thermally under stripper conditions. Ind Eng Chem Res 54(31):7751–7755.
Doi:10.1021/acs.iecr.5b01935
105. Chen X, Freeman SA, Rochelle GT (2010) Foaming of aqueous piperazine and
monoethanolamine for CO2 capture. Int J Greenhouse Gas Control 5(2):381–386. Doi:10.
1016/j.ijggc.2010.09.006
106. Richner G, Puxty G (2012) Assessing the chemical speciation during CO2 absorption by
aqueous amines using in situ FTIR. Ind Eng Chem Res 51(44):14317–14324. Doi:10.1021/
ie302056f
107. Artanto Y, Jansen J, Pearson P, Do T, Cottrell A, Meuleman E, Feron P (2012) Performance
of MEA and amine-blends in the CSIRO PCC pilot plant at Loy Yang Power in Australia.
Fuel 101:264–275. Doi:10.1016/j.fuel.2012.02.023
108. Azzi M, Day S, French S, Halliburton B, Jackson P, Lavrencic S, Riley K, Tibbett A (2010)
CO2 Capture Mongstadt—Project A—Establishing sampling and analytical procedures for
potentially harmful components from post-combustion amine based CO2 capture. Task 2:
Procedures for manual sampling. EP 105456, CSIRO, Australia
109. Gassnova (2015) http://www.gassnova.no/en/ccs-projects/full-scale-mongstad. Accessed
July 2015
110. Bui M, Gunawan I, Verheyen TV, Meuleman E, Feron P (2014) Dynamic operation of
post-combustion CO2 capture in Australian coal-fired power plants. In: Dixon T, Herzog H,
Twinning S (eds) 12th international conference on Greenhouse Gas Control Technologies,
Ghgt-12, vol 63. Energy Procedia, pp 1368–1375. Doi:10.1016/j.egypro.2014.11.146
Process Implications of CO2 Capture Solvent Selection 67

111. Freeman SA (2011) Thermal degradation and oxidation of aqueous piperazine for carbon
dioxide capture. Ph.D. thesis, University of Texas at Austin
112. Notz R, Mangalapally HP, Hasse H (2012) Post combustion CO2 capture by reactive
absorption: Pilot plant description and results of systematic studies with MEA. Int J
Greenhouse Gas Control 6:84–112. Doi:10.1016/j.ijggc.2011.11.004
113. Nielsen PT, Li L, Rochelle GT (2013) Piperazine degradation in pilot plants. In: Dixon T,
Yamaji K (eds) Ghgt-11, vol 37. Energy Procedia, pp 1912–1923. Doi:10.1016/j.egypro.
2013.06.072
Useful Mechanisms, Energy Efficiency
Benefits, and Challenges of Emerging
Innovative Advanced Solvent Based
Capture Processes

Wojciech M. Budzianowski

Abstract Innovative advanced solvent based capture processes (ASBCPs)


employing sophisticated separation mechanisms have emerged as essential solu-
tions that may dramatically reduce high energy requirement of CO2 separation in
capture plants. This study systematically characterises useful mechanisms of several
such ASBCPs, being mostly at relatively low technology readiness levels. Based on
improved understanding of these emerging ASBCPs it is shown how they can
contribute to achieving energy efficiency benefits and thus increase CO2 capture
efficiency. It is followed by an analysis of practical examples of all ASBCPs
recently presented in academic and patent literature. Finally, major challenges of
discussed emerging ASBCPs are identified showing potential directions for further
research and development. Overall, innovative ASBCPs employ very different
mechanisms translating to different energy efficiency benefits in various CO2
capture applications. The provided technological account along with the identified
challenges may be useful for capture process developers which will be able to bring
the most promising ASBCPs to higher technology readiness levels and finally they
can be employed in the practice.

Nomenclature
5OCB 4,4′-pentyloxy cyanobiphenyl
7OCB 4,4′-heptyloxy cyanobiphenyl
AA Aqueous ammonia
AAS Amino acid salt
ASBCP Advanced solvent based capture process
BDA 1,4-butanediamine
BTCA Benzotriazole-5-carboxylic acid
CA Carbonic anhydrase

W.M. Budzianowski (&)


Consulting Services, Wrocław, Poland
e-mail: wojciech.budzianowski@gmail.com
W.M. Budzianowski
Renewable Energy and Sustainable Development (RESD) Group, Wrocław, Poland

© Springer International Publishing AG 2017 69


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_4
70 W.M. Budzianowski

CAP Chilled ammonia process


CAPEX Capital expenditure
CO2BOL CO2 binding organic liquid
DEEA 2-(diethylamino) ethanol
DES Deep eutectic solvent
DMCA N,N-dimethylcyclohexylamine
DMF Dimethylformamide
DPA Dipropylamine
DsBA Di-sec-butylamine
EPD N-ethyl piperidine
GAP-0 1,3-bis(3-aminopropyl)-1,1,3,3-tetramethyldsiloxane
HA Hexylamine
IL Ionic liquid
IPADM-2BOL 1-((1,3-dimethylimidazolidin-2-ylidene)amino)propan-2-ol
MAPA 3-(methylamino)propylamine
MEA Monoethanolamine
MOF Metal organic framework
OPEX Operating expenditure
PC Pulverised coal
PCH3 4,4′-propylcyclohexyl benzonitrile
PCH7 4,4′-heptylcyclohexyl benzonitrile
TETA Triethylenetetramine
TMG 1,1,3,3-tetramethylguanidine
TRL Technology readiness level
TSIL Task specific ionic liquid
%wt Percent by weight

1 Introduction

Gas-liquid absorption is among the main technologies seriously considered for


large scale CO2 capture plants, especially if they are operated in a post-combustion
mode [1]. The technology has been developed in recent few years and is now
becoming one of leading decarbonisation options. In particular developments of
advanced solvent based capture processes (ASBCPs) create opportunities for dra-
matic reduction of the energy penalty associated with removing CO2 from flue
gases. The potential of ASBCPs reaches however far beyond flue gases. ASBCPs
may be used in various sectors from cement production through refineries to
industries processing biomass and various carbonaceous materials. Due to their
versatility ASBCPs have potential for decarbonisation of different CO2 intensive
industries.
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 71

Innovative advanced solvent based capture processes employ various chemical


and/or physical phenomena capable of enhancing a capture process [2] thus min-
imising energy requirements for the entire CO2 capture chain. Since reactive sol-
vents are able to strongly bind with CO2 and thus efficiently separate it from diluted
gases, gas-liquid chemical absorption is more efficient in post-combustion CO2
capture where CO2 is present in relatively low concentrations [3, 4]. CO2 capture by
gas-liquid absorption requires temperatures in between solid sorbent and cryogenic
captures. Such a moderate thermal regime usually alleviates capture energy penalty
due to improved thermal compatibility with flue gas temperatures and thus reduced
process irreversibility. It may further translate to thermodynamic benefits achieved
in energy conversion cycles integrated with CO2 capture. Solid sorbents seem more
suitable for pre-combustion and chemical looping applications while cryogenic
capture cannot be efficiently integrated into thermal power plants [5]. Moreover,
solid sorbents have limited regenerability and often degrade after a few cycles. The
problem with membranes and cryogenic capture is their limited maturity for real
scale applications.
Some ASBCPs may be also highly environmentally benign. For example, by
employing ASBCPs in which the solvent is not in direct physical contact with the
capture installation and flue gases one can minimise solvent degradation, vapori-
sation and corrosion. Since real CO2 capture plants will process huge amounts of
flue gases even minimal volatility would lead to high overall emissions. The
physical separation of the solvent and infrastructure may be achieved by
microencapsulation or solvent membranes. ASBCPs are also suitable for process
intensification. For instance, CO2 absorption is intensified by precipitating CO2
comprising solids from the solution thus counteracting solvent saturation and
maximising the absorption driving force. In addition, thermal regeneration may be
also intensified e.g. by using a polarity swing phenomenon enabling CO2 release
from the solvent at low temperatures.
Figure 1 presents ASBCPs for CO2 capture by gas-liquid absorption investi-
gated in this study. The emphasis within this current study is put on useful

Fig. 1 Advanced solvent based capture processes for CO2 separation by gas-liquid absorption
investigated in this study
72 W.M. Budzianowski

mechanisms and energy efficiency benefits from applying ASBCPs followed by the
identification of challenges that need to be addressed through future research and
developments activities.

2 Precipitating Solvents

2.1 Useful Mechanisms

The useful mechanism employed by precipitating solvents relies on phase change


upon contact with CO2 containing gases resulting in precipitation of solids from the
solvent solution. The precipitate is usually a salt and forms in a scrubber as a CO2-
rich slurry of very small particles. The precipitated slurry, before it is sent to a
stripper for regeneration and CO2 separation, is thickened by filtration or sedi-
mentation. The obtained CO2-lean solvent fraction is recycled to the scrubber for
continuous CO2 separation.

2.2 Energy Efficiency Benefits

By applying precipitating solvents the energy efficiency of the capture process is


improved because the removal of dissolved CO2 by precipitation counteracts sol-
vent saturation and hence maximises CO2 absorption flux. It reduces solvent flow
and thus energy penalty per unit CO2 captured. In addition, the regeneration step
can also be more energy efficient because the precipitate is decomposed at a higher
pressure than liquid CO2-loaded solvents thereby producing pressurised CO2 stream
requiring less energy for compression. Many precipitating solvents are anhydrous
(e.g. GAP-0) and hence they require less energy for solvent regeneration because no
water with high associated vaporisation enthalpy needs to be evaporated. In contrast
to neat reactive solvents with strong CO2/solvent binding (the formed compound
first needs to be evaporated and then thermally decomposed to yield the CO2 gas
and the solvent which is very energy intensive [6]) most precipitating solvents may
release CO2 without boiling the solvent which improves their energy efficiency.

2.3 Practical Examples

Numerous chemicals may serve as precipitating solvents but only a few have
properties that may lead to developing an energy efficient capture process. A GAP-0
aminosilicone solvent is one of such candidates. It is a liquid characterised by a
relatively high boiling point (265 °C) and low viscosity (4 cP at 25 °C). The
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 73

carbonation product is GAP-0 carbamate which relatively rapidly precipitates [7].


By increasing temperature the GAP-0 carbamate is decomposed into CO2 and the
GAP-0 solvent suitable for reuse. GAP-0 is thermally stable up to 180 °C and
poorly volatile. Higher operating temperatures lead to increased CO2 partial pres-
sure and thus the cycling CO2 loading capacity is increased. A non-volatility
property simplifies the CO2 stripping process (there is no need for evaporated
solvent separation) thereby decreasing OPEX and CAPEX. The GAP-0 based
process is able to reduce energy requirements by up to 50% compared to MEA
(with concentration of 30%) [7, 8]. Since GAP-0 is less reactive towards CO2 than
MEA, the energy requirements reduction is greater under concentrated CO2 gas
stream conditions, i.e. the behaviour is similar as for physical vs. chemical solvents
(physical solvents outperform chemical solvents at high CO2 concentrations). The
reduced energy requirement also arises from the fact that the specific heat of MEA
is approximately 60% greater than that of GAP-0 [3.7 vs. 2.3 kJ/(kg K)] which
translates to reduced sensible heat in the stripping process involving solvent
heating. Characteristics of the GAP-0 based precipitating solvent ASBCP are
outlined as follows. First, CO2 is absorbed by using a spray column with little
solvent. It is followed by separating the precipitated GAP-0 carbamate solids from
flue gases by applying a cyclone. Then separated solids are screw conveyed from
the scrubber operated at 0.1 MPa to the stripper operated at 0.5–2 MPa. Further, the
GAP-0 carbamate precipitate is heated by steam injection to temperatures between
100 and 115 °C. Next, CO2 is stripped from the precipitate by applying two agi-
tated vessels where operating pressure is reduced following from the first to the
second vessel. Finally, the regenerated GAP-0 is recycled to the scrubber.
Hydrous ammonia upon contact with CO2 under lower temperatures and higher
concentrations precipitates ammonium salts [9] such as bicarbonate, carbonate,
sesqui-carbonate, and carbamate. The precipitation is achieved in a crystalliser
which separates the CO2-rich and CO2-lean solvent. Compared to hydrous
ammonium without solid formation (chilled ammonia process), 40% lower mass
flow rate need to be sent to the stripper due to CO2 enrichment, thus reducing the
regeneration energy. The CO2-lean solvent is conveyed to bottom section of the
scrubber while full regenerated solvent from the stripper to the top of the scrubber
which optimises process performance [10].
One another chemical used as a precipitating solvent is an amino acid salt
(AAS), e.g. taurine. The AAS solvent on contact with CO2, after reaching the
solubility limit of the zwitterion, precipitates amino acids. The process benefits
from increased pH (due to acid precipitation) which increases the CO2 absorption
rate. It is achieved in a spray scrubber [11] because spray absorption is superior
compared to packed beds due to lower pressure drops and acceptable CO2
absorption rates. The rate of absorption is not limited by stiff solvent droplets
because the viscosity of AAS is low allowing for the fast renewal of droplet surface.
The CO2 loaded solvent forms a slurry comprising about 15% of solids. The slurry
is conveyed to a thermal stripper operated at about 75 °C where concentrated
CO2 stream is obtained. The energy requirement of this ASBCP is about
74 W.M. Budzianowski

2400 kJ/kgCO2, however this number does not include low-grade heat consumed in
order to dissolve the precipitated solids [12]. A DECAB Plus process employing
K-TAU amino acid was claimed to reduce energy requirement by 65% compared to
MEA benchmark (concentration 30%) [13]. However, for the CO2 capture applying
another AAS (KSAR with concentration of 5 m) the recorded energy efficiency
improvement was marginal [14]. The reason might be that not only solvent but also
plant layout affects energy requirements thus an inefficient plant tends to increase
overall energy requirements. Thus using energy efficiency techniques such as flash
and lean vapour compression a further decrease of the energy requirement can be
obtained, e.g. to 1900 kJ/kgCO2 as shown in [15].
A precipitating solvent based ASBCP may also include porous powder suspended
in glycol. This system is exceptional in that it does not chemically react the CO2 and
the solvent but the CO2 is only physically entrapped in the pores. Such a combined
ASBCP makes use of the suitability of liquids for operation in large capture plants
and potentially reduced energy requirements of solid sorbents since the system
contains suspended solids. The system may consist of 2-methylimidazole glycol as a
liquid and ZIF-8 (MOF) as a powder. The porosity characteristics of ZIF-8 is such
that its pores of 3.4 Å are too small for entrapping the glycol (having a molecular size
of 4.5 Å) while are sufficient for entrapping CO2 molecules. ZIF-8 is very well
soluble in glycol and is characterised by chemical and thermal stability. CO2 can be
stripped from the powder without evaporating the glycol leading to alleviated energy
requirement. The whole system is therefore potentially suitable for energy efficient
operation in multiple cycles.
Oligomers, i.e. polymers highly soluble in solvents, which on contact with CO2
precipitate solids are another class of materials for use as precipitating solvents.
Certain amino silicones in combination with anhydrous glycol have 50% higher
CO2 loading capacity compared to MEA (concentration 30%) [16]. Due to pre-
cipitation and high loading capacity such an ASBCP achieves high CO2 absorption
rate [17]. Nevertheless, new material candidates suitable for use in realistic capture
conditions would have to be proposed or synthesised to make oligomers more
important for capturing CO2.
K2CO3 is a cheap and potentially promising precipitating solvent. It crystallises
on contact with CO2 forming KHCO3 and subsequent crystal regeneration is
obtained by applying increased pressure. The separation of KHCO3 is obtained by
means of a hydrocyclone and the slurry is sent to the regenerator. The reported
energy consumption is approximately 2500 kJ/kgCO2 [18]. In realistic capture
applications 40 wt% K2CO3 + 10 wt% potassium glycine is applied, the latter
acting as a promoter because it can enhance the CO2 regeneration rate by 6
fold [19].
The UNO MK 3 process employs a K2CO3 solvent and uses a hydrocyclone for
rich solvent thickening as well as recycles exhaust gases to obtain a CO2-enriched
stream. It achieves reduced energy requirement of about 3000 kJ/kgCO2 [20].
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 75

2.4 Challenges

The use of precipitating solvents in real capture conditions meets is however


associated with several challenges. For instance the presence of non-CO2 gases
such as SO2 and NOX in flue gases [21, 22] cause co- precipitation and often require
different conditions for removal than CO2 thus complicating the regeneration
step. If non-CO2 gases release from the precipitate they decrease CO2 stream purity.
If they accumulate in the solvent, e.g. as solids, CO2 loading capacity is decreased
over time. Accumulation of non-CO2 gases occurs if they bind stronger than CO2
with the solvent and cannot be fully separated in the regenerator. Another challenge
typical of slurry systems is packed bed clogging by formed solids. The clogging
increases column pressure drop thereby rising OPEX of the plant. The precipitate
build-up on the packing may be slow but over time may fully clog the scrubber.
Consequently, spray columns and wetted wall columns need to be considered
because they do not clog. However, these contactors suffer from lower CO2 capture
rates due to very low gas-liquid contact area and could be inefficient in large capture
plants. For packed beds which usually achieve greater absorption rates new can-
didates for precipitating solvents would have to be implemented in order to over-
come challenges associated with clogging.

3 Two Immiscible Liquid Phases

3.1 Useful Mechanisms

The property of two immiscible liquid phases may be used to create an effective
ASBCP. For an effective system immiscibility of the involved liquids needs to be
temperature dependent. More specifically, under scrubbing temperatures these
solvents must be homogenous while under stripping temperatures must be
heterogeneous. It is schematically displayed in the panel a of Fig. 2. Figure 2 in
panel b displays two immiscible liquid phases with upper critical temperature which
will form two phases in a scrubber if operated at lower temperature while panel c
shows a solvent with two critical temperatures which would from two phases at
moderate temperature. All present day two immiscible liquid phases ASBCPs are
based on the mechanisms given in Fig. 2 panel a.
An interesting ASBCP could be created if the immiscibility of involved liquids
was CO2 concentration dependent. In this case a sufficient trigger to form two
phases would be varied CO2 concentration and no heating would be need for
solvent regeneration. However, again no such solvents have been found so far.
In an ASBCP involving two immiscible liquid phases with lower critical tem-
perature (Fig. 2 panel a) the two resulting liquid phases vary in terms of CO2
solubility and CO2 accumulates only in one phase. From the two phases formed
(CO2-lean and CO2-rich) only the latter is conveyed to the stripper.
76 W.M. Budzianowski

Fig. 2 Phase behaviours of ASBCPs relying on two immiscible liquid phases. Panel a—lower
critical temperature, panel b—upper critical temperature, panel c—upper and lower critical
temperatures. Only panel a represents the temperature dependent case that has been employed in
the practice for capturing CO2

For some already tested two immiscible liquid phases based ASBCPs the
mechanism of their temperature dependent immiscibility is associated with the roles
played by hydrophobic and hydrophilic functional groups comprised by lipophilic
amines. Lipophilic amines have highly varied solubility in water which sharply
decreases at higher temperatures leading to phase splitting. The known solvents
split upon heating to about 80 °C [23].
The whole solvent as a single phase is used to scrub CO2. The CO2 loaded
mixture is conveyed to a coalescence tank where two phases are formed. The CO2-
rich phase is sent to the stripper while the CO2-lean phase is recycled back to the
scrubber, after mixing with the regenerated second phase.

3.2 Energy Efficiency Benefits

The reason for energy efficiency benefits is that the volume of the regenerated CO2-
rich phase is much smaller than that of the CO2-lean phase. In anhydrous systems
less energy is required for solvent heating and vaporisation (water has high sensible
heat and heat of vaporisation [24]). As both involved liquids contribute to CO2
absorption in a scrubber thus mass transfer is enhanced and less solvent needs to be
cycled between a scrubber and stripper. The CO2-rich phase has very high CO2
concentration and thanks to its lower volume the ASBCP requires very little solvent
meaning also less required energy.
The demonstration obtained in the iCAP project with a DEEA + MAPA based
ASBCP revealed that the energy requirement of two immiscible liquid phases is
about 2400 kJ/kgCO2 [25]. Similar results have been obtained by a DMX process
showing energy consumption of about 2300 kJ/kgCO2 [26]. Reduced stripping
temperature in this ASBCP facilitates the utilisation of low-grade heat for solvent
recovery.
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 77

3.3 Practical Examples

Several liquid pairs have been proposed for capture applications. For instance,
water and lipophilic amines such as HA, DMCA, DPA, DsBA, and EPD. Water
forms a CO2-lean phase while lipophilic amines a CO2-rich phase. The CO2 loading
capacity of water is marginal under practical capture operating conditions (it
depends on pH and pressure) while for lipophilic amines it is approximately 1 kmol
CO2/kmol lipophilic amine) [27]. These liquids are immiscible and CO2 concen-
trates in lipophilic amines which only need separation and regeneration.
DEEA and MAPA liquids that form two immiscible phases were tested within
the iCAP project. MAPA consisting of one primary and one secondary amine
groups formed the heavier CO2-rich phase [25, 28]. DEEA being a tertiary alka-
nolamine formed the lighter CO2-lean phase. The DEEA + MAPA system operated
at high pressure [25] which allowed to produce a pressurised CO2 stream. More
recently the 5 M DEEA/2 M MAPA mixture was investigated in the Gløshaugen
plant [29] achieving lower reboiler temperatures and duty (compared to MEA at
30% concentration).
Xu et al. [30] explored the DEEA-BDA mixture having similar characteristics as
the DEEA-MAPA but its validation under real flue gas conditions and scales are to
be demonstrated.
Ye et al. tested the TETA + DEEA blend (1:4) and found overall energy
requirement reduction of 30% compared to MEA (concentration 30%) [31].

3.4 Challenges

Challenges include the need for unconventional solvent handling, e.g. through
separation of two liquid phases in a coalescence tank. In addition, the selection of
immiscible liquid pairs is complex and screening must be very detailed. No solvents
exist representing temperature dependent cases displayed in Fig. 2 panels b, c. Also
CO2 concentration dependent immiscibility has not been explored so far.
Challenges also include solvent degradation and volatility. Especially volatility
limits solvent selection opportunities and control to avoid solvent losses adds to the
plant’s OPEX [29].

4 Catalysed Solvents

4.1 Useful Mechanisms

The rate of CO2 binding reaction may be increased by catalysts. A very fast natural
catalyst is an enzyme carbonic anhydrase (CA). CA impacts primarily the rate of
78 W.M. Budzianowski

reaction (1)—CO2 hydration/HCO3− dehydration [32]—which is the real bottle-


neck of many hydrous reactive solvent based CO2 capturing processes:

CO2 þ H2 O
HCO
3 þH
þ
ð1Þ

Reaction (1) is reactant diffusion controlled under low concentrations (catalysis


is thus not helpful). However, at higher CO2 concentrations typical in CO2 capture
reactants are abundantly available and CA efficiently catalyses the CO2 hydration to
bicarbonate. The turnover rate of is up to 1.5  106 (mol CO2/(s mol CA) [33]. CA
may be added to the solvent due to its water solubility but preferably it should be
immobilised, e.g. on column packing to prevent washing to the high-temperature
stripper where it may become denatured. However, if solvent regeneration is
achieved at low temperatures (below 40–50 °C) CA may be employed to catalyse
also the dehydration of CO2 by bicarbonate decomposition. The impact of CA
catalysis is most pronounced for solvents with low reactivity such as K2CO3 while
solvents with high reactivity such as MEA CA may give marginal benefits [34].

4.2 Energy Efficiency Benefits

Major energy efficiency benefits arising from the use of CA are associated with
increased CO2 reaction rates. This translates to compact designs and reduced sol-
vent circulation leading to reduction in energy penalties. If CO2 stripping could be
catalysed this could also potentially lead to a regeneration temperature reduction.
CA may be also applied in various hybrid processes such membrane immobilisation
and hence greater energy efficiency of these hybrids could be achieved.

4.3 Practical Examples

Due to varying catalytic mechanisms under different operating conditions the cat-
alytic effect of CA depends on many parameters. For poorly reactive solvents such
as K2CO3 benefits are usually significant. CA performance is improved if immo-
bilised on the ceramic packing which minimises wash-out. CO2 absorption rate is
increased by 20-fold. Typical energy requirement is 3500 kJ/kgCO2 [35]. For CA
catalysed K2CO3 solvent flue gas flow rate may be increased by 7-fold compared to
blank K2CO3. This reduces column size and CAPEX. In demonstrations by
Akermin [35] some CA related problems have been overcome. For example, the
process was tested on coal combustion derived flue gas achieving steady operation
during 2800 h. Nevertheless, CA inactivation is still a problem since 54% CA
deactivation has been observed after 1600 h of operation.
In [36] a K2CO3 solvent promoted by a Zn complex containing a cyclic ligand of
1,4,7,10-tetracyclodode cane (Zn-cyclen) was used by applying a membrane
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 79

contactor. Zn-cyclen catalyst is more stable than CA, has a longer life time and its
molecular weight (235 g/mol) is smaller than that of CA. This study found that in
the Zn-cycle catalysed K2CO3 solvent the rate constant of CO2 hydration reaction
was 10 fold increased compared to the non-catalysed solvent.
Other catalysts such as oxoanions (e.g. phosphate, phosphite, arsenate, and
arsenite) have been also proposed [37]. Although significant catalytic effects were
noted these cheap materials require further research addressing toxicity and
chemical stability of oxoanions.

4.4 Challenges

Challenges in using CA is mainly associated with the susceptibility of CA to


denaturation at temperatures higher than 40–50 °C. It limits operating temperatures
especially in the striper or requires CA separation from the solvent or immobili-
sation in the scrubber. Handling of CA through immobilisation or filtration (tested
in HiPerCap [38]) adds complexity to the whole process. Besides, due to high liquid
phase mass transfer resistance CA is only effective when acting precisely at the
gas-liquid interface which necessitates the use of methods such as immobilisation
that expose the CA for more direct gas contact [39].

5 Microencapsulated Solvents

5.1 Useful Mechanisms

In microencapsulated solvents solvent is entrapped within microcapsules. These


microcapsules have highly CO2 permeable shells and the solvent is present in their
cores. Figure 3 illustrates the mechanism of CO2 scrubbing and stripping by such

Fig. 3 The mechanistic


scheme of CO2 scrubbing and
stripping involving
microencapsulated solvents
80 W.M. Budzianowski

microcapsules. The scrubbing process starts with the diffusion of gaseous CO2
through the permeable shell which is followed by the chemical reaction between
CO2 and the inner solvent comprising hydroxyl ions yielding bicarbonates. The
stripping process involves heating of the microcapsules in order to decompose
bicarbonate and the released CO2 permeates through the shell. Consequently, by
physically separating scrubbing and stripping units a relatively pure CO2 stream is
obtained.

5.2 Energy Efficiency Benefits

Main energy efficiency benefits are associated with higher contact area meaning that
less solvent is required and less pumping power is consumed. Solvent microen-
capsulation beneficially separates the solvent from the CO2 capture installation and
flue gases which limits solvent degradation and infrastructure corrosion. Potentially
less energy may be thus required in order to minimise unwanted emissions from
solvent degradation and vaporisation by minimising downstream processing.
Solvent vaporisation control by various additives [40–44] is thus no longer required
with microencapsulated solvents minimising OPEX. The microcapsule structures
enable to exploit solvents more efficiently. They make use of the loading capacity
and selectivity of reactive solvents. Due to small dimensions of microcapsules the
surface area is very high which enhance both scrubbing and stripping steps.
Microencapsulation allows for the use of most advanced solvents including ILs
(their viscosity is no longer a problem due to microencapsulation) or precipitating
solvents (solids may precipitate and decompose within microcapsules [45]. This all
suggests that microencapsulating may be a good solution towards developing truly
energy efficient ASBCPs.

5.3 Practical Examples

Microencapsulated solvents as a relatively new ASBCP have limited demonstra-


tion. From what is available in the open literature it may be concluded that
microcapsules may be produced by applying a microfluidic double-capillary device
[46]. Typical microcapsules diameters are between 0.1 and 0.6 mm. Microcapsules
consist of a few layers. The microcapsule core consists of the solvent (hydrous
K2CO3 or Na2CO3). The rate of CO2 reaction with the solvent is increased by some
25 % by a catalyst (cyclen—Zn-1,4,7,10-tetraazacyclododecanen) placed in the
microcapsule interior core. The overall increase of CO2 absorption rate is more than
5 fold compared to neat solvents due to contributions from catalysis and surface
expansion. The external microcapsule layer consists of silicone material highly
permeable for CO2 gas.
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 81

Flue gases may be passed through a fluidised bed of microcapsules to absorb


CO2. CO2 loaded microcapsules are continuously removed and regenerated at an
elevated temperature [47].

5.4 Challenges

Challenges associated with applying microcapsules especially in realistic capture


conditions are associated with mass transfer resistance across the capsule shell. The
shell may be contaminated by various constituents of flue gases reducing gas
permeability. The contaminants will likely fill the pores of silicone shells reducing
the lifetime of microcapsules. Microencapsulated solvents, in contrast to solvent
membranes, need to be circulated between the scrubber and the stripper which may
decrease lifetime of potentially fragile microcapsules. If direct contact with flue
gases is considered, microcapsules need to be resistant to impurities such as SO2 or
NOX.

6 Solvent Membranes

6.1 Useful Mechanisms

A solvent membrane ASBCP includes a porous membrane separating a solvent film


from flue/sweep gases, Fig. 4. The porous membrane is in contact with pressurised
flue gases enabling CO2 to permeate through the pores and react with the solvent in
the inner solvent membrane. The CO2 loaded solvent molecules diffuse through the
solvent membrane, decompose to release CO2 which passes the porous membrane
and leaves the contactor as the CO2-rich permeate. The CO2 binding reaction needs
to be rapid to intensify CO2 absorption rates by removing free CO2 from the porous
membrane. The solvent membrane ASBCP may beneficially use volatile solvents

Fig. 4 Schematic of the


solvent membrane ASBCP
82 W.M. Budzianowski

such as ammonia [48]. It is unsuitable for the use of precipitating solvents since it is
difficult to deliver thermal energy to the solvent membrane in order to decompose
precipitated solids.

6.2 Energy Efficiency Benefits

This kind of ASBCP enables the use of more expensive solvents being highly
reactive and selective towards CO2. The high reactivity and selectivity will minimise
the need for pressure difference across the membrane and reduce associated energy
requirements. Since the CO2 rich permeate is produced directly from the solvent
membrane contactor, no additional thermal energy is needed to strip the CO2 from
the loaded solvent which potentially reduces the overall energy requirement. Energy
efficiency benefits are also gained by means of more compact scrubber design due to
higher gas-liquid contact area [49]. Since the solvent is used as a solvent membrane
virtually no solvent circulation is needed minimising associated pumping work.
Since the solvent is separated from the gas phases by the porous membranes it does
not vaporise and is kept clean which minimise solvent losses translating to lower life
cycle energy needs associated with solvent production.

6.3 Practical Examples

Many solvent membrane ASBCPs have been investigated over recent years.
Solvents implemented as CO2 capturing liquids inserted in membrane shells include
amine acid salt [50], aqueous ammonia [51, 52] and ionic liquids.
Some solvents achieve benefits while applied in solvent membrane contactors.
For example, a shortcoming of ammonia vaporisation [53] can be overcome due to
separating the solvent and flue gases by an additional membrane barrier.
Ionic liquids such as [emim][Tf2 N], [emim][CF3SO3], [emim][dca] and [thtdp]
[Cl] were tested and CO2 permeabilities in the range of 350–1000 barrers CO2/N2 and
selectivities of 15–61 were reported [54]. The research on fluoroalkyl-functionalised
imidazolium-based solvent membranes revealed CO2 permeabilities of 210–320
and selectivities of 16–27 (CO2/N2) and 13–19 (CO2/CH4) [55]. Unconventional
IL-based solvent membrane tested in [56] showed selectivities 10–52 (CO2/N2), 5–13
(CO2/H2) and 5–23 (CO2/CH4). Tests on membrane systems involving two mem-
brane supports with different hydrophobicity and imidazolium-based ILs as a liquid
membrane revealed mixed-gas selectivities 20–32 (CO2/N2) and 98–200 (CO2/CH4).
The hydrophobic membrane support was found to be more stable than the hydrophilic
one [57]. Selectivities reported in [58] were 21.2 (CO2/N2) and 27 (CO2/CH4) for
[emim][Tf2 N] and [emim][BF4] ILs, respectively. In addition amine-functionalised
IL in a cross-linked Nylon 66 support were analysed [59] and selectivities greater than
15 and CO2 permeabilities ranging from 100 to 1000 barrers were obtained. With
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 83

IL-based solvent membranes in a Nafion matrix selectivity was 26 (CO2/CH4) [60].


Solvent membranes employing ammonium, imidazolium, pyridinium, pyrrolidinium
and phosphonium revealed selectivity of 5–30 (CO2/CH4) [61]. Also amine-
functionalised IL-based solvent membranes enabled to achieve high selectivities [62]
for CO2/CH4 and stable operation.

6.4 Challenges

The required pressure gradient across the solvent membrane is a significant con-
tributor to OPEX and needs to be minimised. Also higher membrane selectivity will
increase the purity of the obtained CO2 stream. Selectivity may be improved by
applying solvent reactive only toward CO2 and exhibiting very low diffusivity of
non-reacted species. The solvent membranes may have suitable applications in
natural gas processing while for flue gases impurities may cause fouling and may
accumulate in solvent membrane. For high CO2 content gas, multiple membranes
may be required which are quite challenging to be implemented.

7 Ionic Liquids

7.1 Useful Mechanisms

Ionic liquids (ILs) may consist of either organic or inorganic salts that make use of
both chemical and physical absorption. ILs reversibly react with CO2 and also have
high physical CO2 loading capacity. The chemical CO2 loading capacity depends on
the involved anions that attach CO2 [63, 64]. The physical CO2 loading capacity is
mainly affected by free volume, size of an IL molecule and the IL’s chemical structure.
The CO2 loaded IL is regenerated at slightly increased temperature enabling the
release of physically as well as chemically bound CO2. They are recyclable and
thermally stable.

7.2 Energy Efficiency Benefits

Energy efficiency benefits are associated with high CO2 loading capacity of ILs which
would reduce the flow rate of circulating liquid. In addition fast reaction rate exhibited
by e.g. polyionic ILs has potential to further reduce required pumping work for solvent
circulation. Besides, ILs have very low vapour pressures meaning that no (energy
intensive) measures need to be applied to minimise solvent vaporisation. ILs are much
less volatile than conventional solvents, which make them an energy-efficient solvent
system. If ILs offer better selectivity, it can be more efficient.
84 W.M. Budzianowski

7.3 Practical Examples

ILs are viscous they are characterised by several times lower liquid mass transfer
coefficients compared to MEA in the same conditions [65] which limits practical
applications of ILs. Therefore, ILs need specific approaches which reduce existing
liquid phase mass transfer limitations such as solvent membranes [66, 67].

7.4 Challenges

Challenges of IL solvents primarily involve high production cost due to complex


chemical structures as well as high viscosity. The high viscosity is responsible for
high pumping cost and mass transfer reduction due to slow diffusion of CO2 in
ionic liquids. If ILs are in direct contact with flue gases they are quickly degraded
which necessitates make-up of the expensive solvent. These are main obstacles for
the use of ILs as neat solvents in large scale capture plants. It seems that ILs would
have to be used only in the context of ASBCPs.

8 Polarity-Swing-Assisted Solvents

8.1 Useful Mechanisms

Polarity-swing-assisted solvents use organic liquids with switchable polarity. These


materials, e.g. IPADM-2BOL, bind CO2 and the shift in polarity allows for more
efficient CO2 stripping. The mechanism of switchable polarity relies on the use of a
non-polar “antisolvent”, e.g. hexadecane. This additive is capable of destabilising
the loaded solvent facilitating CO2 release. The stripping step can therefore be
obtained under temperatures lower than required by neat solvents, e.g. 50–80 °C
[68–70].

8.2 Energy Efficiency Benefits

Due to a swing in solvent polarity the polarity-swing-assisted solvents may achieve


alleviated energy requirement up to 65% compared to MEA benchmark (with
concentration 30%) [71, 72]. Due to reversible CO2 binding [73] the stripping step
is achieved through applying only modest heating or even bubbling with inert gas
[74]. The thermodynamic properties of solvents are very important. For example,
the 1,1,3,3-tetramethylguanidine (TMG) solvent is characterised by slightly lower
reaction rate compared to MEA but has lower activation energy meaning that the
process may be feasible at lower temperatures [75, 76]. If TMG is applied with
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 85

reaction promoters such as piperazine [77] the limited kinetics problem can be
overcome. Overall, the process may be energy efficient. This expectation was
confirmed by investigating the DBU/1-propanol system involving heptane as
“antisolvent” [78]. This system achieved at 75 °C similar stripping rates and CO2
loadings as aqueous amines at 120 °C. Due to lower temperatures solvent degra-
dation and evaporation were less pronounced [71, 72].

8.3 Practical Examples

One example of a polarity swing-assisted ASBCP is the IPADM-2BOL system


comprising decane as “antisolvent”. The bench scale testing during the 4 months
campaign of this ASBCP were described by Zheng et al. [79]. The process favourable
characteristics included little effect of high solvent viscosities on mass transfer rates,
no foaming, minimum solvent evaporation, no measurable solvent degradation. In
addition, by applying a simple coalescence tank antisolvent was separated from the
solvent by phase splitting. Antisolvent carryover was minimal. Water contamination
up to 5 % did not degrade anhydrous solvent performance which suggested that this
ASBCP could be employed for CO2 capturing from humid flue gases.

8.4 Challenges

These ASBCPs have relatively low TRL and sufficient technical characteristics
under realistic capture conditions are not available. Materials selection is the main
challenge, especially for large scale applications. The already demonstrated sol-
vents required relatively high circulation rates implying the need for larger column
sizing. High solvent viscosity implied increased pumping work [79]. In order to
fully benefit from the polarity swing-assisted property further fundamental studies
are therefore required.

9 Slurry Solvents

9.1 Useful Mechanisms

Small solid particles such as metal organic frameworks (MOFs) having remarkable
capacity for CO2 sorption may be mixed with an appropriate liquid carrier to obtain
a slurry with properties of a fluid and a solid [80], see Fig. 5. NMR measurements
from [80] shows that the tested MOF significantly enhances bicarbonate/carbonate
formation contributing to favourable CO2 loading capacity of the formed slurry.
86 W.M. Budzianowski

Fig. 5 The mechanism of


CO2 scrubbing and stripping
by slurry solvents

Thus slurries make benefit from high surface area of porous materials retaining
benefits of liquid solvents.

9.2 Energy Efficiency Benefits

Energy efficiency benefits mainly relate to the ease of solvent handling in a capture
plant while benefiting from solid properties of the slurry. In addition, if efficient
solid particles could be found, the resulting slurry solvent based ASBCP would
consume little energy for solvent pumping while capture CO2 efficiently. The ease
of slurry solvent recovery will also be important. They usually can be regenerated at
low temperatures thus achieving reduced energy requirement compared to 30%
aqueous MEA benchmark. MOFs can be usually regenerated below the boiling
temperature of typical solvents thus reducing energy requirement.

9.3 Practical Examples

One of practical examples includes the metal organic framework complex {Zn[N
[CH2(2-py)]3](l-OH)}2(NO3)2. It achieves CO2 loading capacity of around 2
molCO2/molMOF. However, due to its more than 8 times higher molecular weight
it underperforms MEA solvents in terms of CO2 loading capacities on a mass basis.
It has lower regeneration temperature compared to MEA, 80 °C is sufficient to
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 87

achieve complete regeneration including no carbamate accumulation which is a


problem for MEA solutions at this temperature. If this MOF is dissolved in water
the regeneration at 80 °C means operation below the solvent boiling temperature
reducing energy requirement in stripping. The kinetics of CO2 absorption is similar
to MEA [80].
In [81] a flexible and porous MOF consisting of Co3(OH)2(BTCA)2 with DMF
as guest molecules exhibiting good CO2 uptake (223.7 mg/g at 273 K and
104.7 mg/g at 298 K) and better selectivity for CO2/N2 (up to 80) was investigated.
It could be thus used as a slurry solvent.

9.4 Challenges

Main challenges are associated with difficulties to find small solid particles with
sufficiently high CO2 loading capacity on a mass basis. These molecules are usually
characterised by high molecular mass and hence even if have high CO2 loading
capacity on a molar basis, their loading capacity on mass basis can be small relative
to common solvents such as MEA. Also the stripping step of some slurry solvents
may require harsh conditions increasing the overall energy requirement of the
process.

10 Liquid Crystals

10.1 Useful Mechanisms

Liquid crystals are molecules that form a structured crystalline phase and liquid
phase. The former is referred to as a nematic phase and the latter as isotropic phase.
The shift between the two phases is possible upon a trigger such as a (preferably
small) temperature change. The isotropic phase is formed at higher temperatures.
CO2 capture makes use of a difference between free volume of the solvent between
the isotropic (higher free volume) and nematic (lower free volume) phases [82]. The
nematic phase is characterised by orientational ordering but no positional ordering.
The solubility of CO2 is higher in the isotropic phase than in the nematic phase
making liquid crystals suitable for CO2 scrubbing and stripping cycles, see Fig. 6.

10.2 Energy Efficiency Benefits

Heating the liquid crystal nematic phase just a few degrees is already sufficient to
absorbed CO2. Given the very small temperature trigger this kind of ASBCP has
potential to consume much less energy than required in conventional absorptions.
88 W.M. Budzianowski

Fig. 6 Mechanisms of the


operation of liquid crystals
based ASBCPs

Both phases have similar viscosity [83] meaning that work requirement for solvent
pumping may be limited.
Scrubbing/stripping cycles are different with liquid crystals than with conven-
tional solvents since CO2 release is obtained at a lower temperature than CO2
scrubbing. This property may have implications for energy efficiency benefits
however no work has addressed this process opportunity so far. Theoretically flue
gases could be used to heat the scrubbing process while CO2 stripping could be
subsequently achieved by cooling the solvent to at ambient temperature meaning no
energy requirements for regeneration.

10.3 Practical Examples

Binary and ternary mixtures of chemicals forming liquid crystals are considered for
CO2 capture applications. Researches carried out by de Groen et al. [83] show that
the combination of PCH3 and PCH7 is more promising compared to 5OCB and
7OCB. The former binary system has higher phase transition enthalpy and asso-
ciated CO2 concentration difference, however the latter has higher CO2 loading
capacity. Further researches are required to find improved liquid crystals.
Simultaneous experimental, theoretical and molecular simulation work is needed.
PCH5 liquid crystals have CO2 loading capacity of about 5 %wt and scrubbing
is obtained at 40 °C. The nematic phase is formed at 25 °C. CO2 stripping from the
nematic phase requires enhancement, e.g. by bubbling with an inert gas [84] which
complicates process feasibility.
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 89

10.4 Challenges

Major challenges are associated with low TRL and insufficient demonstration. It
means that potential shortcomings has not yet been revealed.
Among confirmed challenges is the low phase transition enthalpy (from nematic
to isotropic) for tested liquid crystals. This leads to a small solubility difference of
absorbed CO2 in the two phases. The small concentration difference may limit the
effectiveness of scrubbing/stripping cycles [83].

11 Deep Eutectic Solvents

11.1 Useful Mechanisms

Deep Eutectic Solvents (DESs) consist of a salt and a second compound as H2


donor forming a low melting point mixture. DESs have low melting temperatures
due intermolecular H2 bonds [85]. Due to charge delocalization in DESs the melting
points are even lower than that of individual components [86–88]. The H2 bonds
play a remarkable role in more efficient binding of CO2 molecules by DESs
compared to classical ILs.

11.2 Energy Efficiency Benefits

Energy efficiency benefits are similar to those of ILs and largely depend on the
specific DESs used. The role of a H2 donor molecules is essential in minimising
energy needs in scrubbing/stripping cycles [89]. More process level studies are
needed to more clearly demonstrate potential energy efficiency benefits associated
with DESs under practical CO2 capture conditions.

11.3 Practical Examples

Practical examples of DESs include choline chloride salt and urea in molar ratio of
1:2 [90]. Choline chloride salt may also form DESs with carboxylic acids and
polyols [85, 91]. Choline chloride may form DESs with levulinic acid/furfuryl
alcohol for which detailed CO2 loading capacity data can be found in [92]. Several
potential DESs have been reviewed by [93] including multiple halide slats and
hydrogen bond donors but little process level information has been provided.
90 W.M. Budzianowski

Table 1 A summary of useful mechanisms, energy efficiency benefits and challenges of


investigated ASBCPs
Advanced solvent Useful mechanisms Energy efficiency Challenges
based capture process benefits
Precipitating solvents Solids precipitate Practically infinite Complex slurry
upon contact with CO2 loading handling required
CO2. The capacity due to Reactants are
precipitated slurry is precipitation required
thickened and Stable and fast
conveyed to a absorption rate
stripper associated with rapid
precipitation of CO2
loaded compounds
Two immiscible liquid Temperature CO2 accumulation in Unconventional
phases triggered formation one phase reduces solvent handling is
of two immiscible energy requirement required involving a
liquid phases with in solvent pumping coalescence tank
one low-volume and regeneration Very specific
high CO2 Low regeneration chemicals are
concentration phase temperature enables required to fit the
is conveyed to a to use low-grade operating window of
stripper thermal energy CO2 capture
Catalysed solvents Carbonic anhydrase Reaction rate A CA based ASBCP
is used to catalyse increase by up to 10 requires
CO2 reaction with fold for high CO2 temperatures below
water concentrations 40–50 °C to prohibit
protein denaturation
CA handling is
complex, e.g. it
requires
immobilisation on
packing materials or
filtration from the
spent solvent
Microencapsulated The solvent is High surface area Contaminated flue
solvents encapsulated in facilitates mass gases may gradually
microcapsules which transfer block the pores of
separates the solvent Solvents have the silicone shells
from flue gases and prolonged lifetime May require
plant infrastructure and more expensive circulation of
solvents may be used microcapsules
Can handle between the scrubber
precipitating and stripper may
solvents reduce their lifetime
Solvent membranes The porous Capable of Innovative
membrane separates minimising solvent membrane materials
the solvent from flue volatility facilitating required
gases. Flue gases the utilisation of The required
(optionally volatile solvents pressure gradient
pressurised) are such as ammonia
(continued)
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 91

Table 1 (continued)
Advanced solvent Useful mechanisms Energy efficiency Challenges
based capture process benefits
conveyed to the The size of the across the solvent
channel comprising absorption increases OPEX
the porous equipment reduced Membrane
membrane. CO2 due to high surface wettability and
permeates through to volume ratio fouling
the pores and reacts Reactive liquids
with the solvent in enable achieving
the inner space. The very high
CO2 loaded solvent selectivities
molecules diffuse Due to no direct
through the solvent contact between a
membrane to the solvent and flue
sweep gas side, gases, more
decompose to release expensive solvent
CO2 which leaves materials may be
the system as a employed, e.g. ILs or
permeate stream catalysed solvents
Ionic liquids Chemical and High CO2 loading High solvent
physical absorption capacity due production cost limit
of CO2 physicochemical application only to
nature of absorption ASBCPs (too costly
Completely for use as a neat
reversible reaction solvent)
with CO2 facilitates
stripping
Polarity-swing-assisted Polarity shift due to No thermal Difficulties in finding
solvents “antisolvent” regeneration reduces suitable polarity
addition chemically overall energy swing materials for
destabilises the CO2 requirement realistic CO2 capture
loaded solvent and conditions
facilitates stripping Very low TRL and
lacking technical
details
Slurry solvents Small solid particles Fluid nature of the Small solid particles
mixed with a solvent slurry alleviates with sufficiently high
may be operated like pumping work CO2 loading
a fluid adding to MOFs can be capacity on a mass
operational regenerated below basis needs to be
flexibility of capture the boiling found
plants temperature of
typical solvents thus
reducing energy
requirement
Liquid crystals Difference in CO2 Small temperature Very small CO2
solubility between trigger makes solubility difference
isotropic and possible low heat between the two
nematic phases that demand phases leads to
(continued)
92 W.M. Budzianowski

Table 1 (continued)
Advanced solvent Useful mechanisms Energy efficiency Challenges
based capture process benefits
shift upon a (small) Similar (possibly inefficiencies of
temperature low) viscosity of multiple
difference both phases reduces scrubbing/stripping
solvent pumping cycles
work
If scrubbing could
use heat of flue gases
and stripping could
be achieved at
ambient temperature,
the process would
require nearly no
energy for solvent
regeneration but
such systems have
not been found
Deep eutectic solvents Consist of a salt and Released energies Require high
a second compound for CO2 capture are pressures and low
acting as hydrogen larger than those of temperatures in
donor forming a low classical ILs absorption
melting mixture
The H2 bonds play a
remarkable role in
more efficient
binding of CO2
molecules by DESs
compared to
classical ILs

11.4 Challenges

From scarce literature information it is seen that in order to achieve meaningful CO2
loading capacity low temperatures and high pressures need to be applied. For
example, for DES consisting of choline chloride and glycerol in a 1:2 molar ratio at
313 K CO2 loading capacity is 1.5 molCO2/kgDES at 30 bar and only about 0.15
molCO2/kgDES at 3 bar. It suggests that DESs operate like physical solvents and in
the practice would require pressurised absorption in which high energy requirement
is associated with flue gas compression if post combustion is to be used. Since there
is almost no potential to reduce compression work (compression already approa-
ches thermodynamic limits) little progress toward energy efficient CO2 capture with
DESs requiring flue gas pressurisation can be expected. In general state-of-the-art
DESs based ASBCPs lack demonstration activities that could provide reliable
process related information for CO2 capture under realistic conditions.
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 93

12 Summary

Table 1 provides useful mechanisms, energy efficiency benefits and challenges of


all discussed ASBCPs.
The investigated ASBCPs are at different TRL levels. The ASBCPs that are at
the lowest TRL require further fundamental studies to expound useful mechanisms
and tune operating parameters [94]. The ASBCPs at higher TRL after demonstra-
tion can be qualified for real CO2 capture applications. By applying these qualified
energy efficient ASBCPs the energy sector can be decarbonised [95, 96].

13 Conclusions and Outlook

The current study focused on energy efficient advanced solvent based capture
processes (ASBCPs) for capturing CO2 by gas-liquid absorption. The analysis of 11
types of different ASBCPs was presented by explaining useful mechanisms, energy
efficient benefits, practical examples and challenges. Several described ASBCPs
demonstrated significant reductions in terms of energy requirement for CO2 capture
compared to MEA as the solvent benchmark. However, many emerging ASBCPs
cannot be employed in the practice because of incompatibility with harsh operating
conditions and inability to overcome challenges associated with flue gas impurities.
The identified challenges of ASBCPs suggest that further researches are needed
for most of them. Overcoming these challenges will enable to achieve meaningful
progress toward commercialisation and wide-scale adoption of ASBCPs.
Candidates for ASBCPs need to be systematically screened, researched and
demonstrated. This study highlights certain groups of promising ASBCPs. For
instance precipitating solvents, two immiscible liquid phases, catalysed solvents
and microencapsulated solvents as well as their combinations. Other discussed
ASBCPs are interesting but need more fundamental insights in order to achieve
progress toward potential commercialisation.

Acknowledgments This study has been supported by the members of the Renewable Energy and
Sustainable Development (RESD) Group (Poland) under the project RESD-RDG03/2016 which is
gratefully acknowledged.

References

1. Abu-Zahra MRM, Abbas Z, Singh P, Feron P. (2013) Carbon dioxide post-combustion


capture: solvent technologies overview, status and future directions. In Méndez-Vilas E
(ed) Materials and processes for energy: communicating current research and technological
developments, vol 1. Formatex Research Center, pp 923–934
94 W.M. Budzianowski

2. Guo BS, Jing GH, Zhou ZM (2015) Regeneration performance and absorption/desorption
mechanism of tetramethylammonium glycinate aqueous solution for carbon dioxide capture.
Int J Greenhouse Gas Control 34:31–38
3. Vasudevan S, Farooq S, Karimi IA, Saeys M, Quah MCG, Agrawal R (2016) Energy penalty
estimates for CO2 capture: comparison between fuel types and capture-combustion modes.
Energy 103:709–714
4. Budzianowski WM (2012) Benefits of biogas upgrading to biomethane by high-pressure
reactive solvent scrubbing. Biofuels Bioprod Biorefin 6(1):12–20
5. Fu C, Gundersen T (2012) Using exergy analysis to reduce power consumption in air
separation units for oxy-combustion processes. Energy 44(1):60–68
6. Liu H, Liu B, Lin LC, Chen G, Wu Y, Wang J, Gao X, Lv Y, Pan Y, Zhang X, Zhang X,
Yang L, Sun C, Smit B, Wang W (2014) A hybrid absorption-adsorption method to efficiently
capture carbon. Nat Commun 5:5147
7. Perry RJ, Wood BR, Genovese S, Obrien MJ, Westendorf T, Meketa ML, Farnum R,
McDermott J, Sultanova I, Perry TM, Vipperla RK, Wichmann LA, Enick RM, Hong L,
Tapriyal D (2012) CO2 capture using phase-changing sorbents. Energy Fuels 26(4):
2528–2538
8. Gonzalez-Salazar MA, Perry RJ, Vipperla RK, Hernandez-Nogales A, Nord LO,
Michelassi V, Shisler R, Lissianski V (2012) Comparison of current and advanced post-
combustion CO2 capture technologies for power plant applications. Energy Procedia 23:3–14
9. Yu J, Wang S, Yu H (2014) Modelling analysis of solid precipitation in an ammonia-based
CO2 capture process. Int J Greenhouse Gas Control 30:133–139
10. Gazzani M, Sutter D, Mazzotti M (2014) Improving the efficiency of a chilled ammonia CO2
capture plant through solid formation: a thermodynamic analysis. Energy Procedia 63:
1084–1090
11. Fernandez ES, Goetheer ELV (2011) DECAB Process development of a phase change
absorption process. Energy Procedia 4:868–875
12. Sanchez Fernandez E (2013) Novel process designs to improve the efficiency of
postcombustion carbon dioxide capture. Ph.D. thesis, Delft University of Technology
13. Sanchez Fernandez E, Heffernan K, van der Ham L, Linders MJG, Eggink E, Schrama FNH,
Brilman DWF, Goetheer ELV, Vlugt TJH (2013) Conceptual design of a novel CO2 capture
process based on precipitating amino acid solvents. Ind Eng Chem Res 52(34):12223–12235
14. Aronu UE, Kim I, Haugen G (2014) Evaluation of energetic benefit for solid-liquid phase
change CO2 absorbents. Energy Procedia 63:532–541
15. Sanchez-Fernandez E, Heffernan K, van der Ham L, Linders MJG, Brilman DWF,
Goetheer ELV, Vlugt TJH (2014) Analysis of process configurations for CO2 capture by
precipitating amino acid solvents. Ind Eng Chem Res 53(6):2348–2361
16. Perry R. (2010) Exploring more cost-effective methods for CO2 capture. http://ge.geglobal
research.com/blog/exploring-more-cost-effective-methods-for-co2-capture/. Accessed Dec
2015
17. Perry R (2010) Phase changing solvents for CO2 capture. http://ge.geglobalresearch.com/
blog/phase-changing-solvents-for-co2-capture/. Accessed Dec 2015
18. Smith K, Xiao G, Mumford K, Gouw J, Indrawan I, Thanumurthy N, Quyn D, Cuthbertson R,
Rayer A, Nicholas N, Lee A, da Silva G, Kentish S, Harkin T, Qader A, Anderson C,
Hooper B, Stevens G (2014) Demonstration of a concentrated potassium carbonate process
for CO2 capture. Energy Fuels 28(1):299–306
19. Smith K, Lee A, Mumford K, Li S, Indrawan I, Thanumurthy N, Temple N, Anderson C,
Hooper B, Kentish S, Stevens G (2015) Pilot plant results for a precipitating potassium
carbonate solvent absorption process promoted with glycine for enhanced CO2 capture. Fuel
Process Technol 135:60–65
20. Anderson C, Hooper B, Qader A, Harkin T, Smith K, Mumford K, Pandit J, Ho M, Lee A,
Nicholas N, Indrawan I, Gouw J, Xiao J, Thanumurthy N, Temple N, Stevens G, Wiley D
(2014) Recent developments in the UNO MK 3 process—a low cost, environmentally benign
precipitating process for CO2 capture. Energy Procedia 63:1773–1780
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 95

21. Lupiáñez C, Guedea I, Bolea I, Díez LI, Romeo LM (2013) Experimental study of SO2 and
NOX emissions in fluidized bed oxy-fuel combustion. Fuel Process Technol 106:587–594
22. Asif M, Kim WS (2014) Modeling and simulation of the combined removal of SO2 and CO2
by aqueous ammonia. Greenhouse Gases Sci Technol 4(4):509–527
23. Zhang J, Qiao Y, Agar DW (2012) Intensification of low temperature thermomorphic biphasic
amine solvent regeneration for CO2 capture. Chem Eng Res Des 90:743–749
24. Yu YS, Lu HF, Zhang TT, Zhang ZX, Wang GX, Rudolph V (2013) Determining the
performance of an efficient non-aqueous CO2 capture process at desorption temperature below
373K. Ind Eng Chem Res 52(35):12622–12634
25. Monteiro JGM, Pinto DDD, Svendsen HF. (2011) Phase change solvents. In: EU-China
Workshop on Innovative CCS Technologies, Beijing, 19–20 Sept 2011
26. Raynal L, Bouillon PA, Gomez A, Broutin P (2011) From MEA to demixing solvents and
future steps, a roadmap for lowering the cost of post-combustion carbon capture. Chem Eng J
171(3):742–752
27. Zhang J, Nwani O, Tan Y, Agar DW (2011) Carbon dioxide absorption into biphasic amine
solvent with solvent loss reduction. Chem Eng Res Des 89:1190–1196
28. Monteiro JGMS, Pinto DDD, Zaidy SAH, Hartono A, Svendsen HF (2013) VLE data and
modelling of aqueous N, N-diethylethanolamine (DEEA) solutions. Int J Greenhouse Gas
Control 19:432–440
29. Pinto DDD, Knuutila H, Fytianos G, Haugen G, Mejdell T, Svendsen HF (2014) CO2 post
combustion capture with a phase change solvent. Pilot plant campaign. Int J Greenhouse Gas
Control 31:153–164
30. Xu Z, Wang S, Qi G, Liu J, Zhao B, Chen C (2014) CO2 absorption by biphasic solvents:
comparison with lower phase alone. Oil Gas Sci Technol 69(5):851–864
31. Ye Q, Wang X, Lu Y (2015) Screening and evaluation of novel biphasic solvents for
energy-efficient post-combustion CO2 capture. Int J Greenhouse Gas Control 39:205–214
32. Pendersen-van Elk NJMC, Derks PWJ, Fradette S, Veersteg GF (2012) Kinetics of absorption
of carbon dioxide in aqueous MDEA solutions with carbonic anhydrase at 298 K. Int J
Greenhouse Gas Control 9:385–392
33. Li L, Fu ML, Zhao YH, Zhu YT (2012) Characterization of carbonic anhydrase II from
Chlorella vulgaris in bio-CO2 capture. Environ Sci Pollut Res 19(9):4227–4232
34. Kunze AK, Dojchinov G, Haritos VS, Lutze P (2015) Reactive absorption of CO2 into
enzyme accelerated solvents: from laboratory to pilot scale. Appl Energy 156:676–685
35. Akermin Inc. (2013) Final Scientific/Technical Report—Advanced Low Energy Enzyme—
Catalyzed Solvent for CO2 Capture. http://www.netl.doe.gov/File%20Library/Research/Coal/
carbon%20capture/post-combustion/FE0004228-Final-Report-01-07-2014.pdf
36. Saeed M, Deng L (2016) Post-combustion CO2 membrane absorption promoted by mimic
enzyme. J Membr Sci 499:36–46
37. Phan DT, Maeder M, Burns RC, Puxty G (2014) Catalysis of CO2 absorption in aqueous
solution by inorganic oxoanions and their application to post combustion capture. Environ Sci
Technol 48(8):4623–4629
38. Kvamsdal HM, Kim I, van Os P, Pevidac C, Hägg MB, Browne J, Robinson L, Feron P
(2014) HiPerCap: a new FP7 project for development and assessment of novel and emerging
post-combustion CO2 capture technologies. Energy Procedia 63:6166–6172
39. Yong JKJ, Stevens GW, Caruso F, Kentish SE (2015) The use of carbonic anhydrase to
accelerate carbon dioxide capture processes. J Chem Technol Biotechnol 90:3–10
40. Li K, Yu H, Tade M, Feron P (2014) Theoretical and experimental study of NH3 suppression
by addition of Me(II) ions (Ni, Cu and Zn) in an ammonia-based CO2 capture process. Int J
Greenhouse Gas Control 24:54–63
41. Yu J, Wang S, Yu H (2013) Experimental studies on suppression of ammonia vaporization by
additives. Greenhouse Gases Sci Technol 3(5):415–422
42. Ma S, Song H, Chen G, Wang Y, Zang B (2013) Cobalt(II) as an additive inhibiting ammonia
escape in carbon capture using ammonia solution. Greenhouse Gases Sci Technol 3(5):
392–396
96 W.M. Budzianowski

43. Ma S, Song H, Zang B, Chen G (2013) Experimental study on additives inhibiting ammonia
escape in carbon capture process using ammonia method. Chem Eng Res Des 91(12):
2775–2781
44. Ma S, Chen G, Han T, Zhu S, Yang J (2015) Experimental study on the effect of Ni(II)
additive on ammonia escape in CO2 capture using ammonia solution. Int J Greenhouse Gas
Control 37:249–255
45. Stolaroff J, Vericella J, Baker S, Spadaccini C, Duoss E, Aines R. Micro-encapsulation of
advanced solvents for post-combustion carbon capture. In: 3rd Post Combustion Capture
Conference (PCCC3)
46. Vericella JJ, Baker SE, Stolaroff JK, Duoss EB, Hardin JO, Lewicki J, Glogowski E,
Floyd WC, Valdez CA, Smith WL, Satcher JH Jr, Bourcier WL, Spadaccini CM, Lewis JA,
Aines RD (2015) Encapsulated liquid sorbents for carbon dioxide capture. Nat Commun
6:6124
47. Aines RD, Bourcier WL, Spadaccini CM, Stolaroff JK (2015) Polymer-encapsulated carbon
capture liquids that tolerate precipitation of solids for increased capacity. US 8945279B2
48. Asif M, Kim WS (2015) Process simulation of ammonia-based CO2 capture and regeneration
in packed column. Int J Glob Warming 8(3):401–415
49. Favre E, Svendsen HF (2012) Membrane contactors for intensified post-combustion
carbon-dioxide capture by gas-liquid absorption processes. J Membr Sci 407–408:1–7
50. Lu JG, Ji Y, Zhang H, Chen MD (2010) CO2 capture using activated amine acid salt solutions
in a membrane contactor. Sep Sci Technol 45:1240–1251
51. Budzianowski WM (2016) Explorative analysis of advanced solvent processes for energy
efficient carbon dioxide capture by gas-liquid absorption. Int J Greenhouse Gas Control
49:108–120
52. Budzianowski WM (2015) Single solvents, solvent blends, and advanced solvent systems in
CO2 capture by absorption: a review. Int J Glob Warming 7(2):184–225
53. Budzianowski WM (2011) CO2 reactive absorption from flue gases into aqueous ammonia
solutions: the NH3 slippage effect. Environ Prot Eng 37(4):5–19
54. Scovazzo P (2009) Determination of the upper limits, benchmarks, and critical properties for
gas separations using stabilized room temperature ionic liquid membranes (SILMs) for the
purpose of guiding future research. J Membr Sci 343:199–211
55. Bara JE, Gabriel CJ, Carlisle TK, Camper DE, Finotello A, Gin DL, Noble RD (2009) Gas
separations in fluoroalkyl-functionalized room-temperature ionic liquids using supported
liquid membranes. Chem Eng J 147:43–50
56. Cserjesi P, Nemestothy N, Belafi-Bako K (2010) Gas separation properties of supported
liquid membranes prepared with unconventional ionic liquids. J Membr Sci 349:6–11
57. Neves LA, Crespo JG, Coelhoso IM (2010) Gas permeation studies in supported ionic liquid
membranes. J Membr Sci 357:160–170
58. Scovazzo P, Havard D, McShea M, Mixon S, Morgan D (2009) Long-term, continuous
mixed-gas dry fed CO2/CH4 and CO2/N2 separation performance and selectivities for room
temperature ionic liquid membranes. J Membr Sci 327:41–48
59. Myers C, Pennline H, Luebke D, Ilconich J, Dixon JK, Maginn EJ, Brennecke JF (2008) High
temperature separation of carbon dioxide/hydrogen mixtures using facilitated supported ionic
liquid membranes. J Membr Sci 322:28–31
60. Yoo S, Won J, Kang SW, Kang Y, Nagase S (2010) CO2 separation membranes using ionic
liquids in a Nafion matrix. J Membr Sci 363:72–79
61. Iarikov DD, Hacarlioglu P, Oyama ST (2011) Supported room temperature ionic liquid
membranes for separation. Chem Eng J 166:401–406
62. Hanioka S, Maruyama T, Sotani T, Teramoto M, Matsuyama H, Nakashima K, Hanaki M,
Kubota F, Goto M (2008) CO2 separation facilitated by task-specific ionic liquids using a
supported liquid membrane. J Membr Sci 314:1–4
63. Mumford KA, Wu Y, Smith KH, Stevens GW (2015) Review of solvent based
carbon-dioxide capture technologies. Front Chem Sci Eng 9(2):125–141
Useful Mechanisms, Energy Efficiency Benefits, and Challenges … 97

64. Tang J, Tang H, Sun W, Plancher H, Radosz M, Shen Y (2005) Poly (ionic liquid) s: a new
material with enhanced and fast CO2 absorption. Chem Commun 26:3325–3327
65. Ziobrowski Z, Krupiczka R, Rotkegel A (2016) Carbon dioxide absorption in a packed
column using imidazolium based ionic liquids and MEA solution. Int J Greenhouse Gas
Control 47:8–16
66. Ramdin M, de Loos TW, Vlugt TJH (2012) State-of-the-art of CO2 capture with ionic liquids.
Ind Eng Chem Res 51(24):8149–8177
67. Zhang X, Zhang X, Dong H, Zhao Z, Zhang S, Huang Y (2012) Carbon capture with ionic
liquids: overview and progress. Energy Environ Sci 5:6668–6681
68. Heldebrant DJ, Yonker CR, Jessop PG, Phan L (2009) CO2-binding organic liquids
(CO2BOLs) for post-combustion CO2 capture. Energy Procedia 1(1):1187–1195
69. Heldebrant DJ, Koech PK, Ang MTC, Liang C, Rainbolt JE, Yonker CR, Jessop PG (2010)
Reversible zwitterionic liquids, the reaction of alkanol guanidines, alkanol amidines, and
diamines with CO2. Green Chem 12:713
70. Jessop PG, Heldebrant DJ, Li X, Eckert CA, Liotta CL (2005) Green chemistry: reversible
nonpolar-to-polar solvent. Nature 436:1102
71. Koech PK, Zhang J, Kutnyakov IV, Cosimbescu L, Lee SJ, Bowden ME, Smurthwaite TD,
Heldebrant DJ (2013) Low viscosity alkanolguanidine and alkanolamidine liquids for CO2
capture. RSC Adv 3(2):566–572
72. Mathias PM, Afshar K, Zheng F, Bearden MD, Freeman CJ, Andrea T, Koech PK,
Kutnyakov I, Zwoster A, Smith AR, Jessop PG, Nik OG, Heldebrant DJ (2013) Improving
the regeneration of CO2-binding organic liquids with a polarity change. Energy Environ Sci 6
(7):2233–2242
73. Heldebrant DJ, Yonker CR, Jessop PG, Phan L (2008) Organic liquid CO2 capture agents
with high gravimetric CO2 capacity. Energy Environ Sci 1(4):487–493
74. Heldebrant DJ, Koech PK, Rainbolt JE, Zheng F, Smurthwaite T, Freeman CJ, Oss M, Leito I
(2011) Performance of single-component CO2-binding organic liquids (CO2BOLs) for post
combustion CO2 capture. Chem Eng J 171(3):794–800
75. Öztürk MÇ, Yüksel Orhan Ö, Alper E (2014) Kinetics of carbon dioxide binding by
1,1,3,3-tetramethylguanidine in 1-hexanol. Int J Greenhouse Gas Control 26:76–82
76. Yüksel Orhan Ö, Öztürk MÇ, Şeker A, Alper E (2015) Kinetics and performance studies of a
switchable solvent TMG (1,1,3,3-tetramethylguanidine)/1-propanol/carbon dioxide system.
Turk J Chem 39(1):13–24
77. Li L, Conway W, Puxty G, Burns R, Clifford S, Maeder M, Yu H (2015) The effect of
piperazine (PZ) on CO2 absorption kinetics into aqueous ammonia solutions at 25.0 °C. Int J
Greenhouse Gas Control 36:135–143
78. Zhang J, Kutnyakov I, Koech PK, Zwoster A, Howard C, Zheng F, Freeman CJ,
Heldebrant DJ (2013) CO2-binding-organic-liquids-enhanced CO2 capture using
polarity-swing-assisted regeneration. Energy Procedia 37:285–291
79. Zheng F, Heldebrant D, Mathias P, Koech P, Bhakta M, Freeman C, Bearden M, Zwoster A
(2016) Bench scale testing and process performance projections of CO2 capture by CO2BOLs
with and without polarity swing assisted regeneration. Energy Fuels 30(2):1192–1203
80. Heyn RH, Aronu UE, Vevelstad SJ, Hoff KA, Didriksen T, Arstad B, Blom R (2014) Use of
metal-organics based solvents for CO2 capture. Energy Procedia 63:1805–1810
81. Ren H-Y, Zhang X-M (2016) Enhanced selective CO2 capture upon incorporation of
dimethylformamide in the cobalt Metal-Organic Framework [Co3(OH)2(btca)2]. Energy Fuels
30(1):526–530
82. Oyarzun B, van Westen T, Vlugt TJH (2013) The phase behaviour of linear and partially
flexible hard-sphere chain fluids and the solubility of hard spheres in hard-sphere chain fluids.
J Chem Phys 138:204905
83. de Groen M, Vlugt TJH, de Loos T (2015) Binary and ternary mixtures of liquid crystals with
CO2. AIChE J 61(9):2977–2984
84. Gross J, Jansens PJ. (2008) A method for the removal of a gas from a process gas stream by
means of liquid crystals. Patent WO2008147181A1
98 W.M. Budzianowski

85. Zhang Q, de Oliveira Vigier K, Royer S, Jerome F (2012) Deep eutectic solvents: syntheses,
properties and applications. Chem Soc Rev 41:7108–7146
86. Carriazo D, Serrano MC, Gutierrez MC, Ferrer ML, del Monte F (2012) Deep-eutectic
solvents playing multiple roles in the synthesis of polymers and related materials. Chem Soc
Rev 41:4996–5014
87. Paiva A, Craveiro R, Aroso I, Martins M, Reis RL, Duarte ARC (2014) Natural deep eutectic
solvents—solvents for the 21st Century. ACS Sustain Chem Eng 2:1063–1071
88. Florindo C, Oliveira FS, Rebelo LPN, Fernandes AM, Marrucho IM (2014) Insights into the
synthesis and properties of deep eutectic solvents based on cholinium chloride and carboxylic
acids. ACS Sustain Chem Eng 2:2416–2425
89. García G, Atilhan M, Aparicio S (2015) A theoretical study on mitigation of CO2 through
advanced deep eutectic solvents. Int J Greenhouse Gas Control 39:62–73
90. Abbott AP, Capper G, Davies DL, Rasheed RK, Tambyrajah V. (2003) Novel solvent
properties of choline chloride/urea mixtures. Chem Commun 70–71
91. Wagle DV, Zhao H, Baker GA (2014) Deep eutectic solvents: sustainable media for
nanoscale and functional materials. Acc Chem Res 47:2299–2308
92. Lu M, Han G, Jiang Y, Zhang X, Deng D, Ai N (2015) Solubilities of carbon dioxide in the
eutectic mixture of levulinic acid (or furfuryl alcohol) and choline chloride. J Chem
Thermodyn 88:72–77
93. García G, Aparicio S, Ullah R, Atilhan M (2015) Deep eutectic solvents: physicochemical
properties and gas separation applications. Energy Fuels 29:2616–2644
94. Budzianowski WM (2016) A review of potential innovations for production, conditioning and
utilization of biogas with multiple criteria assessment. Renew Sustain Energy Rev 54:
1148–1171
95. Budzianowski WM (2012) Negative carbon intensity of renewable energy technologies
involving biomass or carbon dioxide as inputs. Renew Sustain Energy Rev 16(9):6507–6521
96. Budzianowski WM (2012) Value-added carbon management technologies for low CO2
intensive carbon-based energy vectors. Energy 41(1):280–297
Phase Change Solvents for CO2
Capture Applications

Kathryn A. Mumford, Kathryn H. Smith and Geoffrey W. Stevens

Abstract Solvent systems that separate into two phases upon absorption of CO2,
one rich and one lean in CO2, have significant potential to exhibit reduced energy
requirements. This reduction stems primarily from the ability to separate the two
phases such that only the stream containing CO2 is heated in the regenerator
resulting in reduced sensible and latent heating requirements. Thus, this class of
solvents are currently under intense investigation in universities and industries
globally. This chapter presents recent developments in Solid-Liquid and
Liquid-Liquid phase change solvents for CO2 capture, including laboratory, pilot
scale and commercial system installations.

Nomenclature
A Amine (in Fig. 2)
AAS Amino acid salt
ABS Absorber
AMP 2-Amino-2-methyl-1-propanol
COND Condenser
DEA Diethanolamine
DEEA 2-(diethylamino)ethanol
DMCA Dimethylcyclohexylamine
DPA Dipropylamine
HEX Heat exchanger
HP CO2 High pressure carbon dioxide
Kn Equilibrium constant
LP CO2 Low pressure carbon dioxide
MAPA 3-(methylamino)propylamine
MEA Monoethanolamine
NMP N-methyl-2-pyrrolidone

K.A. Mumford (&)  K.H. Smith  G.W. Stevens


The University of Melbourne, Melbourne, Australia
e-mail: mumfordk@unimelb.edu.au

© Springer International Publishing AG 2017 99


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_5
100 K.A. Mumford et al.

P Pump
REB Reboiler
SEP Separator
TBS Thermomorphic biphasic solvent
TEGDME Triethylene glycol dimethyl ether
TEPA Tetraethylenepentamine
TETA Triethylenetetramine

1 Introduction

High regeneration energy requirements for solvents used in CO2 capture processes
are a key challenge in capture technology process development. Solvents that
undergo a phase change upon CO2 absorption (such as precipitation or generation
of two immiscible phases) have recently attracted attention as a promising option
for reducing the regeneration energy requirement. It is known that certain com-
ponents in solvent systems may form a precipitate upon absorption of CO2. This
precipitation results in the formation of a slurry which may then be separated into
two streams: a stream rich in CO2 and a stream lean in CO2. Potentially, the use of
precipitating solvents may have several advantages over traditional, liquid based
systems resulting from; their ability to maintain absorption at higher CO2 loadings
resulting in higher cyclical loadings and lower solvent recirculation rates; increased
kinetics and smaller contacting equipment; and reduced energy consumption during
regeneration due to a reduction in latent heat requirements, as compared to tradi-
tional liquid amine based systems. A number of precipitating solvent systems are
being investigated by different companies and research institutions including;
chilled ammonia (Alstom [1]), amino acids (TNO [42], NTNU [35]) and potassium
carbonate (UNO Technology [15], University of Melbourne [33, 34], Shell [23]), in
addition to systems which utilise organic solvents as diluents to the reactive solute
as opposed to water. Processes under development include both the precipitation of
the carbon dioxide containing species and non-carbon dioxide containing species
[8]. As yet, no precipitating processes are operating at an industrial scale for carbon
dioxide capture, as the design and operation of suitable equipment for precipitating
systems requires further development and optimisation. Another class of biphasic
solvents results in two immisicible liquid phases following absorption of CO2. One
liquid phase will have a low CO2 content while the other phase will be high in CO2.
This resulting CO2 rich phase can be regenerated with a much lower energy penalty
than traditional liquid processes. Commercial examples of these types of biphasic
solvents include the DMX process, thermomorphic solvents and mixed amine (e.g.
DEEA/MAPA) systems. This chapter will discuss recent advances in phase change
solvents for CO2 capture, including solid-liquid and liquid-liquid processes.
Phase Change Solvents for CO2 Capture Applications 101

2 Solid-Liquid Phase Change Solvents for CO2 Capture

A number of precipitating solvent processes have been developed for CO2 capture
applications, including chilled ammonia, amino acid salts and carbonate solvents.
All of these processes take advantage of the significant reduction in the energy
penalty of the regeneration process when using a slurry or precipitating solvent.
Other benefits of precipitating CO2 capture processes include a concentration step
for the solids which leads to higher CO2 partial pressure and a lower solvent flow
rate to the regenerator compared to competing processes [23].

2.1 Precipitating Chilled Ammonia Process

The use of chilled ammonia to capture carbon dioxide was first patented by Gal
[10, 11] in 2006. This process is based upon the use of ammonia and other trace
chemicals to absorb carbon dioxide from flue gases at low temperatures and
pressures. Post-combustion flue gas streams emitted from power stations or other
low pressure CO2 containing streams from industrial sources have been identified
as the most likely market for this technology. Unlike traditional solvents, such as
monoethanolamine (MEA), the ammonia solvent solution stability is not adversely
impacted by oxygen or other trace contaminants, and operation at low temperatures
allows use of waste heat for regeneration purposes and also decreases the ammonia
slip in the absorber. Additionally, it is reported that the heat of absorption of CO2
by ammonia is significantly lower than for amines.
In the chilled ammonia process, CO2, ammonia and water combine to form
ammonium carbonate ððNH4 Þ2 CO3 Þ, ammonium bicarbonate ðNH4 HCO3 Þ and
ammonium carbamate ðNH2 CO2 NH4 Þ all of which have the potential to form
precipitates. The main reactions associated with this process are presented as
reaction (1)–(6).

CO2ðgÞ $ CO2ðlÞ ð1Þ

2NH3ðaqÞ þ H2 OðlÞ þ CO2ðaqÞ $ ðNH4 Þ2 CO3ðaqÞ $ ðNH4 Þ2 CO3ðsÞ ð2Þ

NH3ðaqÞ þ H2 OðlÞ þ CO2ðaqÞ $ NH4 HCO3ðaqÞ $ NH4 HCO3ðsÞ ð3Þ

ðNH4 Þ2 CO3ðaqÞ þ CO2ðaqÞ þ H2 OðlÞ $ 2ðNH4 ÞHCO3ðaqÞ $ 2ðNH4 ÞHCO3ðsÞ ð4Þ

ðNH4 Þ2 CO3ðaqÞ $ NH2 CO2 NH4ðaqÞ þ H2 OðlÞ $ NH2 CO2 NH4ðsÞ ð5Þ

NH4 HCO3ðsÞ $ NH3ðaqÞ þ H2 OðlÞ þ CO2ðlÞ ð6Þ

The overall capture process operates in a similar fashion to typical solvent


absorption systems. First the flue gas passes through a direct contact cooler to
102 K.A. Mumford et al.

reduce the temperature to between 2 and 10 °C. Then the cooled gas enters the
bottom of the absorber and lean solvent, comprising water, ammonia (approxi-
mately 29 wt%) and a low concentrations of CO2 enters from the top. The tem-
perature of the absorber is kept low to prevent ammonia slip. CO2 in the gas phase
reacts according to the reaction scheme outlined in reactions (1)–(6) and the CO2
lean flue gas exits through the top of the absorber after passing through an acid
water wash system to remove any residual ammonia. The CO2 rich solvent stream
exits at the bottom of the absorber. If sufficient CO2 has been absorbed the solu-
bility limit of the product species can be reached and a precipitate formed. As the
absorber is run at low temperatures the reaction rate of this system is lower than
traditional amine systems and hence a larger than typical absorber is often required.
The CO2 rich solvent stream that has exited the absorber passes through a heat
exchanger to raise its temperature and then enters the regenerator at the top where it
is heated to between 100 and 150 °C where the CO2 is released. The use of the
higher stripping temperature allows the CO2 to be released at a higher pressure.
Alstom’s chilled ammonia process has been trialled at a number of sites through
their extensive partnerships [37], including;
• 20 MWe pilot at AEP’s Mountaineer Power Plant [37]
• 20 MW application at Statoil Hydro’s Mongstad Test Centre in Norway [18].

2.2 Precipitating Amino Acid Salts

Amino acids are a subset of amines whereby the side chain contains a carboxylic
acid group i.e. R ¼ OOC  R0 . This class of solvent has advantages over other
amine types due to their higher resistance to oxidative degradation and negligible
vapour pressure, both of which lead to a more environmentally benign capture
process [12]. Primary and secondary amino acids react with CO2 as per the scheme
presented in Fig. 1. As shown, in these systems 2 mol of amino acid ðRNH2 Þ react
with one mole of CO2 to form a carbamate ðRNHCO 2 Þ and a zwitterion species
þ
ðRNH3 Þ. The carbamate and zwitterion species may then undergo hydrolysis to
form amino acid, bicarbonate ðHCO 3 Þ and zwitterion. These species may then
either form carbonate at high pH, or be regenerated to the starting amino acid and
CO2 via the addition of heat. When using amino acid salts for CO2 absorption
neutralization using equimolar potassium hydroxide is usually necessary. In recent

Fig. 1 Schematic of RNHCO2- RNH3+


carbamate, bicarbonate and carbamate
carbonate formation using
amines (primary or 2RNH3+ CO32-
2RNH2 + CO2 H 2O carbonate
secondary)
amine
pH
+ -
heat RNH3 HCO3 + RNH2
bicarbonate
Phase Change Solvents for CO2 Capture Applications 103

years, interest has grown in the potential use of different amino acid salts for
post-combustion CO2 capture. In fact, the CO2 absorption characteristics for
common amino acids and their properties have been under extensive investigation
[16, 19]. Siemens Energy also brought commercial interest to Amino Acid salt
based solvents in 2010 through the development of their PostCap technology that
was tested at a 2.5 MW pilot facility in 2009 [27] and later at Polk Power Plant
[28]. Siemens reports that the energy consumption of this process is approximately
73% of the conventional monoethanolamine (MEA) process.
More recently, developments that use amino acids have moved towards inducing
precipitation of either the zwitterion ðRNH3þ Þ, a non-carbon dioxide containing
species, or bicarbonate, a carbon dioxide containing species, with the view to
driving the CO2 capture reaction forward. The relative proportion of the precipitate
product depends upon a number of parameters including; the amine concentration;
solution pH, temperature and the chemical stability of the carbamate formed.
Primary amino acids tend to form stable carbamates and hence at equilibrium a
higher proportion of these species is formed, as is the case with taurine [8].
Conversely, secondary amino acids, which have a higher level of steric hindrance,
tend to form both the zwitterion and bicarbonate, as in the case of be proline,
sarcosine and b-alanine [20]. Tertiary amino acids, with the highest steric hin-
drance, are generally unable to form stable carbamates at all and so all of the
reaction products for these species are bicarbonates [16, 20]. It may also be noted
from Fig. 1 that the maximum absorption of CO2 is achieved when all of the
absorbed CO2 exists as bicarbonate as only one mole of amine is required per mole
of CO2 [14]. Also of importance is the regeneration of the amino acid. It has been
found that some amino acid solutions, particularly glycine and alanine, fail to
produce a sufficiently lean solution after desorption which means that they do not
commence reabsorption at a favourable point in the absorption curve [14].
Therefore careful consideration of the regeneration process, as well as absorption
process, is valuable.
Systems under development that utilise precipitating amino acid salts include;
• The DECAB Plus process—precipitation of the protonated amine
• CASPER (CO2-capture And Sulphur Precipitation for Enhanced Removal)—
CO2 and SO2 precipitation process.

2.2.1 DECAB and DECAB Plus Process

The DECAB and DECAB Plus Processes are based on the precipitation of a pri-
mary amino acid, such as taurine, as a zwitterion [9]. The DECAB process refers to
a process whereby a spray tower is used to handle solids during absorption. The
precipitates are re-dissolved before desorption, which takes place in a conventional
stripper [8]. The DECAB Plus process refers to a process where the supernatant is
separated from the slurry and recycled to the absorber. This acts to enhance the
104 K.A. Mumford et al.

release of CO2 in the remaining liquid inside the regenerator via an induced pH
swing when the precipitate dissolves upon heating [31].
The inventors report that upon loading with carbon dioxide, the solubility limit of
the zwitterion may be reached, resulting in its precipitation from solution. As a
consequence of the precipitation of the acidic species, the pH rises, resulting in the
ability of the solvent to absorb more CO2 from the gas stream. When the solvent
exits the absorber, the supernatant is separated from the slurry, and returned to the
absorber. The slurry stream, which now has a reduced K+ concentration compared to
the amino acid, is heated so that the precipitate dissolves, resulting in a reduction in
pH. The lower pH promotes the hydrolysis of the carbamate species, resulting in the
formation of amino acid salt and bicarbonate, which reverts to the zwitterion species
and carbon dioxide at lower pH, thereby resulting in a reduction in energy required
for solvent regeneration. Fernandez [9] explored this concept in detail with 4 M
aqueous potassium taurate, a primary amine. They found that this process reduces
the energy penalty by approximately 35% compared to the MEA baseline [9].
Large scale implementation of this technology has been limited due to two main
reasons:
– Design and optimisation of appropriate contacting equipment for the precipi-
tating system has not yet been finalised
– Although the chemical equilibrium of this system has been investigated in
detail, the reaction kinetics have not yet been fully investigated [31].

2.2.2 CASPER (CO2-Capture and Sulphur Precipitation


for Enhanced Removal)

CASPER is a CO2 capture and sulphur precipitation process for simultaneous CO2
and SO2 removal from flue gas [21]. It uses a potassium amino acid based aqueous
solution and precipitates sulphur as K2SO4. It was tested by CSIRO in 2013 at Loy
Yang Power Station, Victoria, Australia and it was found that the performance of
3 M potassium b-alanate was satisfactory as compared to MEA in the same pilot
plant facility.

2.3 Precipitating Carbonate Solvent Systems

Carbon dioxide capture processes using potassium carbonate have been imple-
mented in industry for many years. Potassium carbonate solvents have a number of
advantages including; low vapour pressure, low toxicity and minimal oxidative
degradation. The main challenge associated with potassium carbonate capture
systems when applied to post-combustion capture streams is the slow reaction rate.
However it has been found that the use of promoters which have an active N–H
Phase Change Solvents for CO2 Capture Applications 105

group can accelerate the reaction rate significantly [6, 38, 39]. Amines and, their
subset, amino acids have been successfully utilised to achieve this goal.
Similar to precipitating amino acid salt (AAS) based solvent systems, precipi-
tating potassium carbonate systems may further improve CO2 absorption efficiency
and reduce regeneration energy requirements [24, 40]. Generally, in promoted
potassium carbonate systems, as the concentration of the amino acid is kept low
(10 wt% amino acid in 30–45 wt% potassium carbonate [38]) it does not reach its
solubility limit and the bicarbonate only is precipitated. Fernandes and co-workers
[7] have developed a conceptual diagram that describes the interplay between the
important reactions associated with this process. Specifically it describes how the
species composition changes with acidity, water and amino acid concentration.
A representation of this diagram is shown in Fig. 2.
Lee et al. [17] and Thee [38] have investigated the interactions between car-
bonate and amino acid promotors extensively including the reaction rate and
vapour–liquid–solid equilibria. Their results corroborated with the model of
Fernandes [7]. Specifically, they found that there was a reaction between the dis-
solved carbon dioxide and amino acid [38] (K7 and K8) which acted to increase the
overall reaction rate and additionally that the amino acid does form a carbamate
[17] but only the bicarbonate is of sufficient concentration to reach its solubility
limit and form a precipitate at post-combustion capture conditions. Therefore, if the
solids containing stream is concentrated, and separated from the supernatant there
will be a lower liquid flow to the regenerator and thus a lower latent energy
requirement.

K4 K3
H2CO3 HCO3- CO32-

KCO2
K1 K2

+H2O
+H+
CO2 + H2O CO2(aq) + OH-
K5
+RNH3
K7
K9
K10

K6
ACO2H ACO2- RNH3+ RNH2
K8

Fig. 2 General reaction scheme for amine, CO2, carbonate in aqueous solution, where Kn
represents an equilibrium constant and A—amine [7]
106 K.A. Mumford et al.

2.3.1 UNO MK3

The CO2CRC and the University of Melbourne have been investigating promoted
potassium carbonate solvent systems for CO2 capture for a number of years. The
most recent development has been a precipitating promoted potassium carbonate
system termed UNO MK3 and this process is being commercialised by UNO
Technology Pty Ltd. The process was developed in the laboratory at bench scale
and then tested in pilot plants with a capacity of up to 1 tonne/day of CO2 using
both synthetic and actual flue gas [33, 34]. In the precipitating system design,
potassium bicarbonate is precipitated from a promoted potassium carbonate solvent
following CO2 absorption and subsequent cooling. The precipitate can then be
separated from the liquid phase via a hydrocyclone for selective regeneration of the
KHCO3 species. A process flow diagram of the process can be found in Fig. 3. In
this way, less water is passed to the regeneration stage and thus drives down the
energy requirements from over 3 GJ/tonne CO2 for a liquid based system to less
than 2.5 GJ/tonne for a precipitating based system [2].
The most widely reported system has been the use of 10 wt% potassium glycine,
as a promotor, in a 40–45 wt% K2CO3 solution. The addition of glycine was found

Fig. 3 UNO MK 3 process flow diagram [33]


Phase Change Solvents for CO2 Capture Applications 107

to improve the CO2 recovery rate by up to 6 times in a laboratory based pilot plant
trial whilst also slightly increasing the pressure drop and holdup which was thought
likely due to a reduction of the surface tension of the solvent [33]. A glycine
promoted precipitating K2CO3 solvent process has also been demonstrated using a
pilot plant that captured CO2 from flue gas generated from a brown coal based
pulverised coal combustion power station in Victoria, Australia [32].

2.3.2 Shell Carbonate Slurry Process

Shell have been developing a carbonate slurry process that uses potassium car-
bonate solvent with a crystallisation and concentration step (see Fig. 4) following
CO2 absorption [23, 41]. These studies have reported a potential *50% reduction
in regeneration energy and the potential for lower nitrosamine emissions compared
to traditional amine solvent processes [22]. The Shell carbonate process has been
tested using a bench scale pilot plant that has a capacity of 25 kg/day CO2. An
accelerator is used to enhance the mass transfer of CO2 to the liquid phase. The
critical steps of this process were reported to be the crystal formation and solids
handling equipment used to operate as a precipitating solvent process.

Fig. 4 Process flow diagram of the shell carbonate slurry process [23]
108 K.A. Mumford et al.

3 Non-aqueous and Liquid-Liquid Phase Change Solvents


for CO2 Capture

3.1 Non-aqueous Solvents

Recently there has been focus on the development of non-aqueous solvent systems
for CO2 capture applications. Typically these systems utilise well accepted solvents
such as MEA or DEA, but instead of being dissolved in water, they are dissolved in
organic solvents, such as ethanol or butanol. This change impacts the chemical
reactivity of these solvents considerably and has the potential to provide significant
benefits, including;
• Reduced corrosivity
• Increased reaction rates.
– solubility of CO2 in alcohols such as ethanol is significantly higher than in
water [46].
– alcohols such as ethanol can facilitate chemical reactions between solutes
and CO2 [46].
• Decreased regeneration temperatures
• As organic solvents are used, the formation of bicarbonate can’t occur and an
unstable carbamate is formed instead. This species, if it is unstable, requires
less energy to regenerate [46]. Additionally, as lower temperatures are
required for regeneration, waste heat, from other sources may also be used for
this duty.
Recently, developments in these solvent systems have focused on loading the
solvent with CO2 until a precipitate is formed. However, not all systems form
crystal like precipitates as shown in Table 1. As indicated, alkanolamines such as
MEA and DEA form gum structures rather than crystals upon absorption of CO2,
whereas alkylamines such as TETA or TEPA almost always give crystals. It is
postulated that the hydroxyl groups of MEA and DEA are likely causing this due to
their ability to form strong hydrogen bonded networks. Therefore this aspect needs
to be considered whilst selecting appropriate systems for use.
Zheng et al. [46] performed an extensive study investigation into different sol-
vent combinations, and in particular found that TETA in ethanol was a strong
candidate. They found that TETA in ethanol exhibits a higher reaction rate and
absorption capacity as compared to TETA in water solutions, additionally 81.8% of
the absorbed CO2 in the solid phases was TETA carbamate, whereas in water no
precipitate is observed. Additionally, the decomposition of TETA-carbamate,
releasing CO2 commenced at approximately 90 °C.
Other systems investigated include DEA in various ionic liquids
(1-ethyl-3-mthylimidazolium bis(trifluoromethylsulfonyl)imide) [emim][Tf2N],
1-butyl-3-mthylimidazolium bis(trifluoromethylsulfonyl)imide) [bmim][Tf2N],
Table 1 Physical state of products of amine-CO2 reactions in various solvents
Monoethanolamine Diethanolamine Triethylenetetramine Tetraethylenepentamine
(MEA) (DEA) (TETA) (TEPA)
Ethanol Gum [46] Gum [46] Powder [46] Powder [46]
1-butanol Gum [46] Gum [46] Powder [46] Powder [46]
1-pentanol Gum [46] Gum [46] Powder [46] Powder [46]
1-hexanol Gum [46] Gum [46] Powder [46] Powder [46]
Phase Change Solvents for CO2 Capture Applications

Iso-octanol Gum [46] Gum [46] Powder [46] Powder [46]


IL [emim][Tf2N] [bmim][Tf2N] Powder [13]
[hmim][Tf2N]
109
110 K.A. Mumford et al.

1-hexyl-3-mthylimidazolium bis(trifluoromethylsulfonyl)imide) [hmim][Tf2N])


[13] which upon loading with CO2 formed a DEA carbamate. It is also interesting
to note that in the case of [emim][Tf2N] and [hmim][Tf2N], the precipitate rose to
the surface quickly due to the hydrophobicity of the ionic liquid. Therefore it is
postulated that in an industrial setting separation of the phases would somewhat
simpler as compared to other potential solvent systems [13].
AMP/DEA in 1,2-propandiol/ethanol [5] formed an AMP-carbamate which was
found to regenerate at 80 °C, and AMP in N-methyl-2-pyrrolidone (NMP) and
triethylene glycol dimethyl ether (TEGDME) which also formed AMP carbamates
was found to regenerate at 75 and 90 °C respectively [36]. It is worth noting that
the TETA-carbamate regeneration temperature was higher than that of DEA-
carbamate and AMP-carbamate, which suggests TETA-carbamate is more stable
than DEA-carbamate and AMP-carbamate [46].

3.2 Liquid-Liquid Phase Change Solvents

A number of processes have been developed with liquid-liquid phase change sol-
vent properties. When the solvent or solvent blend absorbs CO2, two immiscible
liquid phases are formed; one that is rich in CO2 and another that is low in CO2.
The CO2 rich phase can be separated by gravity and then regenerated with a much
lower regeneration energy requirement. A number of different commercial pro-
cesses and solvent blends have been proposed.

3.2.1 DMXTM Process (IFP Energies Nouvelles)

A demixing process for CO2 capture, known as DMXTM, was originally developed
by IFP Energies noevelles. This process uses demixing solvents to form two
immiscible liquid phases following CO2 absorption at specific CO2 loading and
temperature conditions [30]. The apparent CO2 loading achievable with this process
is comparable to the standard amine process however with the demixing solvent the
salts are formed with CO2 concentrated in the resulting lower liquid phase. This
results in the lower or heavy aqueous liquid phase having a high CO2 loading which
can be separated via a decantation step and then only this CO2 rich phase needs to
be sent to the regeneration column. The light amine phase, which has a very low
CO2 content, is mixed with the regenerated solvent from the regeneration column
and then returned to the absorber. Refer to Fig. 5 for a process flow diagram of the
DMXTM process. The cyclic capacity of the process is increased and the mass of
solvent to be regenerated is reduced resulting in a lower energy penalty. DMX-1
has been used as the demixing solvent due to its thermodynamic capacity as well as
favourable degradation properties. The reported energy consumption for this pro-
cess is 2.3 GJ/tCO2 with reports that this could be reduced to as low as 2.1 GJ/tCO2
following optimised heat integration, indicating significant operating cost
Phase Change Solvents for CO2 Capture Applications 111

Fig. 5 Process flow diagram of the DMXTM process [30]

reductions. The capital cost or CAPEX of this process is estimated to be similar to a


standard amine capture plant. Although there are reported capital cost savings from
operating with a higher capacity solvent (such as a smaller diameter column, pumps
and heat exchangers) this benefit is counterbalanced by the cost of the decanter
equipment and higher column height from reduced reaction kinetics.

3.2.2 Thermomorphic Biphasic Solvent (TBS) Systems

Thermomorphic biphasic solvent (TBS) systems release absorbed CO2 in the


regeneration column at much lower temperatures (e.g. 80 °C or lower) than stan-
dard amine solvent processes (which typically operate around 120–130 °C). These
solvent systems also have a high CO2 capacity (e.g. 0.9 mol CO2/mol absorbant)
[44]. The heating and agitation during the regeneration process results in two liquid
phases being formed. When returned to the absorber the system reverts back to a
single liquid phase. Lipophilic amines, such as N,N-dimethylcyclohexylamine
(DMCA) and dipropylamine (DPA) are typically used in this process. When the
organic phase is formed in the regenerator, it acts as the extractant which removes
the amine from the aqueous phase resulting in dissociation of the bicarbonate and
carbamate in the loaded aqueous phase. These lipophilic amines have high chemical
stability and the use of waste heat can be considered as the regeneration temperature
is low at around 80 °C. This concept has been tested in a bench-scale absorption
column (2.5 cm diameter with structured packing) and stirred tank regenerator [45].
Three solvent combinations were studied with regeneration temperatures from 50 to
95 °C resulting in energy consumption reductions of more than 35% compared to
the conventional MEA based solvent process (Fig. 6).
112 K.A. Mumford et al.

Fig. 6 Process flow diagram of a thermomorphic biphasic solvent (TBS) system [45]

3.2.3 Mixed Amines—DEEA/MAPA System

Amine blends and phase change solvents offer promising improvements for the
CO2 absorption process [25]. When some amine solvents are mixed together they
can form two phases after CO2 absorption and the cyclic loading can be much
higher than the standard MEA process. An example of an amine blend with these
properties occurs when 2-(diethylamino) ethanol (DEEA) and 3-(methylamino)
propylamine (MAPA)are mixed together. DEEA is a tertiary alkanolamine and
MAPA has two amine functional groups (primary and secondary). The amine
mixture will lower the overall heat of absorption while the primary/secondary
amine will improve the capture rate and the tertiary amine will improve the capture
capacity of the solvent. A 5M DEEA/2M MAPA blend will form two phases upon
loading with CO2 and after separation nearly all the CO2 is present in the lower
phase which contains the MAPA. Only this heavy phase needs to be sent to the
regenerator leading to lower operational costs for stripping CO2 from the solvent.
Arshad et al. studied the performance of this solvent blend and reported that it has a
lower regeneration energy at similar temperature compared to the standard MEA
process [3, 4]. A pilot plant at the Gloshaugen (NTNU/SINITEF) facility has been
used to test this solvent blend for post-combustion CO2 capture [29]. A process
flow diagram of this pilot plant can be found in Fig. 7. The pilot plant performed
well with no issues with the high viscosity or foaming. Due to the presence of the
Phase Change Solvents for CO2 Capture Applications 113

Fig. 7 Process flow diagram of the Gloshaugen pilot plant testing the DEEA/MAPA solvent
process [29]

tertiary amine, absorption of CO2 was relatively fast, CO2 stripping was easier and
the reboiler was operated at a lower specific reboiler duty and temperature than the
standard 30 wt% MEA process. Issues that need to be considered with these
blended amine process are solvent degradation and solvent volatility in order to
avoid further environmental issues and reduce solvent replacement requirements.
Other blends of amines that result in phase change solvents have been proposed,
including 2 M 1,4-butanediamine (BDA) and 4 M 2-(diethylamino)-ethanol
(DEEA) by Xu et al. [43].
Another non-aqueous amine solvent process, called the “Self-Concentrating
Absorbent CO2 Capture Process” is being developed by 3H Company [26]. When
the proprietary amine solvent mixture reacts with CO2, the mixture separates into
two distinct phases: a CO2-rich liquid phase and a dilute lean phase. The process is
proposed to reduce the total regeneration energy by as much as 70%. The solvent
volume required would be lower than an MEA system as the solvent has a high
working capacity which also results in lower pumping requirements, lower auxil-
iary power demands, and reduced equipment size [26]. As the solvent is
non-aqueous corrosion problems are also minimised. Demonstration of this process
is proposed at an E-ON power plant in the United States as a next stage of com-
mercialization development (Fig. 8).
114 K.A. Mumford et al.

Fig. 8 Self-concentrating amine absorbent process concept [26]

4 Conclusions

This chapter provides an update on recent developments within a rapidly devel-


oping field, namely Phase Change Solvents for CO2 capture. This has included both
solid-liquid and liquid-liquid system developments within academia and industry. It
was shown that the benefit of these energy efficient solvent systems over traditional
liquid only systems stems from their potential to separate the CO2 rich phase from
the CO2 lean phase, thus improving solvent capacities, reaction kinetics and
reducing energy penalties.

References

1. Alstom (2007) Chilled ammonia-based wet scrubbing for post-combustion CO2 capture. In:
401/021507 DNRN (ed)
2. Anderson C, Harkin T, Ho M et al (2013) Developments in the CO2CRC UNO MK 3
process: a multi-component solvent process for large scale CO2 Capture. In: Dixon T,
Yamaji K (eds) Ghgt-11. Elsevier Science, Amsterdam, pp 225–232
3. Arshad MW, Fosbøl PL, Von Solms N et al (2013) Freezing point depressions of phase
change CO2 solvents. J Chem Eng Data 58:1918–1926
4. Arshad MW, Von Solms N, Thomsen K et al (2013) Heat of absorption of CO2 in aqueous
solutions of DEEA, MAPA and their mixture. Energy Procedia, 1532–1542
5. Barzagli F, Mani F, Peruzzini M (2013) Efficient CO2 absorption and low temperature
desorption with non-aqueous solvents based on 2-amino-2-methyl-1-propanol (AMP). Int J
Greenh Gas Control 16:217–223
6. Cullinane JT, Rochelle GT (2004) Carbon dioxide absorption with aqueous potassium
carbonate promoted by piperazine. Chem Eng Sci 59:3619–3630
7. Fernandes D, Conway W, Burns R et al (2012) Investigations of primary and secondary
amine carbamate stability by H-1 NMR spectroscopy for post combustion capture of carbon
dioxide. J Chem Thermodyn 54:183–191
Phase Change Solvents for CO2 Capture Applications 115

8. Fernandez ES, Goetheer ELV (2011) DECAB: process development of a phase change
absorption process. 10th international conference on greenhouse gas control technologies.
Vol 4, pp 868–875
9. Fernandez ES, Heffernan K, Van Der Ham LV et al (2013) Conceptual design of a novel
CO2 capture process based on precipitating amino acid solvents. Ind Eng Chem Res 52:
12223–12235
10. Gal E (2008) ultra cleaning of combustion gas including the removal of Co2. In:Google
Patents
11. Gal E, Jayaweera I (2010) Chilled ammonia based CO2 capture system with water wash
system. In: Google Patents
12. Goetheer ELV, Nell L (2009) First pilot results from TNO’s solvent development workflow.
Carbon Capture J 8:2–3
13. Hasib-Ur-Rahman M, Siaj M, Larachi F (2012) CO2 capture in alkanolamine/
room-temperature ionic liquid emulsions: a viable approach with carbamate crystallization
and curbed corrosion behavior. Int J Greenh Gas Control 6:246–252
14. Hook RJ (1997) An investigation of some sterically hindered amines as potential carbon
dioxide scrubbing compounds. Ind Eng Chem Res 36:1779–1790
15. Hooper B, Stevens G, Kentish S et al (2011) A process and plant for removing acid gases
16. Kumar PS, Hogendoorn JA, Feron PHM et al (2003) Equilibrium solubility of CO2 in
aqueous potassium taurate solutions: part 1. crystallization in carbon dioxide loaded aqueous
salt solutions of amino acids. Ind Eng Chem Res 42:2832–2840
17. Lee A, Wolf M, Kromer N et al (2015) A study of the vapour-liquid equilibrium of CO2 in
mixed solutions of potassium carbonate and potassium glycinate. Int J Greenh Gas Control
36:27–33
18. Lombardo G, Agarwal R, Askander J (2014) Chilled ammonia process at technology center
Mongstad—first results. Energy Procedia 51:31–39
19. Majchrowicz ME, Brilman DWF (2012) Solubility of CO2 in aqueous potassium L-prolinate
solutions-absorber conditions. Chem Eng Sci 72:35–44
20. Majchrowicz ME, Brilman DWF, Groeneveld MJ (2009) Precipitation regime for selected
amino acid salts for CO(2) capture from flue gases. Greenhouse Gas Control Technol
9(1):979–984
21. Misiak K, Sanchez CS, Van Os P et al (2013) Next generation post- combustion capture:
combined CO2 and SO2 removal. Energy Procedia 37:1150–1159
22. Moene R (2013) CCS—innovation in shell (innovation is bringing an insightful ideal
successfully to the market). In: CLIMIT SUMMIT 2013. Oslo
23. Moene R, Schoon L, Van Straelen J et al (2013) Precipitating carbonate process for energy
efficient post-combustion CO2 capture. Energy Procedia 37:1881–1887
24. Moene R, Schoon L, Van Straelen J et al (2013) Precipitating carbonate process for energy
efficient post-combustion CO2 capture. Ghgt-11 37:1881–1887
25. Monteiro JGMS, Pinto DDD, SaH Zaidy et al (2013) VLE data and modelling of aqueous N,
N-diethylethanolamine (DEEA) solutions. Int J Greenhouse Gas Control 19:432–440
26. Mosser M (2012) Post-combustion CO2 capture for existing PC boilers by self-concentrating
amine absorbent. In: National Energy Technology Laboratory
27. National Energy Technology Laboratory (2013) DOE/NETL advance carbon dioxide capture
R&D program: technology update. In: Department of Energy, p 716
28. Pennenergy (2014) Clean coal: carbon capture pilot begins at Polk IGCC plant
29. Pinto DDD, Knuutila H, Fytianos G et al (2014) CO2 post combustion capture with a phase
change solvent. Pilot plant campaign. Int J Greenhouse Gas Control 31:153–164
30. Raynal L, Bouillon PA, Gomez A et al (2011) From MEA to demixing solvents and future
steps, a roadmap for lowering the cost of post-combustion carbon capture. Chem Eng J
171:742–752
31. Sanchez-Fernandez E, De Miguel Mercader F, Misiak K et al (2013) New process concepts
for CO2 capture based on precipitating amino acids. Energy Procedia
116 K.A. Mumford et al.

32. Smith K, Harkin T, Mumford K et al (2016) Outcomes from a precipitating potassium


carbonate solvent absorption pilot plant for CO2 capture from a brown coal fired power station
in Australia. Fuel Processing Technology. doi:10.1016/j.fuproc.2016.08.008
33. Smith K, Lee A, Mumford K et al (2015) Pilot plant results for a precipitating potassium
carbonate solvent absorption process promoted with glycine for enhanced CO2 capture. Fuel
Process Technol 135:60–65
34. Smith K, Xiao G, Mumford K et al (2013) Demonstration of a concentrated potassium
carbonate process for CO2 capture. Energy Fuels 28:299–306
35. Svendsen HF, Tobiesen FA, Mejdell T et al (2008) Method for capturing Co2 from exhaust
gas. In: Google Patents
36. Svensson H, Hulteberg C, Karlsson HT (2014) Precipitation of AMP carbamate in CO2
absorption process. Energy Procedia 63:750–757
37. Telikapalli V, Kozak F, Francois J et al (2011) CCS with the Alstom chilled ammonia process
development program—field pilot results. 10th international conference on greenhouse gas
control technologies 4:273–281
38. Thee H, Nicholas NJ, Smith K et al (2014) A kinetic study of CO2 capture with potassium
carbonate solutions promoted with various amino acids: glycine, sarcosine and proline. Int J
Greenh Gas Control 20:212–222
39. Thee H, Suryaputradinata Y, Mumford KA et al (2012) A kinetic and process modeling study
of CO2 capture with MEA-promoted potasium carbonate solutions. Chem Eng J 210
40. Van Straelen JPT (2011) Process for the removal of carbon dioxide from a gas. In: Google
Patents
41. Van Straelen JPT (2015) Process for the removal of carbon dioxide from a gas. In: Google
Patents
42. Versteeg GF, Kumar PS, Hogendoorn JA et al (2011) Method for absorption of acid gases. In:
Google Patents
43. Xu Z, Wang S, Qi G et al (2014) CO2 absorption by biphasic solvents: comparison with lower
phase alone. Oil Gas Sci Technol 69:851–864
44. Zhang J, Nwani O, Tan Y et al (2011) Carbon dioxide absorption into biphasic amine solvent
with solvent loss reduction. Chem Eng Res Des 89:1190–1196
45. Zhang J, Qiao Y, Agar DW (2012) Improvement of lipophilic-amine-based thermomorphic
biphasic solvent for energy-efficient carbon capture. Energy Procedia 23:92–101
46. Zheng SD, Tao MN, Liu Q et al (2014) Capturing CO2 into the precipitate of a
phase-changing solvent after absorption. Environ Sci Technol 48:8905–8910
Aqueous Amino Acid Salts and Their
Blends as Efficient Absorbents for CO2
Capture

Azmi Mohd Shariff and Muhammad Shuaib Shaikh

Abstract The increase in global population and industrialization has led to an


increase in global energy consumption exponentially. Over 85% of global energy is
supplied by burning fossil fuel, which releases large volume of CO2 emissions in the
atmosphere. Increasing of CO2 emissions is the major cause for the catastrophic
climate change, which has led to increased demand for efficient and effective CO2
capture. CO2 absorption by chemical solvents is the most widely used technique
commercially nowadays. Alkanolamine solvents such as monoethanolamine
(MEA) and methyldiethanolamine (MDEA) are the most commonly used absorbents
for CO2 removal from various gas streams. However, it is well known that these
solvents suffer from variety of drawbacks such as limited CO2 loading capacity,
equipment corrosion, toxic nature and highly volatile. Moreover, these absorbents
are easily degradable, require high regeneration energy, and cause flooding problems
in the operation. Therefore, better and efficient solvents should be searched for the
removal of CO2 from exhaust gas streams. Aqueous amino acid salts and their blends
are the promising solvents for CO2 capture as compared to alkanolamine. In this
chapter, amino acid salts and their blends are introduced and their performance
analysis as potential solvents for commercial possibilities are discussed. Based on
the analysis, these absorbents show superior performance as an alternative to the
conventional alkanolamines for CO2 capture. These solvents are environmental
friendly with higher CO2 loading capacity, faster reaction kinetics and require less
regeneration energy compares to the commercial amines. Besides, these solvents are
non-volatile, less corrosive and oxidative stable. Moreover, aqueous amino acid salts
are more effective by blending with additives such as piperazine.

A.M. Shariff (&)  M.S. Shaikh


Research Centre for CO2 Capture (RCCO2C), Department of Chemical Engineering,
Universiti Teknologi PETRONAS, 31750 Tronoh, Perak, Malaysia
e-mail: azmish@petronas.com.my

© Springer International Publishing AG 2017 117


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_6
118 A.M. Shariff and M.S. Shaikh

Nomenclature
Abbreviations
%wt. Percent by weight
AAAS Aqueous amine amino acid salt
AAS Amino acid salt
ALA Alanine
AMP 2-amino-2-methyl-1-propanol
ARG Arginine
ARG Arginine
ASN Asparagine
ASP Aspartate
ASTM American society for testing and materials
CAPEX Capital expenditure
CASPER CO2 capture and sulfur precipitation for enhanced removal
CO2 Carbon dioxide
CYS Cysteine
DEA Diethanolamine
DGA Diglycolamine
DIPA Diisopropylamine
GHG Greenhouse gas
GLN Glutamine
GLU Glutamate
GLY Glycine
H 2S Hydrogen sulfide
HIS Histidine
ILU Isoleucine
K-AABA Potassium salt of DL-a-amino butyric acid
K-ALA Potassium salt of alanine
K-ASN Potassium salt of L-asparagine/asparaginate
K-BALA Potassium salt of b-alanine
K-DiMGLY Potassium salt of diethyl or dimethylglycine
K-GLU Potassium salt of glutamate
K-GLY Potassium salt of glycine
K-LYS Potassium salt of lysine
kPa Kilo pascal
K-PRO Potassium salt of proline
K-SAR Potassium salt of sarcosine
K-SER Potassium salt of serine
K-TAU Potassium salt of taurine
K-THR Potassium salt of threonine
LEU Leucine
Li-PRO Lithium salt of proline
Li-SAR Lithium salt of sarcosine
LYS Lysine
Aqueous Amino Acid Salts and Their Blends … 119

MDEA N-methyldiethanolamine
MEA Monoethanolamine
MET Methionine
MET Methionine
Na-ALA Sodium salt of alanine
NA-BALA Sodium salt of b-alanine
Na-GLY Sodium salt of glycine
Na-PH Sodium phenolate
Na-PRO Sodium salt of proline
Na-SAR Sodium salt of sarcosine
Na-SO3 Sodium sulfite
Na-TAU Sodium salt of taurine
Na-VO3 Sodium metavanadate
NH3 Ammonia
NOAA National Oceanographic and Atmospheric Administration
OPEX Operating expenditure
pH Power of hydrogen ion
PHE Phenylalanine
PostCap Post combustion capture technology
ppm Parts per million
PRO Proline
PZ Piperazine
SARMAPA Sarcosine with 3-(methylamino propylamine)
SER Serine
SO2 Sulphur dioxide
TEA Triethanolamine
THR Threonine
TIPA Tri-isopropanolamine
TRP Tryptophan
TYR Tyrosine
VAL Valine
VLE Vapor liquid equilibrium

Units and Symbols


Ea Activation energy (Kg/mol)
k2 Forward second order reaction rate (m3 mol-1S-1)
kov Overall reaction rate constant (S-1)
LD50 Lethal dose (mg/Kg)
M Molarity (mol/litre)
T Temperature (K/°C)
a Loading (mol/mol)
120 A.M. Shariff and M.S. Shaikh

1 Introduction

The energy demand is increasing worldwide after an industrial revolution. About


85% of global energy demand is fulfilled by burning fossil fuels to generate
electricity in power plants [1, 2]. Therefore, the huge amount of greenhouse gases
(GHG) are released in the atmosphere, causing a global environmental problem,
termed as global warming. Carbon dioxide (CO2) is the main contributor to global
warming [2, 3]. According to NOAA, 2014, the atmospheric CO2 concentration is
396.21 ppm, which exceeds the tolerable limit (350 ppm). It is expected that the
CO2 concentration may rise beyond 400 ppm by the year 2016, if the suitable
mitigation techniques are not established [4]. Based on the rising level of CO2 in the
atmosphere, the average temperature may increase globally from 1.4 to 5.8 °C by
the year 2100, according to the different climate prediction models. The harmful
effects of global warming includes the rise of sea level, melting of glaciers and ice
caps, climate change, and infrastructure destruction [5]. Hence, it has become a
global panic to lessen the CO2 emissions by introducing the effective technologies
[6]. In this connection, a lot of research work has been carried out worldwide. As
the results, various technologies have been established such as absorption,
adsorption, membrane and cryogenic processes [7–9]. Among these, the most
extensively functioning technology is absorption by chemical solvents due to the
great deal of research carried out on liquid solvents and their practical applicability
[2, 10–14].
The conventional chemical solvents used commercially for CO2 absorption are
alkanolamines (also called as amines) such as monoethanolamine (MEA), dietha-
nolamine (DEA), diisopropanolamine (DIPA), triethanolamine (TEA), methyl-
diethanolamine (MDEA) [15–18]. Though these alkanolamines have been used
widely for CO2 absorption, however, several shortcomings have been identified and
reported [19, 20]. The problems concerned with these solvents include shorter life
due to poor resistance to oxidative and thermal degradation. Corrosion of equip-
ment and flow lines, loss of solvent due to high vapor pressure, higher energy of
regeneration and toxicity are among the key problems allied with such solvents.
Moreover, these solvents have inadequate cyclic CO2 loading capacity, except for
tertiary amines such as MDEA, which has higher cyclic capacity compared to
conventional MEA [18, 21–25]. These drawbacks are summarized in Table 1 with
their effects. The shortcomings offered by these solvents confine their use for
commercial applications [20, 26]. Therefore, it is mandatory for the scientists and
researchers to discover the new solvent systems, which can compensate the
shortcomings of the existing ones.
The amino acid salt solutions have emerged as one of the promising solvents for
CO2 capture as compared to amines. Amino acid salt solutions are environmental
friendly with higher CO2 loading capacity, faster reaction kinetics and require less
regeneration energy. Besides, these solvents are non-volatile, non-flammable, less
Aqueous Amino Acid Salts and Their Blends … 121

Table 1 Drawbacks of amines and their possible effects [15, 18, 21–25, 28–30, 32, 96, 116]
Drawbacks of amines Effects
Limited CO2 loading • High volume of solvent required
(Except tertiary amines i.e., MDEA, which • Increase the overall solvent cost
has better loading than MEA, but still less • Increase the size of the plant equipment and
than amino acid salts) accessories
• Overall increase in CAPEX and OPEX
Low cyclic capacity • Leads to high circulation rate of the solvent
(Except MDEA) • Increase the energy consumption
• Increase the dimensions of the process
equipment
• Increase the CAPEX and OPEX
High regeneration energy • Increase the operating cost due to high
(Except MDEA which require less energy requirements
regeneration energy than MEA, but still
higher than amino acid salts)
High circulation rate • Increase the dimensions of equipment
(solvent heat exchangers, amine pumps,
absorber etc.)
• High energy consumption in plant operation
Thermal and oxidative degradation • Increase solvent replacement costs
(Except MDEA, which has less degradation • Shorter life of the solvent, thus increase the
than MEA) CAPEX
Low absorption rate • Increase the size of the absorber and other
plant accessories
• Require large volume of solvent
• Increase the CAPEX and OPEX
Formation of heat stable salts • Enhance the corrosion of equipment
• Require extra cost for corrosion repair
Low surface tension • Cause flooding and foaming
• Lower the mass transfer efficiency
Corrosive • Decrease equipment and accessories life
• Increase the cost required for restoration of
corroded equipment
• Increase maintenance cost due to unplanned
downtime
Toxic • Health and environmental risks
High volatility and flammable • High vapor loss of the solvent
• Increase the solvent replacement cost
• Increase the overall CAPEX and OPEX
• Fire and other related hazards due to
flammable nature

corrosive and oxidative stable [27–31]. Several studies have been conducted to
authenticate the performance of these solvents. In this chapter, the performance
analysis of various aqueous amino acid salts and their blends have been carried out
based on the literature data for evaluating their commercial possibilities.
122 A.M. Shariff and M.S. Shaikh

2 Chemistry

The amino acids are known as amphiprotic in nature, which can either donate or
accept the proton. Thus, the amino acids can act as acid or a base, therefore, called as
ampholytes. The structures of amino acids contain at least one basic carboxyl or
sulphonyl group and one acid amino group. In an aqueous solution, amino acids
exist as a zwitterion (form II in reaction 1) in the absence of the other solutes. The
compounds exist as a zwitterion when they are positively and negatively charged due
to presence of separate functional groups. When the acid is added to the amino acid
solution, the zwitterion picks up the proton (form I in reaction 1), while the addition
of base to amino acid solution removes a proton from the ammonium group and
leaves the molecule with final negative charge (form III in reaction 1). The zwit-
terion and deprotonated form of amino acid is shown in reaction 1 [27, 32].
H þ H þ
HO2 CRNH3þ $  O2 CRNH3þ $  O2 CRNH2
ð1Þ
ðIÞ ðIIÞ ðIIIÞ

At this stage, the amino acid solution contains deprotonated amino group, which
is actually reactive with acid gases such as CO2. There are twenty types of amino
acids, which are listed in Table 2 along with corresponding chemical formula and
linear structure.
Table 2 Types of amino acids [113]
Amino acid Chemical Linear structure Molecular
formula weight
(g/mol)
Alanine (ALA) C3H7NO2 CH3–CH(NH2)–COOH 89.0935
Arginine (ARG) C6H14N4O2 HN=C(NH2)–NH–(CH2)3–CH 174.2017
(NH2)–COOH
Asparagine (ASN) C4H8N2O3 H2N–CO–CH2–CH(NH2)–COOH 132.1184
Aspartate (ASP) C4H7NO4 HOOC–CH2–CH(NH2)–COOH 133.1032
Cysteine (CYS) C3H7NO2S HS–CH2–CH(NH2)–COOH 121.1590
Glutamate (GLU) C5H9NO4 HOOC–(CH2)2–CH(NH2)–COOH 147.1299
Glutamine (GLN) C5H10N2O3 H2N–CO–(CH2)2–CH(NH2)–COOH 146.1451
Glycine (GLY) C2H5NO2 NH2–CH2–COOH 75.0669
Histidine (HIS) C6H9N3O2 NH–CH=N–CH=C–CH2–CH 155.1552
(NH2)–COOH
Isoleucine (ILE) C6H13NO2 CH3–CH2–CH(CH3)–CH(NH2)– 131.1736
COOH
Leucine (LEU) C6H13NO2 (CH3)2–CH–CH2–CH(NH2)–COOH 131.1736
Lysine (LYS) C6H14N2O2 H2N–(CH2)4–CH(NH2)–COOH 146.1882
Methionine C5H11NO2S CH3–S–(CH2)2–CH(NH2)–COOH 149.2124
(MET)
(continued)
Aqueous Amino Acid Salts and Their Blends … 123

Table 2 (continued)
Amino acid Chemical Linear structure Molecular
formula weight
(g/mol)
Phenylalanine C9H11NO2 Ph–CH2–CH(NH2)–COOH 165.1900
(PHE)
Proline (PRO) C5H9NO2 NH–(CH2)3–CH–COOH 115.1310
Serine (SER) C3H7NO3 HO–CH2–CH(NH2)–COOH 105.0930
Threonine (THR) C4H9NO3 CH3–CH(OH)–CH(NH2)–COOH 119.1197
Tryptophan (TRP) C11H12N2O2 Ph–NH–CH = C–CH2–CH(NH2)– 204.2262
COOH
Tyrosine (TYR) C9H11NO3 HO–Ph–CH2–CH(NH2)–COOH 181.1894
Valine (VAL) C5H11NO2 (CH3)2–CH–CH(NH2)–COOH 117.1469

3 CO2 Capture

Although the alkanolamines are very popular and widely used in industries for acid
gas removal from variety of sour gas streams, however, the drawbacks listed in
Table 1 restrict their applications for CO2 removal. The major downsides of
alkanolamines such as high-energy consumption and high process cost stimulate the
interest to search another type of solvents for CO2 capture that would minimize the
energy usage and the cost.
Amino acid salt solutions as effective substitutes have been investigated
extensively for the removal of CO2. The reactivities of amino acid salts are similar
to that of amines because of having identical functional groups in their molecules
[27, 28, 30]. However, based on the various attractive and valuable characteristics
identified by extensive research on these systems, amino acid salt solutions have
been proposed as the potential alternative to amine-based systems [20]. They have
the adequate life cycle because of possessing good ability to resist thermal and
oxidative degradation [27–30]. The presence of ionic structure in amino acid salt
systems lowers the volatility to almost negligible value, which prevent their loss
even at high-temperature operations. Moreover, these solvents are able to regen-
erate, expected to be less corrosive, eco-friendly and easily available at the com-
mercial level. Another benefit of amino acid salts is high surface tension, which
makes such solvents appropriate for membrane gas absorption system [27, 29, 30,
33–35]. The summary of the benefits of amino acid salts as potential solvents for
CO2 removal is given in Table 3 with their possible industrial impact.
The CO2 reacts with the amino acid salt solutions through a zwitterion mech-
anism because of the similar functional group as amines [27]. The reaction of CO2
with amino acid salt solution is shown in Fig. 1.
124 A.M. Shariff and M.S. Shaikh

Table 3 Benefits of amino acid salts and their potential impact [27–30, 32–35, 39–42]
Benefits of amino acid salts Potential impact
High CO2 loading • Less volume of solvent required
• Lower solvent circulation rates
• Reduce the equipment size
• Decrease the CAPEX and OPEX
Improved cyclic capacity • Reduce circulation rates of the solvent
• Decrease the energy consumption
• Reduce the equipment and accessories size
• Decrease the CAPEX and OPEX
Low circulation rate • Reduce the dimensions of equipment (solvent heat
exchangers, amine pumps, absorber etc.)
• Less electricity required for pumps and other equipment
• Overall cost reduction
Resistant to thermal and • Reduce solvent replacement costs
oxidative degradation • Increase solvent life and decrease the operating cost
High absorption rate • Decrease the size of the absorber and other plant accessories
• Require less volume of the solvent
• Reduce the CAPEX and OPEX
High surface tension • Prevent flooding and foaming
• Enhance mass transfer rates and efficiency
Less corrosive • Increase equipment and accessories life
• Reduce the cost required for restoration of corroded system
• Reduce overall cost
Less viscous • Increase the pumping efficiency
• Enhance mass transfer
• Reduce pumping cost
Environmental friendly • Minimum health and environmental effects
Negligible volatility and not • No vapor loss and no fire hazards
flammable • Reduces the solvent replacement cost
• Reduces over all CAPEX and OPEX

Fig. 1 Reaction of CO2 with aqueous amino acid salt [27]

In the reaction scheme shown in Fig. 1, initially, the zwitterion and carbamate
are formed during the absorption reaction. Later, the carbamate undergoes
hydrolysis, which results in the formation of deprotonated amino acid and
bicarbonate/carbonate depending upon the pH of the solution. The deprotonated
amino acid later can react with CO2. The extent of carbamate hydrolysis can be
Aqueous Amino Acid Salts and Their Blends … 125

determined by certain parameters such as concentration, carbamate stability and pH


of the solution [36–38]. The reaction equilibria shown in Fig. 1 favors the for-
mation of carbamate and bicarbonate at low temperature conditions. However, at
high temperatures, this equilibria favors the liberation of amino acid and CO2.

4 Absorption Processes

The first amino acid used for the removal for CO2 was glycine (GLY) in
Giammarco-Vetrocoke process. In this process, GLY was used as an activator in the
alkali carbonate solution to improve the CO2 removal performance. As a result,
better absorption performance was found together with less amount of steam
required for solvent regeneration as compared to the amines [15]. Similarly, another
process based on amino acid salts known as Alkacid process, has been developed
by BASF. There are three major process variations in this process, Alkacid “M”,
Alkacid “dik” and Alkacid “S”. In these process variations, Alkacid “M” uses the
sodium salt of alanine (Na-ALA) for the absorption of CO2 or H2S, Alkacid “dik”
utilizes the potassium salts of diethyl or dimethylglycine (K-DiMGLY) for the
selective removal of H2S from the gas streams containing CO2. Alkacid “S” uses
sodium phenolate (Na-PH) solution as an absorbent for the removal of the con-
taminants other than CO2 and H2S such as ammonia, carbon disulfide, mercaptan,
dust and tars. These absorbents showed better absorption capacity and required less
steam for solvent regeneration as compared to alkanolamine (DEA, MDEA).
Moreover, these solvents also showed better stability and less corrosiveness [15].
Another two processes were developed by TNO (The Netherlands) for separation of
CO2 utilizing aqueous amino acid salts as reactive absorbents. In the first process,
membrane gas absorption was used based on the commercial polypropylene hollow
fiber membrane. The conventional solvents like alkanolamines were not compatible
enough with these membranes. The new solvents (CORAL liquids) based on
potassium salts of amino acids such as taurine (K-TAU) and glycine (K-GLY) have
shown better operational performance with these membrane modules owing to the
high surface tension of amino acid salts solutions compared to the amines. Besides,
these new absorbents show better resistance towards thermal and oxidative
degradation and negligible vapor pressure [43, 44]. Another proprietary
post-combustion carbon capture technology (PostCap) has been developed by
Siemens AG. The technology uses aqueous amino acid salt solutions, but the type
of amino acid salt is undisclosed. The use of the solvent developed by Siemens AG
showed that the process requires less energy with enhanced performance in the CO2
capture process. The amino acid based solvent (undisclosed) showed high chemical
stability, environmentally friendly, and flexible operation with good economics of
the process under a wide range of operating parameters. This process was also
validated in a coal-fired pilot scale power plant in Frankfurt, Germany, while their
second pilot plant is under the planning phase in Florida, USA [45]. Another
process known as DECAB process has been developed by TNO (The Netherlands)
126 A.M. Shariff and M.S. Shaikh

using alkaline salts of amino acids. The process takes an extra benefit of the fact
that when the CO2 absorbs in aqueous salts of amino acids, the precipitation forms
under certain conditions in the process such as solvent concentration, CO2 loading,
temperature, etc. [43, 46]. These precipitations can be the amino acid itself
according to Kumar et al. [47]. Due to the formation of precipitation, the CO2
equilibrium partial pressure will remain nearly constant at certain CO2 loading.
Therefore, a possibility of higher loadings of the solvent may be achieved with the
benefits for the solvent circulation and requirement of energy for regeneration.
Moreover, the absorber size could essentially be reduced as the higher driving force
is observed. However, it does not necessarily follow that the precipitation leads to
an increase in the driving force for the reaction, as the liquid nevertheless remains
saturated. The precipitation of the amino acid itself would represent a loss of
reactant [47]. In this CO2 removal process, spray tower was used instead of the
other contactors to handle the possible blockage of the equipment [43, 46]. Based
on the above analysis, it can be said that, the formation of precipitation during CO2
capture process is one of the drawback of amino acid salts, if these are used in
ordinary absorption columns/contactors. Moreover, the formation of precipitation
could potentially block the equipment, pumps and other accessories in the
absorption plant, which could lead to the plant shut down [43, 46]. Moreover, the
formation of precipitation also depends on the type of amino acids and their
respective concentration. The higher concentrations of amino acid salts cannot be
used due to the formation of precipitation during CO2 absorption. This is one of the
limitations of the amino acid salts [43, 51]. The summary of the CO2 absorption
processes developed by various academic and commercial entities is given in
Table 4.

Table 4 Processes for acid gas removal using amino acid salts [32]
Name of the process Amino acid salts used Development Refs.
stage
Giammarco-Vetrocoke • GLY as activator in alkali Commercial [15]
process carbonates
Alkacid process developed • Na-ALA Commercial [15]
by: BASF • K-DiMGLY
• Na-PH
MGA process developed • Alkaline salts of TAU and Commercial [43, 44]
by: TNO GLY (CORAL liquids)
Post Cap process • Potassium salts of amino acids Commercial [45]
developed by: Siemens AG (types: not disclosed)
DECAB process • CORAL liquids Development [43, 46]
developed by: TNO
DECAB plus process • Solvents under development Development [48]
developed by: TNO
CASPER process • Aqueous K-BALA Development [49]
developed by: iCAP
Aqueous Amino Acid Salts and Their Blends … 127

In recent times, TNO has developed another process called as DECAB Plus
which achieves an extra driving force for stripping of CO2. This is achieved by
reduction in pH values due to the formation of precipitates [48]. Another process
known as CASPER (CO2-capture And Sulphur Precipitation for Enhanced
Removal) recently developed by iCAP consortium applies precipitation in CO2
saturated aqueous potassium salt of b-alanine (K-BALA) for the simultaneous
removal of CO2 and SO2 from flue gas streams. The CASPER solvent achieved
robust improvements in the removal of CO2 and SO2 from the flue gases [49].

5 Performance for CO2 Capture

The performance of amino acid salts for CO2 capture is dependent on various
factors. There are so many studies available in literature on amino acid salts for the
removal of CO2 from various gas streams. These studies include the solubility of
CO2 in amino acid salts, kinetic study, thermal degradation, corrosion study, tox-
icity analysis, energy and other considerations. It is very essential to assess the
solvent performance based on the studies conducted by various researchers in order
to evaluate their commercial possibilities.

5.1 Solubility of CO2

The solubility of CO2 in aqueous amino acid salt solutions provides the information
on chemical solubility and loading capacity of acid gases (CO2/H2S). For the
efficient design of CO2 removal process using new formulated solvents, it is very
essential to assess the CO2 loading of the solvent at various temperatures and
pressures for better process performance. CO2 loading data, therefore, helps in the
selection of appropriate solvent for commercial applications [32, 47]. Various
studies have been conducted on solubility of CO2 in aqueous amino acid salt
solutions, as summarized in Table 5; however, there is a scarcity of the systematic
information on comprehensive assessment and comparison of these solvents. In this
section of the chapter, the solubility of CO2 in various amino acid salt solutions has
been briefly reviewed and discussed systematically in order to make it easy to
assess the solvent for commercial applications.
Kumar et al. [47] reported the equilibrium solubility of CO2 in aqueous K-TAU
solutions at 298 and 313 K in low CO2 partial pressure range (0.1–6.0 kPa). The
range of concentration studied was 500–4000 mol m−3. The crystallization was
observed during the CO2 absorption at a higher range of CO2 partial pressures and
at high amino acid salt concentration (2 M and above). Furthermore, the influence
of crystallization on the vapor liquid equilibria was also explained. It was found that
even at a low range of CO2 loadings (0.2–0.5 mol of CO2/mol of AAS); the
crystallization occurred, depending on the further amino acid salt’s concentration. It
128 A.M. Shariff and M.S. Shaikh

Table 5 Various studies on CO2 solubility in amino acid salts


Aqueous amino acid salts Concentration Temperature (K) Pressure Refs.
K-TAU 500–4000 mol/m3 298–313 0.1–6.0 kPa [47]
Na-GLY 10–30 wt% 303.15–323.15 0.1–200 kPa [50]
Na-GLY 1–30 wt% 298.15–313.15 100–2500 kPa [51]
K-GLY 0.1–3.0 mol/dm−3 293–351 6  104 Pa [23]
K-THR 1.0 mol/dm−3 313.15 6  104 Pa [23]
K-GLY 1–3 mol/L 298 – [52]
K-SER 14.3 mass% 313.15–373.15 0.1–400 kPa [54]
SARMAPA 1.0–5.0 M 313–373 4.08–99.1 kPa [53]
K-PRO 0.5–3.0 mol/dm−3 285–323 70 kPa [58]
K-ALA 2.5 M 298–313 Atmospheric [57]
K-PRO 2.5 M 298–313 Atmospheric [57]
K-SAR 4M 313.15–353.15 0–812.5 kPa [59]
K-GLY 1.0–1.8 mol/kg 313–393 10–100 kPa [60]
K-TAU 1.0–1.8 mol/kg 313–393 10–100 kPa [60]
K-PRO 7.5–27.5 wt% 313.2–353.2 1000 kPa [61]
K-AABA 6.9–25.6 wt% 313.2–353.2 1000 kPa [61]
K-LYS 0.5–2.5 mol/dm−3 298–313 0–45 kPa [62]
K-LYS 2.5 mol/L 313.15–333.15 0.05–17.5 kPa [63]
K-PRO 2.5 mol/L 313.15–333.15 0.05–17.5 kPa [63]
K-ASN 8.5–34% 313.2–353.2 Up to 950 kPa [64]
K-GLU 9.2–36.8% 313.2–353.2 Up to 950 kPa [64]

was reported that the crystallization product is the protonated amine or the zwit-
terionic form of the amino acid. The formation of crystallization was observed to
have the positive effect on the CO2 absorption capacity [47]. Similarly, Song et al.
[50] studied the solubility of CO2 in aqueous sodium glycinate (Na-GLY) at three
different temperatures (303.15, 313.15, and 323.15 K) at partial pressure range of
0.1–200 kPa followed by different concentrations (10, 20, 30 wt%). It was
observed that the solubility decreased with the rise of temperature and Na-GLY
concentration; however, with increasing the pressure the solubility increased. It was
concluded that the CO2 solubility in aqueous Na-GLY solution (10 wt%) is better
than various aqueous alkanolamine solvents such as MEA, AMP, and TIPA [50].
Likewise, another study on CO2 solubility in aqueous Na-GLY was conducted by
Harris et al. [51] over a wide range of concentrations at 298.15 and 313.15 K. The
study was conducted at slightly higher pressures up to 2500 kPa. The finding of this
study also shows that the Na-GLY has a better absorption capacity than aqueous
MEA and other alkanolamine solvents [51].
Portugal et al. [23] measured the solubility of CO2 in aqueous K-GLY solutions
at 293 to 353 K for the concentrations ranging from 0.1 to 3.0 mol dm−3. The study
was conducted up to the partial pressure of 6  104 Pa. The solubility showed
Aqueous Amino Acid Salts and Their Blends … 129

increasing trend with an increase in pressure. However, the effect of temperature on


solubility of CO2 in aqueous K-GLY solutions was not substantial from 293 to
323 K, which could be a limitation for the regeneration of the solvent. The
noticeable effect of temperature on the absorption capacity was observed above
323 K, and the effect was more obvious at 353 K. Furthermore, the solubility data
was also compared with the work of Song et al. [50] for aqueous Na-GLY, and it
was concluded that there is no significant difference between the absorption
capacity of aqueous Na-GLY and aqueous K-GLY at 313 and 323 at 1.06 mol
dm−3. The effect of concentration was observed to be same as previous authors. As
expected, the absorption capacity of aqueous solutions of K-GLY was found to be
more than that of conventional MEA when compared [23]. Subsequently, in the
same study, Portugal et al. [23] also investigated the solubility of CO2 in aqueous
potassium threonate (K-THR) for one concentration (1 mol dm−3) at 313 K. It was
concluded based on the results that aqueous K-GLY has better CO2 absorption
capacity as compared to aqueous K-THR for the same temperature and concen-
tration range [23]. Zhang et al. [52] conducted another study on CO2 absorption in
aqueous K-GLY for the concentration range of 1 to 3 mol L−1 at 298 K; and the
results were compared with aqueous MEA and MDEA. It was found that the CO2
absorption in aqueous K-GLY is higher than in aqueous MDEA but slightly lower
than aqueous MEA. The later findings of the Zhang et al. [52] do not comply with
the observations of Portugal et al. [23]. Additionally, the regeneration study of CO2
absorption in aqueous K-GLY was also carried out at 378 K and compared with
aqueous MEA and MDEA. It was concluded that aqueous K-GLY has improved
regeneration rate as compared to MEA but lesser than MDEA [52]. Aronu and
co-researchers [53] have studied the VLE measurement of CO2 in aqueous amine
amino acid salts (AAAS) solutions; 3-(methylamino) propylamine/sarcosine
(SARMAPA) at 40 to 100 °C for the concentration range of 1–5 M, and a total
pressure of 4.08–99.1 kPa. For 5 M loaded SARMAPA, the CO2 equilibrium was
measured at 40–120 °C. Moreover, the reaction heat between aqueous SARMAPA
(5 M) and CO2 was also estimated. The enthalpy of the present system was
79.2 kJ/mol of CO2 at loading of 0.2, which is slightly lower than 30 wt% MEA,
85.4 kJ/mol at 0.2 loading as reported by Lee et al. [55]. However, it is higher than
59.8 kJ/mol of MDEA at 0.1 loading for the concentration of 4.28 kmol m−3 [56].
Song et al. [54] have reported the solubility of CO2 in aqueous solutions of
potassium serinate (K-SER) at three different temperatures (313.15, 343.15, and
353.15 K) for only one concentration, 14.3 mass% over a partial pressure range of
0.1–400 kPa. The solubility of K-SER was also compared with 15.3 mass% MEA.
As a general trend, solubility of both solvents showed pressure dependent behavior,
however, the values of solubility reduced with increasing the temperature. It was
found that CO2 solubility in aqueous K-SER was much better than MEA at
313.15 K (absorber conditions) above partial pressure of 4.0 kPa. However, K-SER
solution has a lower solubility at 373.15 K than that of MEA, which suggested that
the cyclic capacity of CO2 in aqueous K-SER would be higher than that of MEA
[54]. Lim et al. [57] have studied two types of amino acid salts, i.e., potassium salt
of alanine (K-ALA) and proline (K-PRO) at 298 and 313 K for 2.5 M
130 A.M. Shariff and M.S. Shaikh

concentration. The trend of solubility with respect of temperature is same as found


by various researchers [50, 51]. The absorption capacity of the two amino acids in
this study was better than MEA. Besides, the heat of absorption of these two amino
acid salts was also determined and compared with MEA and DEA. It was found that
the aqueous K-ALA and K-PRO have lower absorption heat than MEA and DEA. It
was concluded that K-ALA is better than K-PRO in terms of capacity and heat of
absorption [57].
Majchrowicz and Brilman [58] have also measured the solubility of CO2 in
K-PRO for four different concentrations (0.5, 1.0, 2.0, and 3.0 mol dm−3) at 285
and 323 K and up to 70 kPa partial pressure. The effect of temperature, pressure
and amino acid salt concentration on CO2 solubility followed the same trend as
described previously. Moreover, the precipitation was also encountered at higher
concentration (3-mol dm−3) at 285 K over 6-kPa pressure. The effect of this pre-
cipitation was also determined, which showed that the higher CO2 loading can be
achieved at a lower partial pressure [58]. This finding also supported the work of
Kumar et al. [47]. The detailed understanding on the benefits of precipitations has
also been explained in DECAB process as mentioned in Sect. 4. The solubility of
CO2 in aqueous 1 mol dm−3 was compared with various other amino acid salts and
MEA. It was found that the aqueous K-TAU has lowest capacity while the K-THR
has a loading similar to MEA. The loading capacity of K-PRO and K-GLY is 19%
and 23% higher respectively than MEA. Moreover, the integral heat of absorption
for K-PRO is of the same magnitude as that of Na-GLY and other amines, but it is
less than MEA and slightly higher than MDEA. Thus, it was concluded that K-PRO
is a potential solvent for CO2 capture from industrial gases [58]. Aldenkamp et al.
[60] have reported the solubility of CO2 in K-TAU and K-GLY for two concen-
trations (1 and 1.8 mol kg−1) at absorber and desorber conditions (313, 333, 353,
373, and 393 K). The partial pressure range was studied as 10–100 kPa. The CO2
loading capacity of K-GLY was observed to be higher than K-TAU solutions at all
range of temperatures [60].
In a very recent time, Chang et al. [61] reported the solubility of CO2 in aqueous
K-PRO and DL-a-amino butyric acid (K-AABA) at various temperatures and
1000-kPa pressure. According to their findings, the solubility of both amino acid
salts is higher and comparable to the commercial alkanolamines. Moreover, it is
reported that K-AABA has higher loading capacity as compared to K-PRO [61].
Another recent work by Mazinani and his co-researchers [62] have reported the
solubility of CO2 in potassium lysinate (K-LYS) solution. The effect of tempera-
ture, concentration and pressure is same as reported in various other research works.
The results suggest that the amino acid salt studied can also be potential solvent for
CO2 capture [62]. Another study on K-LYS and K-PRO reported that the solubility
of CO2 in K-LYS and K-PRO is higher than that of MEA. Moreover, K-PRO has
lower loading as compared to K-LYS, therefore, K-LYS (2.5 M) is proposed for
CO2 absorption in place of 30 wt% MEA [63]. Chen et al. [64] have studied the
amino acid salts such as potassium L-asparaginate (K-ASN) and L-glutaminate
(K-GLU) up to 950-kPa pressure and at different temperatures. These amino acid
salts have significant CO2 loading and comparable to alkanolamines [64]. Increase
Aqueous Amino Acid Salts and Their Blends … 131

in gas solubility in liquids with increasing the pressure is universal. The CO2 partial
pressures in flue gases are generally less than 15 kPa-a, so the high-pressure studies
are not particularly relevant to the problem of CO2 capture from flue gases. The
high-pressure studies are significant for CO2 separation from natural gas. In addi-
tion, the reduction in solubility with increasing temperature is also universal, and
essential for the thermal swing recovery process. A requirement for a high des-
orption temperature by a particular amino acid (requiring a high steam pressure)
could be a distinct disadvantage to the power station supplying the steam.
There are various types of amino acid salts for the absorption of CO2 from
various gas streams reported by many groups of researchers. Even though, these
amino acid salts offer better absorption characteristics as compared to amine sol-
vents, but the improvements are still required to enhance the absorption charac-
teristics of amino acid salt solutions. The small improvements in the CO2
absorption performance could potentially benefit the overall process cost.
Therefore, the improvements are always required in the existing solvents to obtain
the more appropriate solvent for CO2 capture. To achieve these goals, various
researchers have proposed to use the solvent blends based on amino acid salts and
other solvents, including alkanolamines to combine all the favorable characteristics
in one solvent blend for sufficiently improving the absorption performance of the
CO2 capture process [59, 65–67]. Although the amino acids blended with alka-
nolamines perform better, but it could also add to some extent the drawbacks
mentioned in Table 1, depending upon the amount of the amine added to the amino
acid salt solutions.
Mazinani et al. [65] reported the solvent blend consists of aqueous Na-GLY and
MEA for CO2 removal. The solubility of CO2 in Na-GLY + MEA is higher than
MEA at a partial pressure above 20 kPa. The mixed solvent demonstrated the better
potential compared to MEA alone; however, the addition of MEA could offer to
some degree the limitations described above in Table 1 [65].
Lu et al. [67] proposed the solvents, which are blends of K-GLY + piperazine
(PZ) and K-GLY + phosphate (K3PO4). The CO2 loading of K-GLY + PZ and
K-GLY + K3PO4 blends is higher than a single K-GLY solvent. Kang et al. [59]
investigated the potassium sarcosinate (K-SAR), K-ALA + PZ blends and
K-SER + PZ blends. The results show that the blend of K-ALA + PZ has highest
solubility than the other solvent blends at 313.15 and 353.15 K. While at tem-
perature 353.15 K, the blends of K-ALA + PZ and K-SER + PZ showed better
solubility than K-SAR absorbent [59]. Privalova et al. [66] investigated three dif-
ferent amino acid salts for 15 wt% concentration blended with 5 wt% PZ at 300 K
over 1 bar pressure, as given in Table 6. According to the findings, addition of a
small amount of PZ in aqueous solutions of amino acid salts remarkably increases
the absorption performance. Park et al. [68] have also reported five different amino
acid salts blended with PZ. Out of five, three solvent blends (1.5 M ALA + 1 M
PZ, 1.5 M SER + 1 M PZ and 4 M SAR) were identified as potential CO2
absorbents especially for membrane contactors [68]. From the above studies, it is
obvious that blending of amino acid salts and amines offer better CO2 absorption
characteristics, and at the same time add to a some extent the drawbacks associated
132 A.M. Shariff and M.S. Shaikh

Table 6 Various studies on CO2 solubility in amino acid salt blends


Amino acid salt blends Concentration Temperature Pressure Refs.
Na-GLY + MEA 0.5–2 M SG 298–313 K 0–35 kPa [65]
0.5–2 M MEA
K-SER + PZ 1.5 M + 1 M 313.15–353.15 K 0–665.5 kPa [59]
K-ALA + PZ 1.5 M + 1 M 313.15–353.15 K 0.2–1041.7 kPa [59]
K-ALA + PZ (15 + 5) wt% 300 K 1.0 bar [66]
K-BALA + PZ (15 + 5) wt% 300 K 1.0 bar [66]
K-SAR + PZ (15 + 5) wt% 300 K 1.0 bar [66]
K-GLY + PZ 4M+1M 40 °C – [68]
K-TAU + PZ 2.5 M + 1.5 M 40 °C – [68]
K-SAR + PZ 3M+1M 40 °C – [68]
K-ALA + PZ 1.5 M + 1 M 40 °C – [68]
K-SER + PZ 1.5 M + 1 M 40 °C – [68]

with amines. Therefore, it is required to carefully formulate the blends of amino


acid salts and amines so that the goal of effective CO2 removal could be sustained.
The solubility data available in the literature for different types of amino acid
salts and their blends have been compared at 313 K as shown Fig. 2. It is evident
form Fig. 2 that almost all the amino acid salt solutions and their blends have higher
CO2 loading as compared to commercial amine (MEA). Among the amino acid
salts studies, K-LYS shows highest solubility of CO2 as compared to various other

Fig. 2 CO2 loading of various amino acid salt solutions and their blends at 313 K
Aqueous Amino Acid Salts and Their Blends … 133

amino acid salt solutions. Moreover, the other amino acid salts have comparable
CO2 loading at reported temperature for the given range of pressure, and concen-
tration. Therefore, the amino acid salts are potential alternatives to alkanolamines
for efficient absorption of CO2. This comprehensive comparison of CO2 loading of
the various amino acid salts could be very helpful for the selection of appropriate
solvent based on amino acid salts for CO2 capture from various streams of gas.

5.2 Kinetics of CO2 Absorption

Kinetics of CO2 absorption show that how fast the reaction between CO2 and amino
acid salts occurs. Reaction kinetics is very important for the design and simulation
of the acid gas removal process, especially for the absorber and desorber design [69,
70]. There are various studies available on the kinetics of CO2 absorption in
aqueous amino acid salt solutions. These studies have been conducted to obtain the
kinetic data useful for CO2 removal plant design. Table 7 presents the various
studies conducted on reaction kinetics of CO2 absorption in aqueous amino acid salt
solutions. In this section, the kinetics of CO2 absorption in various amino acid salt
solutions have been briefly reviewed systematically in order to make it easy to
assess the solvent for commercial applications.

Table 7 Kinetics studies of CO2 absorption in aqueous amino acid salts


Aqueous amino acid salts Concentration Temperature (K) Refs.
K-TAU 200–300 mol m−3 285–305 [69]
K-GLY 200–300 mol m−3 295 [69]
Na-GLY 1.0–3.5 kmol m−3 303.15–323.15 [71]
K-GLY 0.1–3.0 M 293–303 [72]
K-THR 0.1–3.0 M 293–313.15 [74]
K-SAR, Li-SAR 0.5–3.0 mol L−1 298 [75]
K-PRO, Li-PRO 0.5–3.0 mol L−1 298 [75]
K-SAR 0.5–3.8 M 298–303 [76]
K-GLY 0.10–0.50 kmol m−3 298–303 [77]
K-SAR 1.0–4.0 kmol m−3 298.15–335.15 [78]
K-ALA 1.0–3.0 M 293.15–313.15 [79]
K-PRO 0.5–3.0 kmol m−3 303–323 [80]
K-GLY 0.5–2.0 M 298.15–335.15 [81]
K-ALA 0.5–2.0 M 298.15 [81]
K-PRO 0.5–3.0 mol dm−3 290–303 [82]
Na-PRO 0.5–3.0 mol dm−3 298 [82]
Na-TAU 5–50 mol m−3 298–313 [83]
Na-PRO 4–12 mol m−3 298–313 [83]
K-PRO 1M 313 [84]
134 A.M. Shariff and M.S. Shaikh

Kumar et al. [69] reported the kinetics of the CO2 absorption in aqueous K-TAU
at 285 to 305 K, while K-GLY at 295 K. Unlike alkanolamines, aqueous K-TAU
and K-GLY showed an increment in partial reaction order from one at molar
concentration higher than 1000 mol m−3. Zwitterion and termolecular mechanism
were studied for better understanding of the kinetics. It is reported that the value of
k2 (forward second order reaction rate) for the zwitterion mechanism of CO2
absorption in aqueous K-TAU is considerably higher (12.60 m−3 mol−1 S−1) as
compared to alkanolamines (4.94 m−3 mol−1 S−1) [70]. Moreover, it was deter-
mined that the reactivity of CO2 with K-GLY is higher than that of K-TAU because
the K-GLY has higher basic strength [69]. Lee et al. [71] investigated the kinetics of
CO2 absorption in aqueous Na-GLY. The second-order reaction rate constant are
reported as 218, 576, and 1034 m3 kmol−1 S−1 respectively at 303.15, 313.15, and
323.15 K. Additionally, energy of activation for the reaction was reported to be
63.8 kJ/mol [70]. Portugal et al. [72] have reported the kinetics of CO2 absorption
in aqueous K-GLY solutions. They have found out that aqueous K-GLY solution
has faster reaction kinetics as compared to MEA. For MEA (1 M) at 298 K, the
overall kinetic constant is reported as 5920 s−1 [73], however, at same conditions,
the value is 13400 s−1 for aqueous K-GLY. Hence, the amino acid salt studied in
this work showed significant potential to be used for absorption of CO2 [72]. In
another study, Portugal et al. [74] reported the kinetics of CO2 absorption in
aqueous K-THR solution. The results were also compared with that of alka-
nolamine (DEA) and K-GLY solution at 298 K for 1 M concentration. It is reported
that K-THR has slower kinetics than the K-GLY; however, it is faster than DEA
and comparable with other amine solutions [74]. Holst et al. [75] have reported the
screening study of CO2 absorption kinetics in various amino acid salt solutions. The
potassium salts of various amino acids such as 6-aminohexanoic acid, BALA,
ARG, GLU, MET, PRO and SAR were used in the initial study at 298 K for
0.5 mol L−1. It is reported that the K-PRO and K-SAR has better absorption
characteristics as compared to the rest of the amino acids studied because they
combine with relatively high apparent rate constant with low values of pKa [75].
Simons et al. [76] investigated the kinetics of CO2 absorption in aqueous K-SAR
and reported the influence of concentration, temperature and CO2 loading on the
rate of reaction. Moreover, the temperature has the positive effect on the reaction
rate constant because it increased with the rise in temperature. Similarly, the overall
reaction rate constant increases with an increase in amine concentration.
Furthermore, it was observed that the values of overall and apparent reaction rate
constant decrease with increasing the CO2 loading of the amino acid salt solution.
The reaction rate constant for K-SAR is higher than that of MEA [76]. Vaidya et al.
[77] have studied the kinetics of CO2 absorption in two aqueous amino acid salt
solutions, i.e., K-GLY and K-TAU. The second-order reaction rate constant of
K-GLY at 303 K is 6.29 m3 mol−1 s−1. Additionally, K-GLY was also used as
activator /promoter in DEEA to enhance the CO2 removal rate. It was found that
K-GLY is an attractive promoter for enhancing the CO2 absorption rate [77].
Similarly, another research group Aronu et al. [78] carried out the kinetic study of
CO2 absorption in K-SAR at few different temperatures (298.15–335.15 K) and
Aqueous Amino Acid Salts and Their Blends … 135

concentrations (1.0–4.0 kmol m−3). It was observed that the reaction rate constant
increased with increasing the temperature and concentration, and this finding is
similar to the one reported by Simons et al. [76]. Moreover, the value of reaction
rate constant for K-SAR is higher than MEA, but still comparable [78].
Kim et al. [79] reported the kinetic study of CO2 absorption using K-ALA
solution at 393.15 to 313.15 K for 1–3 M concentration. It is reported that K-ALA
solution has lower absorption rate as compared to K-GLY solution. However, due to
the steric hindrance effect, the loading capacity of K-ALA may be improved by
increasing the charge density of N atom of the K-ALA solution. In case of the
desorption, K-ALA has high desorption rate caused by the slow absorption rate. This
could reduce the energy of regeneration [79]. Paul et al. [80] reported the kinetic data
for the absorption of CO2 using K-PRO solution, and found that the studied solvent
has higher second order reaction rate constants as compared to alkanolamines and
few other amino acid salt solutions. The values of overall reaction rate constant (kov)
of K-THR investigated by Portugal et al. [74] and K-TAU reported by Vaidya et al.
[77] is lower than K-PRO studied in this work. However, the values reported for
K-SAR and K-GLY by Portugal et al. [72] and Simons et al. [76] are very close to
K-PRO at 303 K. Similarly, Majchrowicz et al. [82] have also studied the kinetics of
CO2 absorption in K-PRO solution, and Na-PRO solution at 298 K. It is reported
that the reactivity of K-PRO with CO2 is higher than Na-PRO. Moreover, the value
of kov for K-PRO is higher than K-THR and K-TAU while it is comparable with
MEA, K-SAR and K-GLY [82]. Another study on reaction kinetics of CO2 in
aqueous Na-TAU and Na-PRO have been reported by Sodiq et al. [83]. It was found
that the Na-PRO has faster reaction rates than Na-TAU and commercial MEA.
However, Na-TAU showed slower rates than MEA at low concentrations [83].
Likewise, Fang et al. [84] have also reported the kinetics of absorption of CO2 in
K-PRO solutions and influence of total pressure on absorption. It was found that
high pressure has a positive effect on the rate of CO2 absorption. Moreover, K-PRO
was found to be promising solvent for CO2 removal [84].
Various researchers have reported the overall reaction rate constant (kov) for
some amino acid salt solutions. Figure 3 shows the comparison of kov for
absorption of CO2 in various amino acid salt solutions at 303 K for various reported
concentrations. As can be seen from Fig. 3 that K-PRO has the highest kov than
other amino acid salt solutions. K-SAR and K-GLY has somewhat comparable
values of kov. K-ALA and K-THR has the lowest values of kov. Therefore, it is
apparent that the K-PRO is the most promising solvent for bulk removal of CO2
with faster reaction kinetics from other amino acid salt solutions. Moreover, the
activation energy of various amino acid salt solutions reported by various
researchers are given in Table 8, and the comparison between the activation energy
of various amino acid salts solutions and amines is shown in Fig. 4. Activation
energy determines the response of reaction rate to temperature, and it is one of the
keys to determining the optimum operating conditions and loadings. It is clear from
the Fig. 4 that Na-PRO has the lowest energy of activation than even K-PRO
suggesting that the rate of reaction of CO2 with Na-PRO is very fast as also
suggested by Sodiq et al. [83]. The K-PRO and K-SAR has the comparable energy
136 A.M. Shariff and M.S. Shaikh

Fig. 3 Overall reaction rate constants for CO2 absorption in amino acid salt solutions

Table 8 Activation energy Amino acid salts Ea (KJ/mol) References


of aqueous amino acid salts
and amines during CO2 K-TAU 47.40 [69]
absorption Na-GLY 63.80 [71]
K-GLY 48.23 [72]
K-SAR 26.0 [76]
K-PRO 36.5 [80]
K-PRO 43.3 [81]
Na-TAU 48.1 [83]
Na-PRO 12.0 [83]
Amines
MEA 41.2 [85]
DEA 53.1 [85]
MDEA 47.9 [86]
AMP 41.7 [87]
PZ 33.6 [88]
TEA 35.8 [89]

of activation. K-TAU and K-GLY have also the comparable activation energy but
higher than K-PRO and K-SAR. Interestingly, Na-GLY has the highest activation
energy than all the amino acid salts and amines; however, it is comparable with
DEA and MDEA.
Aqueous Amino Acid Salts and Their Blends … 137

Fig. 4 Comparison of activation energy of amino acid salt solutions and amines

5.3 Degradation Tendency

Solvent degradation is a major concern in the CO2 absorption process because of


the significant impact on operational cost and intention to use thermal compression
from high temperature stripping to minimize the operational energy. The degra-
dation of the CO2 capture solvent consume the amine/amino acid salt, as a result;
the makeup rate of the solvent becomes higher. Besides, degradation of solvent also
generates operational problems such as increasing the foaming and corrosion ten-
dency during the absorption-regeneration process [15]. Therefore, the degradation
performance evaluation is very important for the appropriate selection of solvent.
Generally, there are two types of degradation in post-combustion CO2 removal
process such as oxidative and thermal degradation. In an oxidation degradation,
oxidized fragments of amine are formed. These include organic acids, ammonia,
aldehydes, and amines. This degradation is caused in oxygen-rich environment
typically during the flue gases treatment. Such degradation results in toxic products
of degradation and the solvent loss. The other type is the thermal degradation,
caused by either thermal decomposition that normally occurs above temperature
200 °C or by various reactions that are accelerated by high temperatures [90].
Amine based solvents are usually considered to have high oxidative and thermal
degradation [15]. This is one of the drawbacks in post-combustion CO2 absorption
because of the severe negative consequences on the process, and the effects of toxic
degradation products on the environment. Due to these limitations of the amines,
amino acid salts are introduced as alternatives to amines because they are consid-
ered as more stable towards degradation, especially in the presence of oxygen due
138 A.M. Shariff and M.S. Shaikh

to the ionic nature [32, 39, 47]. There are very few studies on the degradation of the
amino acid salts for CO2 capture. Epp et al. [90] studied the oxidative degradation
of the amines (MEA, DGA) and amino acid salt, K-GLY. They have conducted the
experiments in closed vessel containing oxygen as a gas phase. The K-GLY is not
much oxidative stable (degradation: 0.0024 mol NH3/l CO2) as compared to MEA
(degradation: 0.0029 mol NH3/l CO2), because the formation of degradation
products, including ammonia and formaldehyde are almost same in case of MEA
and K-GLY. The results advocated that the benefit of K-GLY in terms of better
resistance to oxidative degradation is not significant as expected [90]. The more
thorough investigation is required in real environment of flue gases for better and
definite findings so that the appropriate solvent may be selected for commercial
CO2 absorption process. Huang et al. [91] reported the thermal degradation of
various amino acid salt solutions such as Na-GLY, Na-SAR, Na-ALA and
Na-BALA at 125, 135 and 145 °C. It is reported that Na-GLY has more degra-
dation rate as compared to MEA at the same amine concentration (5.0 M) and at all
temperatures. These results are unexpected because the amino acid salts were
considered more thermally stable compared to amines as reported in Portugal et al.
[72], and at various places in the literature [32, 39] Moreover, the thermal degra-
dation of other amino acids for 2.5 M concentration was also compared in the study
at 135 °C. Na-GLY and MEA both at 2.5 M concentration were also used as
reference. The results showed that all the studied amino acid salts have higher rates
of amine loss as compared to MEA. Moreover, the thermal stability was found in
the order of MEA > SAR > ALA > BALA. SAR showed the slowest degradation
rate but still almost two times higher than MEA, while BALA has the fastest
degradation at all studied temperatures [91].
From the above studies, it is clear that only few amino acid salts have been tested
for thermal and oxidative degradation. The studied amino acid salts showed less
thermal stability than commercial MEA, which contradicts with the information on
benefits of the amino acid salts summarized in Table 3. Based on these observa-
tions, it is still not straightforward to say that all the amino acid salts are less or high
thermally stable than the amine solvents because of the lack of studies available on
degradation of amino acid salts. Therefore, a thorough study on degradation of
amino acid salts is required for stronger claims in both the cases.

5.4 Corrosion Tendency

Corrosion in the CO2 absorption process is one of the most serious problems
because it reduces the equipment life, causes unscheduled downtime, loss of pro-
duction and sometimes even injury to the personnel [92]. The cost of the production
losses due to the unscheduled downtime of the typical amine plant varies from
$10,000 and $30,000 per day. Besides, the extra expenses are also incurred for the
restoration of corroded systems and other management activities initiated for the
mitigation of the corrosion [93]. The conventional alkanolamines are highly
Aqueous Amino Acid Salts and Their Blends … 139

corrosive towards the equipment and accessories used for CO2 absorption.
Corrosion is one of the problems, which causes the high cost of the CO2 absorption
by amines [94].
Amino acid salts’ solutions are most promising solvent for the absorption of CO2
due to various positive characteristics as mentioned in previous sections. A part
from various CO2 absorption characteristics, it has been shown that the amino acid
salt solutions are also less corrosive as compared to alkanolamines [32]. Therefore,
the cost of the CO2 capture process due to corrosion loss could potentially be
reduced by using the amino acid salt solutions instead of amines.
Since the amino acid salts are still new and emerging solvents for CO2 capture,
therefore, there are very few investigations available in the literature on the cor-
rosion rate of these solvents. Ahn and co-researchers [94] have investigated the
corrosion rate of two amino acid salts such as K-TAU and K-GLY by using the
conventional weight loss method (ASTM E3-80). Corrosion tests were conducted
for different concentrations of amino acid salts (1.5–5 M) and at various temper-
atures (313.15–353.15 K) using carbon steel grade 1018. This type of carbon steel
is widely used for the construction and fabrication of equipment and accessories in
the CO2 absorption plants. The effect of temperature, concentration and CO2
loading was determined for the studied amino acid salts. The results have been
compared with the commercial MEA at 353.15 K and CO2 loading of 0.62 for
MEA and amino acid salts. It is reported that the corrosion rate of MEA increases
with an increase in the concentration because more amounts of molecules absorb
more CO2. The CO2 dissolved in the form of bicarbonate ions cause iron disso-
lution. However, the corrosion rate decreased with an increase in the concentration
of K-GLY. This is contrary to MEA because of the hydrolysis of the carbamate
results in decreased bicarbonate ions, thereby reducing the corrosion. Moreover,
K-TAU showed the same trend as of MEA. The corrosion rate increased with
increasing the temperature and CO2 loading [94]. The corrosion rate of K-TAU is
lower than that of MEA. It was suggested that K-GLY is beneficial at high con-
centration while K-TAU is beneficial at lower concentration [94]. Moreover, the
addition of corrosion inhibitors has been proposed by Ahn et al. [94] to reduce the
corrosion tendency of amino acid salts. In this regard, PZ, sodium metavanadate
(NaVO3) and sodium sulfite (NaSO3) were suggested. Increasing the PZ concen-
tration in amino acid salt further increased the corrosion of carbon steel because of
the blended solution absorbed more CO2 thereby increasing the amount of bicar-
bonate ions and so does the corrosion [94]. However, addition of NaVO3 showed
better inhibition performance (about 99.9%) than NaSO3 (93.7%). The inhibition
performance increased with the rise of inhibitor concentration [94]. Although the
corrosion inhibitors have potential to prevent corrosion, however, few operational
issues can be induced with their use. For example, use of NaVO3 may increase the
solvent degradation by accelerating the reaction rate such as observed by using it
with MEA [95].
Mazinani et al. [65] investigated the corrosion tendency of blend of Na-GLY and
MEA by electrochemical technique (ASTM G5-94) using potentiostat at various
blend ratios without dissolved CO2 at 308 K. The results show that, with increasing
140 A.M. Shariff and M.S. Shaikh

the Na-GLY concentration in the blend, the corrosion rate increases [65]. Recently,
another study on the corrosion tendency of K-LYS was carried out by Mazinani
et al. [62] following the same method as described in their previous investigation
[65]. The results show that the corrosion rate increases with the rise in the con-
centration of K-LYS solution, and follows the same trend as observed previously.
In both studies, the measurement could have shown a higher corrosion rate if CO2
was dissolved in the Na-GLY blends and aqueous solutions of K-LYS.
Based on the analysis of the available literature, amino acid salts have been
considered as less corrosive as compared to alkanolamines. However, the existing
data and knowledge of the corrosion evaluation of amino acid salts are very scarce.
More thorough investigation on corrosion is still required for amino acid salts to
verify their less corrosivity. The data on recommended corrosion inhibitors to be
used with amino acid salt solutions is also very limited; therefore, care must be
exercised in the selection of a proper corrosion inhibitor, so that the solvent actual
performance may not be affected.

5.5 Toxicity

The suitable absorption characteristics alone are not enough for the right selection
of the solvent for CO2 capture, but the possible environmental and health risks
associated with the selected solvents must be known prior to use or commercial
application. Without the proper investigation on toxicity and risks assessment, it
may not be viable to use any newly developed solvent; otherwise, the new envi-
ronmental issues could be raised after solving the climate-change problems.
The most widely used solvents for CO2 absorption are alkanolamines; however,
apart from various operational limitations during CO2 removal, these solvents also
induce various environmental and health risks [21–25]. The new emerging solvents
such as amino acid salt solutions are considered as green and environmental
friendly. They have no harmful effects on the environment because these are nat-
urally present in the environment [25, 96]. Apart from various benefits of amino
acid salts, the environmental friendly characteristics of these solvents also make
them as an alternative to alkanolamine solvents. Shao et al. [97] have reported that
the amino acid group show high biodegradability with lower toxicity than the
amines [97]. The comprehensive toxicity analysis of amino acid salts after
absorption of CO2 has not been reported yet in order to prove their environmental
suitability. However, some of the toxicological information in terms of lethal dose
(LD50) for various amino acids and amines is available, and given in Table 9. The
data have been taken from the safety data sheets of manufacturers. LD50 is the
amount of the chemical /toxic material, which is sufficient to kill the 50% of the test
animals in the certain time. The higher values of LD50 shows the less toxicity while
the lower values denotes the higher toxicity [114].
Aqueous Amino Acid Salts and Their Blends … 141

Table 9 Toxicity of various Amino acid/amines LD50 (mg/kg) Refs.


amino acids and amines
Gly 7930 (oral—rat) [98]
PRO >5110 (oral—rat) [99]
TAU >5000 (oral—rat) [100]
HIS 15,000 (oral—rat) [101]
MET 36,000 (oral—rat) [102]
GLU 7500 (oral—rat) [103]
MEA 700 (oral—mouse) [104]
DEA 710 (oral—rat) [105]
TEA 2200 (oral—rabbit) [106]
PZ 600 (oral—mouse) [107]

The comparison between the toxicity of amino acids and amines is shown in
Fig. 5. The higher values of LD50 indicate the lower toxicity of the chemical while
lower values indicate the high toxicity as mentioned previously.
As can be observed from Fig. 5, the values of LD50 are higher in case of all
amino acids than the amines, suggesting that the amino acids are very less toxic as
compared to various amines. From all the amino acids, MET is less toxic than other
amino acids, while the PZ has the highest toxicity among all other amines. Thus,
amino acids can be considered as eco-friendly solvent as compared to amines, and
be used for CO2 absorption.

Fig. 5 Comparison of toxicity (LD50) of amino acids and amines


142 A.M. Shariff and M.S. Shaikh

5.6 Energy Aspects

During the selection of appropriate solvent for CO2 absorption, it is essential to


assess the energy required for the CO2 absorption process because it is directly
associated with the process efficiency and cost. Since CO2 capture is an energy
demanding process, therefore, it is always important to optimize the energy effi-
ciency of the capture process by developing the new solvents and processes. Some
estimations show that about 30% of the energy produced by power plants would
have to be dedicated to the operation of CO2 capture plant. This makes the cost of
electricity generation to almost double. The doubling of the electricity cost by the
addition of a carbon capture plant is also due to the overall effects of capital and
operating cost. Moreover, these costs vary with location. The major portion of the
energy penalty is due to the high energy required for solvent regeneration, and the
energy required for compression of CO2 for pipelining and sequestration [39, 108].
The commercial alkanolamines solvents require high energy for the absorption of
CO2 apart from their various other operational drawbacks. Therefore, the cost of the
energy required and the cost incurred due to various operational drawbacks make
up the overall process cost at maximum [39]. To overcome these issues, it is always
required to search for the candidate solvent that would save the cost in terms of
minimum energy required and by various other means. Amino acid salt solutions
have been considered as efficient solvents in terms of various benefits described in
previous sections, as well as in terms of the energy requirements. Tobias et al. [109]
reported the process developed by Siemens utilizing the amino acid salt solution,
and it is demonstrated that the amino acid salt solutions have a low absorption
enthalpy that can reduce the energy demand. Siemens efficiently developed the
process for CO2 capture employing the amino acid salt solution with the advantages
of a low energy requirement for regeneration of solvent. Weiland et al. [110]
reported that Na-GLY requires almost same reboiler duty as that of MEA.
However, 85% of the CO2 recovery can be achieved with 1250 m3/h of the solvent
as compared to 1500 m3/h for MEA for a typically configured CO2 capture plant
(3000 tons/day CO2) [110]. This reduction in the flow rate could save the energy
required for process operation. Moreover, some solvent blends based on 30 wt%
KDiMGLY blended with 15 wt% MEA were also tested. The simulated results
showed that 10% of the energy could be saved with 85% of the CO2 recovery as
compared to 30 wt% MEA. However, addition of 15 wt% MEA could arise the
issues linked with amines. In addition to this, another solvent blend ratio consists of
40 wt% KDiMGLY and 5 wt% PZ achieved 20% reduction of the energy with 85%
recovery of CO2 and reduced solvent flow rate of 1250 m3/h. The heat of
absorption (integral values from 0 to 0.3 loading) of MEA, MDEA, PZ, Na-GLY
and KDiMGLY are 84, 58, 76, 85, and 55 kJ/gmol−1 respectively at 25 °C [110]. It
is apparent that KDiMGLY has lowest heat of absorption, thereby may require less
regeneration energy as compared to other solvents studied. Lim et al. [57] have
reported the heat of absorption of K-ALA and K-PRO at 298 K. The values show
that K-ALA has lowest heat of absorption (53.26 kJ/mol) than K-PRO
Aqueous Amino Acid Salts and Their Blends … 143

(90.20 kJ/mol), MEA (81.77 kJ/mol) and DEA (67.06 kJ/mol) respectively. These
results suggest that amino acid salt, especially; K-ALA is expected to achieve better
regeneration performance because of low heat of absorption. Similarly,
Majchrowicz and Brilman [58] have reported the integral enthalpy of CO2
absorption in K-PRO solution as 54.3 kJ/mol at loading of 0.80 and temperature
range of 298.37–313.83 K. This value is lower than Na-GLY (72.5 and 59.5 kJ/mol
at loading, 0.90 and 0.86 over temperature range of 313.15–323.15 K) and MEA
[111]. The results obtained by Majchrowicz and Brilman [58] and Salazar et al.
[111] indicate that Na-GLY and K-PRO are efficient absorbent in terms of energy
required for regeneration. K-PRO may be more cost savings than Na-GLY based on
the results of the enthalpies. Few discrepancies are observed, such that the value of
enthalpy of Na-GLY reported by Weiland et al. [110] is higher than that of Salazar
et al. [111]. Similarly, the enthalpy of K-PRO reported by Lim et al. [57] is quite
higher than the one obtained by Majchrowicz and Brilman [58]. Sanchez-Fernandez
et al. [48] have summarized the results of an energy-efficient process known as
DECAB, DECAB Plus and pH swing. In this process, a precipitating amino acid
such as K-TAU was used. The results demonstrated that the process requires a
reduced amount of energy than MEA. DECAB Plus process was recognized as
more energy efficient (66% of the conventional MEA) [48]. Although the design of
this process requires somewhat larger equipment, especially absorber, the mainte-
nance cost would be significantly lower [48].
Some studies suggested that the amino acid salts are energy efficient solvents,
while the others indicated that these may require moderate energy during CO2
absorption as compared to alkanolamines. Since, there are very limited studies
available in literature on the energy performance of amino acid salt solutions;
therefore, more research should be devoted to energy analysis of amino acid salts to
completely endorse their use. For comprehensive evaluation of energy analysis of
the process employing the amino acid salts, the study of relation between heat of
absorption, overall regeneration heat duty and process parameters is required [112].

5.7 Other Considerations

There are few considerations important for the solvents to be used for CO2
absorption, such as vapor pressure and flammability. The vapor pressure of the
solvent plays an important role in their selection for CO2 absorption. The solvents
that have lower vapor pressure are desirable because it minimize the vapor loss.
While the high vapor pressure can cause the solvent loss by evaporation together
with the clean gas. The loss of the solvent due to higher vapor pressure would incur
an extra process cost [39]. Therefore, it is always required to search, and develop a
solvent with lower or negligible vapor pressure. Commercially used alkanolamines
have higher vapor pressure; therefore, the urge to find the new solvent has
increased, and consequently; a new class of solvents such as amino acid salt
solutions have been introduced. These solvents have near-zero or negligible vapor
144 A.M. Shariff and M.S. Shaikh

pressure because they are salts with ionic structure. Such typical characteristic of
the amino acid salts makes them a prospective alternative to alkanolamines. The
solvent refill cost can be reduced potentially due to the minimum vapor loss during
CO2 absorption [32, 109]. Another benefit of amino acid salt is that they are not
flammable and are less sensitive to oxygen as compared to alkanolamine.
Therefore, these solvents are considered as non-hazardous [109]. Based on the
analysis of these characteristics of amino acid salt solutions, it can be concluded
that these are potential, cost saving, and energy efficient solvents for CO2
absorption.

6 Perspective and Conclusion

The energy and cost efficient CO2 absorption from flue gases is a main technical
challenge in terms of the solvent selection. Conventional alkanolamine solvents
have many operational limitations as detailed out in the above sections. The new
solvents based on amino acid salts are promising in terms of energy and cost. There
have been several individual studies over the recent years on such solvent systems
because of their versatile and promising characteristics. However, in this chapter,
the overall performance analysis of aqueous amino acid salts and their blends was
carried out for evaluating their commercial possibilities. Amino acid salt solutions
and their blends demonstrated a significant potential for CO2 absorption; however,
it is not straightforward to select the potential solvent. There are numerous
parameters to be considered, which affect the overall cost of the process. Therefore,
the sensible and holistic approach is required for choosing the most appropriate
solvent.
The performance analysis of amino acid salt solutions was carried out, and some
parameters were identified to have substantial influence. These parameters include
but not limited to CO2 absorption performance (CO2 loading), kinetics of absorp-
tion, degradation, corrosion, toxicity, energy aspects and various other considera-
tions. Each of the parameters were analyzed and compared with conventional
solvents. The performance analysis was somewhat difficult because of limited
studies available on amino acid salts, and their blends. Besides, few discrepancies
were also found between various studies, which made it challenging to streamline,
and normalize the performance analysis results. Usually, the solvent is selected
based on the comparative analysis. Therefore, a possible comparative analysis was
carried out to meet the objective. However, a comprehensive comparison between
various amino acid salts, and their blends was relatively challenging because of
scarcity of data. Therefore, a complete performance and comparative analysis
would require more research on amino acid salts, and their blends.
From the performance analysis carried out in this chapter, it was found that
almost all the amino acid salts showed good absorption characteristics than alka-
nolamines. Amino acid salts blended with promoters, especially PZ showed
increased CO2 loading than various amino acid salts alone. Moreover, amino acid
Aqueous Amino Acid Salts and Their Blends … 145

salt such as K-PRO and Na-PRO showed fast reaction kinetics as compared to other
amino acids. The activation energy of few amino acid salts such as K-TAU,
Na-TAU and Na-GLY was higher than MEA and PZ, suggesting that these amino
acids have slow reaction kinetics. Na-PRO and K-SAR have the lowest activation
energy than all other amino acid salts and alkanolamines. The degradation rates of
few of the amino acid salts were higher than MEA, while the corrosion resistance
found to be lower. The toxicity of the amino acids was very less as compared to
amines. From the energy analysis, it was found that various amino acid salts and
blends such as K-ALA, KDiMGLY + PZ, and K-PRO require less regeneration
energy than commercial MEA. The results of a new process such as DECAB Plus
showed that the large amount of energy could be saved by using the precipitating
amino acid (K-TAU). Moreover, the amino acid salts have negligible (near-zero)
vapor pressure, and they are nonflammable.
From the above evidence, it can be concluded that amino acid salts could be used
as an effective alternative to alkanolamines for CO2 capture. In summary, based on
the comprehensive analysis and assessment, these solvents are environmental
friendly and energy efficient with higher CO2 loading capacity, faster reaction
kinetics and require less regeneration energy. Besides, these solvents are
non-volatile, non-flammable, less corrosive and oxidative stable to some extent.
Moreover, these solvents can be made more effective by blending various additives
with aqueous amino acid salts. Some of the amino acid salts have a comparatively
significant potential than the others. Therefore, these solvents should be tested for
pilot or commercial-scale demonstration. For the others, additional fundamental and
systematic research is required, especially on thermal degradation, corrosion,
kinetic study, and energy requirement assessment for comprehensive performance
analysis.

Acknowledgments The authors are grateful to Research Centre for CO2 Capture (RCCO2C),
Department of Chemical Engineering, Universiti Teknologi PETRONAS for supporting this work.

References

1. Abu-Zahra MRM, Niederer JPM, Feron PHM, Versteeg GF (2007) CO2 capture from power
plants: part I. A parametric study of the technical performance based on monoethanolamine.
Int J Greenhouse Gas Control 1(2):135–142
2. Dongwoo K, Sangwon P, Hoyong J, Jaehong M, Jinwon P (2013) Solubility of CO2 in
amino-acid-based solutions of (potassium sarcosinate), (potassium alaninate + piperazine),
and (potassium serinate + piperazine). J Chem Eng Data 58(6):1787–1791
3. Yang HQ, Xu ZH, Fan MH, Gupta R, Slimane RB, Bland AE, Wright I (2008) Progress in
carbon dioxide separation and capture: a review. J Environ Sci 20(1):14–27
4. NOAA (2014) Recent global monthly mean CO2. National Oceanographic and Atmospheric
Administration: (NOAA). Silver Springs, MD
5. Williams M (2002) Climate change: information kit. United Nations Environment
Programme (UNEP) and the United Nations Framework Convention on Climate Change
(UNFCCC), Geneva
146 A.M. Shariff and M.S. Shaikh

6. IPPC (Intergovernmental Panel on Climate Change) (2007) IPCC Report. IPCC, Geneva
7. Rubin ES, Rao AB (2002) A technical, economic and environmental assessment of
amine-based CO2 capture technology for power plant greenhouse gas control. Environ Sci
Technol 36(20):4467–4475
8. Khan FM, Krishnamoorthi V, Mahmud T (2011) Modelling reactive absorption of CO2 in
packed columns for post-combustion carbon capture applications. Chem Eng Res Des 89
(9):1600–1608
9. Shaikh MS, Shariff AM, Bustam MA, Murshid G (2015) Measurement and prediction of
physical properties of aqueous sodium L-prolinate and piperazine as a solvent for CO2
removal. Chem Eng Res Des 102:378–388
10. Lucquiaud M, Gibbins J (2011) On the integration of CO2 capture with coal-fired power
plants: a methodology to assess and optimize solvent-based post-combustion capture
systems. Chem Eng Res Des 89(9):1553–1571
11. Mores P, Scenna N, Mussati S (2011) Post-combustion CO2 capture process: equilibrium
stage mathematical model of the chemical absorption of CO2 into monoethanolamine
(MEA) aqueous solution. Chem Eng Res Des 89(9):1587–1599
12. Diamantonis NI, Boulougouris GC, Tsangaris DM, Kadi MJE, Saadawi H, Negahban S,
Economou IG (2013) Thermodynamic and transport property models for carbon capture and
sequestration (CCS) processes with emphasis on CO2 transport. Chem Eng Res Des 91
(10):1793–1806
13. Posch S, Haider M (2013) Dynamic modeling of CO2 absorption from coal-fired power
plants into an aqueous monoethanolamine solution. Chem Eng Res Des 91(6):977–987
14. Sofia D, Coca Llano P, Giuliano A, Iborra Hernández M, García Peña F, Barletta D (2014)
Co-gasification of coal-petcoke and biomass in the puertollano IGCC power plant. Chem
Eng Res Des 92(8):1428–1440
15. Kohl AL, Nielsen R (1997) Gas Purification, 5th edn. Gulf Publishing Company, Houston
16. Hoff KA, Juliussen O, Falk-Pedersen O, Svendsen HF (2004) Modeling and experimental
study of carbon dioxide absorption in aqueous alkanoamines solutions using a membrane
contactor. Ind Eng Chem Res 43(16):4908–4921
17. Rochelle GT (2009) Amine scrubbing of CO2 capture. Science 325:1652–1654
18. Moioli S, Pellegrini LA (2013) Regeneration section of CO2 capture plant by MEA
scrubbing with a rate-based model. Chem Eng Trans 32:1849–1854
19. Lepaumier H, Picq D, Carrette PL (2009) New amines for CO2 capture: II. Oxidative
degradation mechanisms. Ind Eng Chem Res 48(20):9068–9075
20. Shaikh MS, Shariff AM, Bustam MA, Murshid G (2014) Physicochemical properties of
aqueous solutions of sodium l-prolinate as an absorbent for CO2 removal. J Chem Eng Data
59(2):362–368
21. Dawodu OF, Meisen A (1996) Degradation of alkanoamines blends by carbon dioxide.
Can J Chem Eng 74(12):960–966
22. Jamal A, Meisen A (2001) Kinetics of CO2 induced degradation of aqueous diethanolamine.
Chem Eng Sci 56(23):6743–6760
23. Portugal AF, Sousa JM, Magalhaẽs FD, Mendes A (2009) Solubility of carbon dioxide in
aqueous solutions of amino acid salts. Chem Eng Sci 64(9):1993–2002
24. Aronu UE, Hartono A, Svendsen HF (2011) Kinetics of carbon dioxide absorption into
aqueous amine amino acid salt:3-(methylamino) propylamine/sarcosine solution. Chem Eng
Sci 66(23):6109–6119
25. Shuaib Shaikh M, Shariff AM, Bustam MA, Murshid G (2014) Physical properties of
aqueous solutions of potassium carbonate + glycine as a solvent for carbon dioxide removal.
J Serb Chem Soc 79(6):719–727
26. Shaikh MS, Shariff AM, Bustam MA, Murshid G (2015) Physicochemical properties of
aqueous solutions of sodium glycinate in the non-precipitation regime from 298.15 to 343.15
K. Chin J Chem Eng 23(3):536–540
27. Hook RJ (1997) An investigation of some sterically hindered amines as potential carbon
dioxide scrubbing compounds. Ind Eng Chem Res 36(5):1779–1790
Aqueous Amino Acid Salts and Their Blends … 147

28. Simons K, Brilman WDWF, Mengers H, Nijmeijer K, Wessling M (2010) Kinetics of CO2
absorption in aqueous sarcosine salt solutions: influence of concentration, temperature, and
CO2 loading. Ind Eng Chem Res 49(20):9693–9702
29. Hartono A, Aronu UE, Svendsen HF (2011) Liquid speciation study in amine amino acid
salts for CO2 absorbent with 13C-NMR. Energy Procedia 4:209–215
30. Song H-J, Sangwon P, Hyuntae K, Ankur G, Park J-W, Lee S-J (2012) Carbon dioxide
absorption characteristics of aqueous amino acid salt solutions. Int J Greenhouse Gas
Control 11:64–72
31. Shaikh MS, Shariff AM, Bustam MA, Murshid G (2013) Physical properties of aqueous
blends of sodium glycinate (SG) and piperazine (PZ) as a solvent for CO2 capture. J Chem
Eng Data 58(3):634–638
32. Majchrowicz ME (2014) Amino acid salt solutions for carbon dioxide capture. Ph.D.
dissertation, University of Twente, The Netherlands
33. Goan JC, Miller RR, Piatt VR. (1960) Alkazid M as a Regenerative Carbon Dioxide
Absorbent. NRL Report 5465. Naval Research Laboratory, Washington DC
34. Holst JV, Versteeg GF, Brilman DWF, Hogendoorn JA (2009) Kinetic study of CO2 with
various amino acid salts in aqueous solution. Chem Eng Sci 64(1):59–68
35. Nuchitprasittichai A, Cremaschi S (2011) Optimization of CO2 capture process with aqueous
amines using response surface methodology. Comput Chem Eng 35(8):1521–1531
36. Caplow M (1980) Kinetics of carbamate formation and breakdown. J Am Chem Soc 90
(24):6795–6803
37. Ewing SP, Lockshon D, Jencks WP (1980) Mechanism of cleavage of carbamate anions.
J Am Chem Soc 102(9):3072–3084
38. Chakraborty AK, Bischoff KB, Astarita G, Damewood JR (1988) Molecular orbital approach
to substituent effects in amine-CO2 interactions. J Am Chem Soc 110(21):6947–6954
39. Lerche BM (2012) CO2 capture from flue gas using amino acid salt solutions. Ph.D. thesis,
Technical University of Denmark
40. Gabrielsen J (2007) CO2 capture from coal fired power plants, IVCSEP. Ph.D. thesis,
Technical University of Denmark
41. Da Silva EF (2005) Computational chemistry study of solvents for carbon dioxide
absorption. Ph.D. Thesis, Norwegian University of Science and Technology, NO-7491,
Trondheim
42. Darde V (2011) CO2 capture using aqueous ammonia. Ph.D. Thesis, Technical University of
Denmark, Frydenberg, Copenhagen
43. Feron PHM, Asbroek NAM. (2004) New solvents based on amino-acid salts for CO2 capture
from flue gases. GHGT-7, Vancouver, Canada
44. Goetheer ELV, Nell L (2009) First pilot plant results from TNO’s solvent development
workflow. Carbon Capture J 8:2–3
45. Schneider R, Schramm H (2011) Environmental friendly and economic carbon capture from
power plant flue gases: The SIEMENS PostCap technology. First Post Combustion Capture
Conference, Abu Dhabi
46. Fernandez ES, Goetheer ELV (2011) DECAB process development of a phase change
absorption process. Energy Procedia 4:868–875
47. Kumar PS, Hogendoorn JA, Timmer JS, Feron PHM (2003) Equilibrium solubility of CO2
in aqueous potassium taurate solutions: part 2. Experimental VLE data and model. Ind Eng
Chem Res 42(12):2841–2852
48. Sanchez Fernandez E, Heffernan K, van der Ham L, Linders MJG, Eggink E, Schrama FNH,
Brilman DWF, Goetheer ELV, Vlugt TJH (2013) Conceptual design of a novel CO2 capture
process based on precipitating amino acid solvents. Ind Eng Chem Res 52(34):12223–12235
49. Misiak K, Sanchez Sanchez C, van Os P, Goetheer E (2013) Next generation
post-combustion capture: combined CO2 and SO2 removal. Energy Procedia 37:1150–1159
50. Song H-J, Lee S, Maken S, Park JJ, Park JW (2006) Solubilities of carbon dioxide in
aqueous solutions of sodium glycinate. Fluid Phase Equilib 246(1–2):1–5
148 A.M. Shariff and M.S. Shaikh

51. Harris F, Kurnia KA, Mutalib MIA, Murugesan T (2009) Solubilites of carbon dioxide and
densities of aqueous sodium glycinate solutions before and after CO2 absorption. J Chem
Eng Data 54(1):144–147
52. Zhang W, Wang Q, Fang M, Luo Z (2010) Experimental Study on CO2 absorption and
regeneration of aqueous solutions of potassium glycinate. 978-1-4244-4713-8/10/$25.00
©2010 IEEE
53. Aronu UE, Hoff KA, Svendsen HF (2011) Vapor–liquid equilibrium in aqueous amine
amino acids salt solution: 3-(methylamino) propylamine/sarcosine. Chem Eng Sci 66
(17):3859–3867
54. Song HJ, Lee MG, Kim H, Gaur A, Park JW (2011) Density, viscosity, heat capacity,
surface tension, and solubility of CO2 in aqueous solutions of potassium serinate. J Chem
Eng Data 56(4):1371–1377
55. Lee JI, Otto FD, Mather AE (1974) The solubility of H2S and CO2 in aqueous
monoethanolamine solutions. Can J Chem Eng 52(6):803–805
56. Jou FY, Mather AE, Otto FD (1982) Solubility of hydrogen sulfide and carbon dioxide in
aqueous methyldiethanolamine solutions. Ind Eng Chem Process Design Develop 21
(4):539–544
57. Lim J, Kim DH, Yoon Y, Jeong SK, Park KT, Nam SC (2012) Absorption of CO2 into
aqueous potassium salt solutions of L-alanine and L-proline. Energy Fuels 26(6):3910–3918
58. Majchrowicz ME, Brilman DWF (2012) Solubility of CO2 in aqueous potassium L-prolinate
solutions-absorber conditions. Chem Eng Sci 72:35–44
59. Kang D, Park S, Jo H, Min J, Park J (2013) Solubility of CO2 in Amino-Acid-Based
Solutions of (potassium sarcosinate), (potassium alaninate + piperazine), and (potassium
serinate + piperazine). J Chem Eng Data 58(6):1787–1791
60. Aldenkamp N, Huttenhuis P, Penders-van Elk N, Hamborg ES, Versteeg GF (2014)
Solubility of carbon dioxide in aqueous potassium salts of glycine and taurine at absorber
and desorber conditions. J Chem Eng Data 59(11):3397–3406
61. Chang YT, Leron RB, Li MH (2015) Carbon dioxide solubility in aqueous potassium salt
solutions of L-proline and DL-a-aminobutyric acid at high pressures. J Chem Thermodyn
83:110–116
62. Mazinani S, Ramazani R, Samsami A, Jahanmiri A, Bruggen BV, Darvishmanesh S (2015)
Equilibrium solubility, density, viscosity and corrosion rate of carbon dioxide in potassium
lysinate solution. Fluid Phase Equilib 396:28–34
63. Shen S, Yang Y, Wang Y, Ren S, Han J, Chen A (2015) CO2 absorption into aqueous
potassium salts of lysine and proline: Density, viscosity and solubility of CO2. Fluid Phase
Equilib 399:40–49
64. Chen ZW, Leron RB, Li MH (2015) Equilibrium solubility of carbon dioxide in aqueous
potassium L-asparaginate and potassium L-glutaminate solutions. Fluid Phase Equilib
400:20–26
65. Mazinani S, Samsami A, Jahanmiri A (2011) Solubility (at low partial pressures), density,
viscosity, and corrosion rate of carbon dioxide in blend solutions of monoethanolamine
(MEA) and sodium glycinate (SG). J Chem Eng Data 56(7):3163–3168
66. Privalova E, Rasi S, Maki-Arvela P, Eranen K, Rintala J, Murzin DY, Mikkola JP (2013)
CO2 capture from biogas: absorbent selection. RSC Advances 3:2979–2994
67. Lu JG, Fan F, Lui C, Ji Y, Zhang H (2011) Performance evaluation of amino acid salt-based
complex absorbents for CO2 capture. Environ Eng Manage J 10(2):297–304
68. Park S, Song HJ, Park J (2014) Selection of suitable aqueous potassium amino acid salts:
CH4 recovery in coal bed methane via CO2 removal. Fuel Process Technol 120:48–53
69. Kumar PS, Hogendoorn JA, Versteeg GF (2003) Kinetics of the reaction of CO2 with
aqueous potassium salt of taurine and glycine. AIChE J 49(1):203–213
70. Versteeg GF, Van Dijck LAJ, Van Swaaij WPM (1996) On the kinetics between CO2 and
alkanolamines both in aqueous and non-aqueous solutions: an overview. Chem Eng
Commun 144(1):113–158
Aqueous Amino Acid Salts and Their Blends … 149

71. Lee S, Song HJ, Maken S, Park JW (2007) Kinetics of CO2 absorption in aqueous sodium
glycinate solutions. Ind Eng Chem Res 46(5):1578–1583
72. Portugal AF, Derks PWJ, Versteeg GF, Magalhãesa FD, Mendesa A (2007) Characterization
of potassium glycinate for carbon dioxide absorption purposes. Chem Eng Sci 62(23):6534–
6547
73. Glasscock DA, Critchfield JE, Rochelle GT (1991) CO2 absorption desorption in mixtures of
methyldiethanolamine with monoethanolamine or diethanolamine. Chem Eng Sci 46
(11):2829–2845
74. Portugal AF, Magalhães FD, Mendes A (2008) Carbon dioxide absorption kinetics in
potassium threonate. Chem Eng Sci 63(13):3493–3503
75. Holst JV, Versteeg GF, Brilman DWF, Hogendoorn JA (2009) Kinetic study of CO2 with
various amino acid salts in aqueous solution. Chem Eng Sci 64(1):59–68
76. Simons K, Brilman WDWF, Mengers H, Nijmeijer K, Wessling M (2010) Kinetics of CO2
absorption in aqueous sarcosine salt solutions: influence of concentration, temperature, and
CO2 loading. Ind Eng Chem Res 49(20):9693–9702
77. Vaidya PD, Konduru P, Vaidyanathan M (2010) Kinetics of carbon dioxide removal by
aqueous alkaline amino acid salts. Ind Eng Chem Res 49(21):11067–11072
78. Aronu UE, Hartono A, Hoff KA, Svendsen HF (2011) Kinetics of carbon dioxide absorption
into aqueous amino acid salt: potassium salt of sarcosine solution. Ind Eng Chem Res 50
(18):10465–10475
79. Kim M, Song HJ, Lee MG, Jo HY, Park JW (2012) Kinetics and steric hindrance effects of
carbon dioxide absorption into aqueous potassium alaninate solutions. Ind Eng Chem Res 51
(6):2570–2577
80. Paul S, Thomsen K (2012) Kinetics of absorption of carbon dioxide into aqueous potassium
salt of proline. Int J Greenhouse Gas Control 8:169–179
81. Majchrowicz ME, Kersten S, Brilman W (2014) Reactive absorption of carbon dioxide in
L-prolinate salt solutions. Ind Eng Chem Res 53(28):11460–11467
82. Guo D, Thee H, Tan CY, Chen J, Fei W, Kentish S, Stevens GW, Da Silva G (2013) Amino
acids as carbon capture solvents: chemical kinetics and mechanism of the
glycine + CO2 reaction. Energy Fuels 27(7):3898–3904
83. Sodiq A, Rayer AV, Olanrewaju AA, Abu-Zahra MRM (2014) Reaction kinetics of carbon
dioxide (CO2) absorption in sodium salts of taurine and proline using a stopped-flow
technique. Int J Chem Kinet 46:730–745
84. Fang M, Zhou X, Xiang Q, Cai D, Luo Z (2015) Kinetics of CO2 absorption in aqueous
potassium L-prolinate solutions at elevated total pressure. Energy Procedia 75:2293–2298
85. Hikita H, Asai S, Ishikawa H, Honda M (1977) The kinetics of reactions of carbon dioxide
with monoethanolamine, diethanolamine and triethanolamine by a rapid mixing method.
Chem Eng J 13(1):7–12
86. Rangwala HA, Morrell BR, Mather AE, Otto FD (1992) Absorption of CO2 into aqueous
tertiary amine/MEA solutions. Can J Chem Eng 70(3):482–490
87. Alpe E (1990) Reaction mechanism and kinetics of aqueous solutions of
2-amino-2-methyl-1-propanol and carbon dioxide. Ind Eng Chem Res 29(8):1725–1728
88. Bishnoi S, Rochelle GT (2002) Absorption of carbon dioxide in aqueous
piperazine/methyldiethanolamine. AIChE J 48:2788–2799
89. Littel R, Van Swaaij W, Versteeg G (1990) Kinetics of carbon dioxide with tertiary amines
in aqueous solution. AIChE J 36:1633–1640
90. Epp B, Hans F, Monika V (2011) Degradation of solutions of monoethanolamine,
diglycolamine and potassium glycinate in view of tail-end CO2 absorption. Energy Procedia
4:75–80
91. Huang Q, Saloni B, Joseph ER, John P, Selegue KL (2013) Thermal degradation of amino
acid salts in CO2 capture. Int J Greenhouse Gas Control 19:243–250
92. DuPart MS, Bacon TR, Edwards DJ (1993) Hydrocarbon Processing 72(5):89
93. Hawkes EN, Mago BF (1971) Hydrocarbon Process 50(8):109
150 A.M. Shariff and M.S. Shaikh

94. Ahn S, Song HJ, Park JW, Lee JH, Lee IY, Jang KR (2010) Characterization of
metal corrosion by aqueous amino acid salts for capture of CO2. Korean J Chem
Eng 27(5):1576–1580
95. Bello A, Idem RO (2006) Comprehensive study of the kinetics of the oxidative degradation
of CO2 loaded and concentrated aqueous monoethanolamine (MEA) with and without
sodium metavanadate during CO2 absorption from flue gases. Ind Eng Chem Res 45
(8):2569–2579
96. Budzianowski WM (2015) Single solvents, solvent blends, and advanced solvent systems in
CO2 capture by absorption: a review. Int J Global Warming 7(2):184–225
97. Shao R, Stangeland A (2009) Amines used in CO2 capture-health and environmental
impacts. Bellona Report, September 2009
98. Glycine; MSDS Cat. code. SLG1972, SLG2191; [online]; Science Lab.Com, Inc 14025
Smith Road, Houston, Texas 77396USA. http://www.sciencelab.com/msds.php?msdsId=
9927179. Accessed 16 Dec 15
99. L-proline; MSDS Cat code. SLP5488, SLP2662, SLP4471, Science Lab.Com, Inc 14025
Smith Road, Houston, Texas 77396 USA. http://www.sciencelab.com/msds.php?msdsId=
9927237. Accessed 16 Dec 15
100. Taurine; MSDS Cat code. SLT2014, Science Lab.Com, Inc 14025 Smith Road, Houston,
Texas 77396 USA. http://www.sciencelab.com/msds.php?msdsId=9925166. Accessed 16
Dec 15
101. L-Histidine; MSDS Cat code. SLH2317, SLH3163, SLH1843, Science Lab.Com, Inc 14025
Smith Road, Houston, Texas 77396 USA. http://www.sciencelab.com/msds.php?msdsId=
9927189. Accessed 16 Dec 15
102. L-Methionine; MSDS Cat code. SLM1987, SLM3361, SLM1216, Science Lab.Com, Inc
14025 Smith Road, Houston, Texas 77396 USA. http://www.sciencelab.com/msds.php?
msdsId=9927225. Accessed 16 Dec 15
103. L-Glutamine; MSDS Cat code. SLG1462, SLG1963, Science Lab.Com, Inc 14025 Smith
Road, Houston, Texas 77396 USA. http://www.sciencelab.com/msds.php?msdsId=9924160.
Accessed 16 Dec 15
104. Monoethanolamine; MSDS Cat code. SLA4792, SLA2452, SLA3955, Science Lab.Com,
Inc 14025 Smith Road, Houston, Texas 77396 USA. http://www.sciencelab.com/msds.php?
msdsId=9922885. Accessed 16 Dec 15
105. Diethanolamine; MSDS Cat code. SLD1834, SLD3199, Science Lab.Com, Inc 14025 Smith
Road, Houston, Texas 77396 USA. http://www.sciencelab.com/msds.php?msdsId=9923743.
Accessed 16 Dec 15
106. Triethanolamine; MSDS Cat code. SLT3841, SLT1822, Science Lab.Com, Inc 14025 Smith
Road, Houston, Texas 77396 USA. http://www.sciencelab.com/msds.php?msdsId=9927306.
Accessed 16 Dec 15
107. Piperazine; MSDS Cat code. SLP4681, Science Lab.Com, Inc 14025 Smith Road, Houston,
Texas 77396 USA. http://www.sciencelab.com/msds.php?msdsId=9926575. Accessed 16
Dec 15
108. Gabrielsen J (2007) CO2 capture from coal fired power plants, IVCSEP. Ph.D. Thesis,
Technical University of Denmark
109. Jockenhoevel T, Schneider R, Rode H (2008) Development of an economic post-combustion
carbon capture process, GHGT-9-Washington D.C, USA
110. Weiland RH, Hatcher NA, Nava JL (2010) Post-combustion CO2 capture with amino-acid
salts, Optimized Gas Treating, Inc. Clarita, 74535, USA
111. Salazar V, Sanchez-Vincente Y, Pando C, Renuncio JAR, Cabanas A (2010) Enthalpies of
absorption of carbon dioxide in aqueous sodium glycinate solutions at temperatures of
(313.15 and 323.15) K. J Chem Eng Data 55(3):1215–1218
112. Oexmann J, Kather A (2010) Minimizing the regeneration heat duty of post-combustion
CO2 capture by wet chemical absorption: the misguided focus on low heat of absorption
solvents. Int J Greenhouse Gas Control 4(1):36–43
113. Amino acids formulas and molecular weight. Accessed 19 Dec 15
Aqueous Amino Acid Salts and Their Blends … 151

114. Walum E (1998) Acute oral toxicity. Environ Health Perspect 106(2):497–503
115. Shen KP, Li MH (1992) Solubility of carbon dioxide in aqueous mixtures of
monoethanolamine with methyldiethanolamine. J Chem Eng Data 37(1):96–100
116. Tan LS, Shariff AM, Lau KK, Bustam MA (2012) Factors affecting CO2 absorption
efficiency in packed column: a review. J Ind Eng Chem 18(6):1874–1883
Ionic Liquids: Advanced Solvents for CO2
Capture

Xiangping Zhang, Lu Bai, Shaojuan Zeng, Hongshuai Gao,


Suojiang Zhang and Maohong Fan

Abstract As one of promising advanced solvents, ionic liquids (ILs) have become
more attractive for CO2 capture due to their unique properties, special structures and
potential energy saving efficiency. This chapter mainly reviews the research pro-
gress on CO2 capture with ILs, focusing on the CO2 absorption capacity of con-
ventional ILs, task-specific ILs and ILs based mixtures as well as the comparison
and analysis. The influence of cations, anions and functional groups of ILs on the
CO2 absorption was analyzed and the mechanisms of physisorption and
chemisorption were revealed using experimental test and molecular simulation
results. Especially considering the real applications of the new ILs-based capture
technologies, the research on process simulation and assessment of CO2 capture
processes was also reviewed. Finally, we discussed the challenges and opportunities
of transferring the lab-scale research results to practical industries processes, and
also present some perspectives of ILs based novel technologies.

1 Introduction

A large amount of greenhouse gases in atmosphere, especially CO2 emitted from


the fossil fuel combustion in the industrial processes, has contributed to serious
global warming problems, which will lead to climate warming, sea level rising and
aggravated disasters. CO2 capture and storage (CCS) has been regarded as one of
the most promising options to reduce CO2 emissions. In CCS, CO2 separation is
one of the most critical parts [1, 2]. Additionally, in the field of energy gas

X. Zhang (&)  L. Bai  S. Zeng  H. Gao  S. Zhang


State Beijing Key Laboratory of Ionic Liquids Clean Process, State Key Laboratory
of Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy
of Sciences, Beijing, China
e-mail: xpzhang@ipe.ac.cn
M. Fan
Department of Chemical and Petroleum Engineering, School of Energy Resources,
University of Wyoming, Laramie, WY, USA

© Springer International Publishing AG 2017 153


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_7
154 X. Zhang et al.

resources, like natural gas, coal bed gas, shale gas, biogas, and so on, CO2 as
impurities in these gases should be removed before utilizations because CO2
components in these gases might decrease the heating values, demand high energy
consumption for transport and cause pipeline corrosions [3]. Thus, CO2 separation
has been received growing attentions due to its importance in environmental pro-
tection and industrial application. At present, there are already many technologies
for CO2 capture, like physical or chemical solvent scrubbing, cryogenic distillation,
pressure/temperature swing adsorption (P/TSA), gas membrane separation and so
on. Although these technologies have their own advantages, most of them still
suffer from some drawbacks. For examples, amine scrubbing by chemical reaction
is the widely used method for CO2 capture in industry because of the excellent
absorption performances even under low partial pressure of CO2, but there are some
inherent shortcomings, such as the loss of solvent due to thermal or chemical
degradation, equipment corrosion and high energy demands during the solvent
regeneration [4, 5]. Cryogenic distillation requires substantial energy of refrigera-
tion in real applications for CO2 separation [2]. Solid adsorbents used in the process
of adsorption, like activated carbons, zeolites, metal organic frameworks, microp-
orous organic polymers etc., show some advantages for CO2 capture, but the
adsorption capacity, selectivities, stabilities, recycling, etc. of adsorbents still need
to be improved for industrial applications. The membrane gas separation is another
potential technology for CO2 capture because of less equipment space, lower
energy requirement and absence of potentially hazardous chemicals; nevertheless,
the compromise between permeability and selectivity of membranes limits its
large-scale applications [2], as well as their higher costs. Thus, developing novel
solvents/materials and corresponding new technologies for CO2 capture is an
enduring R&D topic and highly desirable.
Ionic liquids (ILs) as advanced solvents have been regarded as prospective
candidates for CO2 capture because of their excellent properties and potential
energy saving efficiency, and being paid more and more attentions in recent years.
ILs are well known as fluid at room temperature and composed entirely of organic
cations and organic/inorganic anions. They have some unique properties, including
negligible vapor pressures, low melted points, high thermal stabilities and tunable
structures, which make them be widely used in many areas, such as organic syn-
thesis, catalysis, electrochemistry, biochemistry, gas separation, especially for CO2
capture. When capturing CO2 with ILs, the negligible vapour pressure of ILs makes
no contamination in gas stream and negligible losses of ILs. More importantly, due
to the negligible vapour pressure of ILs, stripping CO2 from rich solvent could be
carried out by flash operation instead of distillation tower used in amine tech-
nologies, which needs much less energy consumption and lower equipment
investment. For example, it was evaluated that CO2 capture with [bmim][Ac] IL can
reduce the energy losses by 16 % and the economic investment will be 11 % lower
compared to a MEA-based process [6]. It was also investigated by simulations that
the process of IL-amine hybrid solvents ([Bpy][BF4]-MEA) for CO2 capture can
save about 15 % regeneration heat duty compared to a MEA process [7].
Additionally, ILs are designable solvents, thus the physicochemical properties
Ionic Liquids: Advanced Solvents for CO2 Capture 155

could be tuned for aimed applications through structural modifications of cations


and anions and/or changing the combination of cations and anions. Therefore, ILs
as designable solvents according to the real applications is potential option for
energy and cost efficient capturing CO2.
Since it was firstly reported that CO2 could be efficiently dissolved in ILs [8], a
series of conventional ILs with physical absorption and task-specific ILs with
chemical absorption were reported in succession [9–12]. ILs captured CO2 by
physical interaction as result of the special internal microstructure and force of ILs.
According to ATR-IR spectrum, the high CO2 solubility was achieved by the Lewis
acid-base interaction between the anion of IL and CO2 [13, 14], as well as the
hydrogen-bonded interactions between the H proton of C2 atom on the imidazole
ring and CO2 molecule [15]. Quantum chemistry and Molecular dynamics simu-
lation have indicated that the cations and anions form the hydrogen bonding net-
work in the imidazolium-based ILs and the molecules of CO2 can fill in the network
formed by the cations of ILs so that the CO2 dissolves in ILs with high solubility
[16]. Generally, CO2 solubility of ILs is mainly depended on the anion, hence the
hydrogen-bonded interactions induce to a high CO2 solubility. Yu et al. [17] found
that more charge-localized character of [TMG][L], especially the C1 carbocation on
[TMG]+, and the intermolecular-NH2-associated hydrogen bonds can substantially
increase the cation-anion interaction, and the interaction energy of [TMG]L is 65.3–
109.3 kJ/mol higher than some [bmim]+ based ILs, thus the CO2 solubility in
[TMG]L is higher than that in [bmim][PF6]. In order to further improve CO2
absorption capacity, task-specific ILs were designed and developed subsequently
[18–21]. Among these ILs, amino-functionalized ILs were well investigated, for
example, the solubility in [aemmim][Tau] reached 0.9 mol CO2/mol IL [20].
Although significant progress has been made, the gravimetric capacities of these ILs
which is much more interested in real industries, are usually lower than 0.1 g CO2/g
IL [9], so later researchers focused on the improvement of CO2 gravimetric
capacities in ILs. Wang et al. [22] synthesized a series of superbase-derived protic
ILs, and the gravimetric capacities of [MTBDH][Im] reached 0.205 g CO2/g IL, but
the thermal stability of ILs need to be improved. Our research group also developed
a new dual amino-functionalized and the gravimetric capacities can reach 18 %
[23]. Besides the CO2 capacities, the viscosity of IL solvents is another important
property. Generally, the viscosity of task-specific ILs is relatively high and the
viscosity increases after absorption due to the chemical reaction, for example, the
viscosity of [aP4443][AA] increased nearly threefold after absorption of CO2 [21],
and the viscosity of [P66614][Isoleucinate] increased over 200-fold when exposed to
1 bar of CO2 [24]. Therefore, the development of new ILs with low viscosity is
extremely desirable. It was reported that the viscosity of [P66614][2-CNpyr] is less
than 100 cP at 50 °C before exposure to CO2 and changes slightly after absorption
of CO2, which is superior to the amino-functionalized ILs [24]. Additionally,
mixing ILs with other solvents is an efficient way to offset the inherently high
viscosities of ILs and also reduce the energy requirement. For instance, by mixing
ILs with organic amine solvent, which only needs to break the binding energy
between the solvent and CO2, and takes few energy to vapor water, so it is regarded
156 X. Zhang et al.

as one of the promising technology [25]. It was also found that the absorption
capacities had nearly no change when ILs was mixed with solvents, while the
viscosity decreased significantly and the whole energy consumption decreased
about 20 % [10].
In summary, this chapter aims to introduce the research progress on CO2 capture
with ILs, including conventional ILs, task-specific ILs and ILs based mixtures. The
CO2 absorption capacity is summarized and compared, meanwhile, the relationship
between structures and properties of ILs and CO2 absorption mechanism with
experimental characteristic and molecular simulation is discussed in detail. The
process simulation and assessment of CO2 capture with ILs was further reviewed
for the practical application. Some challenges and perspectives of ILs based tech-
nology for CO2 are also presented.

2 Conventional Ionic Liquids for CO2 Separation

The conventional IL [bmim][PF6] for CO2 capture was firstly reported and it can
efficiently and physically absorb 0.75 mol CO2/mole IL at 25 °C and 8.3 MPa. [8].
An interesting result was found that a large quantity of CO2 dissolved in the IL
phase, while less IL existed in the CO2 phase [26], showing that ILs has potential
ability for CO2 separation and receive extensive attentions in recent years. The
conventional ILs usually absorb CO2 through the physical interaction between the
cations/anions and CO2, e.g. electrostatic interaction, van der Waals forces,
hydrogen bonds. Therefore, the structures of cations and anions of ILs play an
important role in CO2 solubility. The CO2 absorption capacities of ILs with
physisorption have been concluded and tabulated in our previous review (Energy
Environ. Sci., 2012, 5, 6668–6681, Table 1).

2.1 The Effect of Cations on CO2 Absorption

Imidazolium-based IL is most widely investigated and reported in literature for CO2


capture [8, 13, 27–29]. Kazarian et al. [13] thought that the high solubility of CO2
in imidazolium-based ILs is mostly attributed to the hydrogen bonding between the
acidic hydrogen on the imidazolium and CO2. In order to understand the effect of
the hydrogen in the C2 position of imidazolium on CO2 solubility, Brenneck et al.
[27] studied three pairs of ILs: 1-n-butyl-3-methylimidazolium hexafluorophos-
phate ([bmim][PF6]) and 1-n-butyl-2,3-dimethylimidazolium hexafluorophosphate
([bmmim][PF6]);1-n-butyl-3-methylimidazolium tetrafluoroborate ([bmim][BF4])
and 1-n-butyl-2,3-dimethylimidazolium tetrafluoroborate ([bmmim][BF4]); and
1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([emim][Tf2N])
and 1-ethyl-2,3-dimethylimidazolium bis(trifluoromethylsulfonyl)imide ([emmim]
[Tf2N]). As shown in Fig. 1, there is some decreased solubility for the
Ionic Liquids: Advanced Solvents for CO2 Capture 157

Fig. 1 Solubility of CO2 in


[emim][Tf2N] (filled square);
[emmim][Tf2N] (open
square);[bmim][PF6] (filled
circle); [bmmim][PF6] (open
circle); [bmim][BF4] (filled
triangle); [bmmim][BF4]
(open triangle) at 25 °C [27]

methyl-substituted IL relative to the hydrogen-substituted IL in each pair, implying


the weak influence of the cations on CO2 solubility.
Furthermore, CO2 solubility in [bmim][Tf2N], [hmim][Tf2N] and [omim][Tf2N]
was measured, and the influence of the length of alkyl chains on CO2 solubility was
investigated by Aki et al. [30]. The results indicated that CO2 solubility marginally
increases with an increase of the alkyl chain length on the cation. The similar results
were obtained from the Yunus’s work [31]. Huang et al. [28] found that CO2
perhaps takes up free space from cavities already available in the rigid and intricate
topography of ILs when CO2 physically dissolves in ILs, but there are no suffi-
ciently large cavities to accommodate CO2 in neat ILs, so subtle rearrangements of
ILs take place to create enough spaces, which results in no obvious volume
expansion during the addition of CO2 and a slight increase of the densities of ILs
(Fig. 2). Therefore, the increase of the side chain length on the cation can result in
the greater free volume, which can increase CO2 solubility in ILs.
In addition, some functional groups are introduced into the cations of ILs to
increase CO2 solubility. Almantariotis et al. [29] studied the effect of fluorination on
CO2 solubility in ILs, and found that the higher solubility of CO2 in the
fluorine-substituted IL [C8H4F13mim][Tf2N] than the normal IL [C8mim][Tf2N] is
due to the larger free volume of the fluorine-substituted IL and the stronger inter-
action between CO2 and the fluorinated alkyl chains (Fig. 3). In fact, the [Cnmim]-
based ILs do not possess both high CO2 solubility and ideal solubility selectivities.
If the cations of ILs contain functional groups such as ethers, nitriles, the CO2
solubility in these ILs was similar to that in their analogues, but the solubilities of
N2 and CH4 were much lower in these ILs. Therefore, the higher CO2/N2 and CO2/
CH4 selectivity could be obtained by using these ILs [25, 32, 33].

2.2 The Effect of Anions on CO2 Absorption

A large number of experimental data and simulation calculations demonstrated that


the anions of ILs play a greater role in CO2 solubility in conventional ILs than
158 X. Zhang et al.

cations. Anthony et al. [34] investigated systematically the influence of cations


(such as imidazolium, ammonium, pyrrolidinium, and phosphonium) and anions
(Tf2N, PF6 and BF4) on CO2 solubility. When the anion is the same, the types of
cations have a slight influence on CO2 solubility. In contrast, the influence of anions
is greater, and the IL with the [Tf2N] anion has a considerably higher affinity to CO2
than either of the [BF4] IL and [PF6] IL, implying that the dominant role of the
anion in CO2 dissolution. Brennecke et al. investigated CO2 solubility in the ILs
with the same cation ([bmim]) and several different anions. The results indicated
that the solubility of CO2 in [bmim] cation based ILs increases in the following
order: [NO3] < [DCA] < [BF4] < [PF6] < [TfO] < [Tf2N] < [Methide] at 25 °C
[30, 35]. Maiti and Sistla et al. also confirmed the results by the COSMO-RS
method [13, 14]. Sergei et al. studied the mechanism of CO2 dissolution in ILs with
the [BF4] and [PF6] anions and CO2 by in situ ATR-IR spectroscopy. The results
indicated that the interaction is a Lewis acid-base type where the anion serves as a
Lewis base, while CO2 acts as a Lewis acid, and the interaction between CO2 and
[BF4] anion should be stronger than that for [PF6] anion, since [BF4] is a stronger
base. However, experimental data showed a higher CO2 solubility in [bmim][PF6]

Fig. 2 Structural rearrangement of ILs due to CO2 dissolution [28]

Fig. 3 Effect of fluorination


groups on CO2 solubility [29]
Ionic Liquids: Advanced Solvents for CO2 Capture 159

rather than [bmim][BF4]. Thus, CO2 solubility in ILs could not solely be explained
by anion-CO2 interactions [13]. In addition, CO2 solubility is higher in the ILs with
fluorine groups in the anion than that in the ILs with nonfluorinated anions. It is to
say that the CO2 solubility will be increased with the number of fluorine groups in
the anion increases. For example, CO2 solubility increases in order of the anions:
[BF4] < [PF6] < [Tf2N] < [Methide] < [eFAP] < [bFAP] [30, 35]. Zhang et al.
[36] using COSMO-RS further predicted that the longer fluoroalkyl chain in the
anion (e.g. FAP anion), the higher CO2 solubility obtained. Bhargava et al. [37]
further studied the interaction of CO2 molecules with various anions of ILs using
density functional theory. As shown in Fig. 4, the optimized structures of
anion-CO2 complexes appeared to be dominated by Lewis acid-base interactions,
but the larger anions such as Tf2N possesses two or three preferred sites for binding
CO2 that could be related to their ability for CO2 dissolution, implying that the free
volume of ILs also plays a significant role in dissolving CO2. The experimentally
obtained data of CO2 solubility in ILs suggests a relationship between the molar
volume of the anion and the solubility [38].

3 Task-Specific Ionic Liquids for CO2 Separation

Although the emerging of ILs provides a new way for CO2 capture, CO2 physical
solubility in conventional ILs is much lower than current commercially available
solvents. Therefore, the design and development of the task-specific ILs for highly
efficient capture of CO2 becomes a hot topic. The task-specific ILs also called
functionalized ILs, include single amino ILs, dual amino ILs and non-amino ILs,
which absorb CO2 through chemical interaction between the basic groups and CO2.
The CO2 absorption capacities of ILs with chemisorption have been concluded and
tabulated in our previous review (Energy Environ. Sci., 2012, 5, 6668–6681,
Table 2). Except for these, some new ILs for CO2 capture with chemisorption are
also reviewed in this part.

3.1 Single Amino ILs

Bates et al. [39] firstly reported the single amino-functionalized IL containing an


amine group on the cation, which is capable of reversibly capture CO2 with a high
capacity of nearly 0.5 mol CO2/mol IL. This is due to the IL could react with CO2
and form carbamate (Fig. 5), the stoichiometry of IL is similar to that of the
aqueous MEA system (one CO2 molecule reacts with two amines). Subsequently,
Zhang et al. [18] reported a series of single amino-functionalized ILs containing an
amine group on the anions, and loaded them on porous silica gel for CO2 capture
due to their high viscosity, and then 0.5 mol of CO2 per mol IL could be captured.
The possible mechanism of chemical absorption of CO2 is that the CO2 is attacked
160 X. Zhang et al.

Fig. 4 The interaction between CO2 and anions: a NO3−, b BF4−, c N(CN)−, d CH3COO−, e PF6−
and f Tf2N− [37]

by the free electron pair of the N atom on the NH2 group and thus forms a hydrogen
bond with another NH2 group. Gurkan et al. [40] also reported two single
amino-functionalized ILs containing an amine group on the anion. In their work, the
ILs showed very high capacity of 0.9 mol of CO2 per mole of IL (Fig. 6).
The introduction of NH2 groups into the imidazolium ring can effectively
improve CO2 capacity, but not for all the cations with NH2 groups. Zhang et al. [41]
studied CO2 absorption performances in the IL 1,1,3,3-tetramethylguanidinium
Ionic Liquids: Advanced Solvents for CO2 Capture 161

lactate ([TMG]L) with NH2 groups on the cation. [TMG]L has very low CO2
physical solubility of only 0.25 wt%, which does not comply with the absorption
molar ratio of 1:2 between CO2 and NH2 group if it follows the same mechanism as
[NH2p-bim][BF4] [42]. The underlying reason is the large FMO energy gap
(9.53 eV) between HOMO-5 of [TMG]L and LUMO of CO2, which is much larger
than the energy gap (6.07 eV) between HOMO of [NH2p-bim][BF4] and LUMO of
CO2 as shown in Fig. 7. It is the carbocation that lowers the HOMO-5 energy of
[TMG]L and weakens its nucleophilicity; as a result, [TMG]L cannot effectively
interact with CO2 [43].

3.2 Dual Amino ILs

In order to enhance CO2 absorption capacity of amino-functionalized ILs, dual


amino groups are tethered to the cations or anions of ILs. Zhang et al. [19]
developed a series of dual amino-functionalized phosphonium ILs containing two
amine groups on the anion and cation, respectively. Xue et al. [20] also reported a
dual amino IL with amino-functionalized imidazolium cation and taurine anion.
The CO2 absorption capacities of them were found to approach 1 mol of CO2 per
mole of IL due to the reaction of both the cation and the anion with CO2. When the
amine is tethered to the cation of the ILs, the amine will react with CO2 according to
1:2 mechanism. When the amine is tethered to the anion of the ILs, the amine will

Fig. 5 Proposed reaction mechanism of CO2 with [NH2p-bim][BF4] [39]

Fig. 6 Proposed reaction mechanism of CO2 with [P66614][Pro] [40]


162 X. Zhang et al.

Fig. 7 The HOMO and LUMO energies for [NH2p-bim][BF4], [TMG]L and CO2 [43]

react with CO2 according to 1:1 mechanism. Therefore, for the dual
amino-functionalized ILs containing two amine groups on the anion and cation, the
theoretical absorption capacity should be about 1.5 mol of CO2 per mole of IL. Xue
et al. [20] thought the reason is that the R-N+H2COO− formed by the amine tethered
anion reacts with CO2 is not stable, it can react with another amine to form
R-NHCOO− (Fig. 8). Zhang et al. [23] reported a dual amino-functionalized IL
containing two amine groups tethered to the cation of IL. The CO2 absorption
capacities of this IL could be up to 1.05 mol of CO2 per mole of IL. FTIR and
NMR spectra proved that the IL could react with CO2 and form carbamate on the
basis of 1:2 stoichiometry (Fig. 9).

3.3 Non Amino ILs

Although the amino-functionalized ILs could improve the CO2 absorption capac-
ities by the reaction of amine in the cation or the anion with CO2, the viscosity of
the amino-functionalized ILs is usually too high, which restricts their further
applications. Moreover, the strong interaction between ILs and CO2 increases the
difficulty of the regeneration. Wang et al. [22, 44–47] proposed a novel strategy for
equimolar CO2 capture by a series of non-amino functionalized ILs. Firstly, Wang
et al. [22] synthesized several superbased-derived protic ILs by neutralization of a
superbase with weak proton donors, such as fluorinate alcohols, imidazoles, pyr-
idines. The results showed that [MTBDH][TFE] and [MTBDH][Im] has very high
absorption capacity of 1.13 and 1.03 mol CO2/mol IL at 23 °C and 0.1 MPa,
Ionic Liquids: Advanced Solvents for CO2 Capture 163

NH2 NHCOO
2 O3S + CO2 SO3
O3S
H3N

Fig. 8 Proposed reaction mechanism of CO2 with the anion [Tau] [20]

Fig. 9 Proposed reaction mechanism of CO2 with DAIL-Br [23]

respectively. Due to the weak proton donor was deprotonated, the ILs will have a
thermodynamic driving force for the reaction with CO2. The mechanism of CO2
capture by these ILs was also proved by spectroscopic investigation and quantum
chemical calculations (Fig. 10).
Subsequently, Wang et al. [46] designed and prepared phenolic ILs for the
efficient and reversible capture of CO2 from phosphonium hydroxide and substi-
tuted phenols. The results showed that [P66614][4-Me-PhO] and [P66614][4-Cl-PhO]
exhibited the high absorption capacity of 0.91 and 0.95 mol CO2/mol IL. Recently,
Wang et al. [47] also proposed another strategy for improving CO2 capture by new
anion-functionalized ILs making use of multiple site cooperative interactions. An
extremely high capacity of CO2 is up to 1.60 mol CO2/mol IL and excellent
reversibility is achieved. The plausible mechanism of CO2 absorption is shown in
Fig. 11.

4 Ionic Liquid Mixed Solvents for CO2 Separation

Although ILs are proposed as potential solvents for CO2 separation due to their
unique properties, there are also some other problems which may limit their
industrial applications, such as higher viscosity of ILs than that of other molecular
solvents. It is reported that the viscosity of ILs could decrease greatly by adding a
small amount of water or organic solvent [48]. Therefore, the mixture of ILs with
water or organic solvents as a new kind of absorbent for CO2 separation has paid
much attention by researchers. At present, the research on ILs mixed solvents for
CO2 capture mainly focused on IL-water solutions, IL-alkanolamine blends,
IL-organic solvents and IL-IL mixtures. The CO2 absorption capacities of some
different IL mixtures are listed in Table 1.
164 X. Zhang et al.

Fig. 10 CO2 absorption by the anions of superbased-derived protic ionic liquids [22]

Fig. 11 CO2 reaction mechanism by [P66614][2-Op] through multiple-site interactions [47]

4.1 IL-Water Systems

The presence of water in ILs could significantly affect the physical properties of ILs
especially viscosity, even a very small amount of water could result in sharp
decrease of viscosity. The first research about the influence of water on the solu-
bility of CO2 in two hygroscopic ILs butylmethylimidazolium nitrate ([Bmim]
[NO3]) and hydroxypropylmethylimidazolium nitrate ([Hopmim][NO3]) was
reported by Bermejo et al. [49]. It was found that when the water concentration is
low, the CO2 solubility in [Bmim][NO3]-water solution is slightly higher than that
in water-free [Bmim][NO3]. However, when the water concentration is high, it will
lead to lower CO2 solubility. It is mainly explained by the positive excess molar
volumes of the aqueous solutions at the working conditions. Later, the investigation
of other IL-water systems for CO2 separation also showed the existed water in ILs
could extremely reduces the viscosity while cause a slight decrease of CO2
absorption capacity [50, 51, 61]. Ventura et al. [50] found that the solubility of CO2
in the aqueous (tri-iso-butyl(rnethyl)phosphonium tosylate) [iBu3MeP][TOS] sys-
tem increase with the increase of IL molar composition revealing a salting-in effect
promoted by the IL. Goodrich et al. [61] also reported that with 14 wt% water, the
CO2 capacity of trihexyl(tetradecyl)phosphonium prolinate ([P66614][Pro]) is
reduced by approximately 0.2 mol of CO2 per mole of IL at 0.25 bar and by
0.1 mol of CO2 per mole of IL at 1 bar. In addition, Wang et al. [51] found that
triethylbutylammonium acetate ([N2224][CH3COO]) has a little larger absorption
capacity than its water complexes due to the different reaction mechanisms.
Ionic Liquids: Advanced Solvents for CO2 Capture 165

Table 1 The CO2 capacities of ILs mixturesa


Ionic liquids mixture CO2 absorption Conditions References
capacity bar/K
molCO2/mol IL Lit.
data
[Bmim][NO3] + H2O *0.1 1.53/323 [49]
(Rmo = 98.01:1.99)
[Hopmim][NO3] + H2O *0.099 2.33/318 [49]
(Rmo = 95.89:4.11)
[iBu3MeP][TOS] + H2O (Rmo = 4:96) *0.030 4.5/289 [50]
[N2224][CH3COO] + H2O (Rmo = 1:2) 0.197 0.1/298 [51]
[P66614][2-CNPyr] + H2O 0.9 0.1/295 [52]
(Rma = 95.5:4.5)
[Hmim][NTf2] + MEA (Rmo = 1:1) 0.5 0.1/313 [53]
[N1111][Gly] + MDEA (Rma = 1:1) 0.56 0.097/298 [54]
[N2222][Gly] + MDEA (Rma = 1:1) 0.64 0.097/298 [54]
[N1111][Lys] + MDEA (Rma = 1:1) 0.69 0.097/298 [54]
[N2222][Lys] + MDEA (Rma = 1:1) 0.74 0.097/298 [54]
[C3OHmim][Cl] + MEA (Rmo = 1:2) 0.396 0.1/308 [55]
[C2OHmim][DCA] + MEA (Rma = 1:3) 0.638 0.185/313 [56]
[Bmim][DCA] + MEA (Rma = 1:3) 0.652 0.233/313 [56]
[Choline][Pro] + PEG 200 (Rma = 1:1) 0.520 0.107/338 [57]
[Emim][Ac] + [Emim][TFA] 0.124 0.1/323 [58]
(Rmo = 49.98:50.02)
[Emim][BF4] + [Omim][NTf2] 0.32 2.18/313 [59]
(Rma = 1:1)
[Bmim][BF4] + [Omim][NTf2] 0.34 2.22/313 [59]
(Rma = 1:1)
[Emim][EtSO4] + [Emim][Ac] 0.19 3.93/298 [60]
(Rmo = 1:1)
a
Rma mass ratio, Rmo mole ratio

Without water, [N2224][CH3COO] can only absorb CO2 via a Lewis acid and base
reaction, while reactions involving [N2224][CH3COO]-nH2O could easily reach the
CO2 saturation because of the formation of a stable acetic acid-H2O compound.
Recently, enhanced CO2 capture in a system of 1-alkyl-3-methylimidazolium tri-
cyanomethanide ([CnC1im][TCM]) and H2O was reported by Romanos [62]. It was
found that the lower water concentrations lead to a reduction of the solubility of
CO2 in the hybrid solvent. However, an increment of water content resulted in an
enhancement of CO2 absorption in the mixture of [CnC1im][TCM] + H2O system
compared to the dry ILs, in contrast to the detrimental influence of water on the
CO2 solubility for most ILs. A molecular exchange mechanism between CO2 in the
gas phase and H2O in the liquid phase was used to explain the enhanced CO2
absorption in the hybrid solvents. When the concentration of water reaches a critical
value, the interaction between IL and H2O molecules will be broken and H2O will
166 X. Zhang et al.

be replaced by CO2. A series of trihexyl(tetradecyl)phosphonium 2-cyano-pyrrolide


([P66614][2-CNPyr]) + H2O mixtures below the saturation limit was studied for
CO2 absorption and showed that an increase of H2O content leads to a slightly
increased solubility of CO2 and the shape of the isotherm was changed dramatically
[52]. The enhanced CO2 solubility could be ascribed to the changes in the activity
of the IL-CO2 complex upon addition of water. The mixture of ILs based on
carboxylate anions formulated with water with a high chemical absorption for CO2
were investigated [63]. The authors thought that the high solubility of CO2 is
attributed to the basicity of the ILs anions activates the reaction between water and
CO2 to form the hydrogencarbonate anion and the conjugate acid of the anion.
Adding a certain amount of water into the ILs will slightly decrease the solubility
of CO2 for most of ILs, but will slightly increase the solubility of CO2 for certain
functionalized ILs. The affect of water on the absorption mechanism of IL + water
should be investigated deeply in the future. Additionally, the long term stability of
the IL + water should be considered for industrial application because the water
could lead to the hydrolysis of ILs at high temperature.

4.2 IL-Alkanolamine Systems

Due to the fact that CO2 solubility in conventional ILs is not satisfactory, thus
functionalized ILs with high CO2 capacity have been developed. However, because
of the higher viscosity and complicated synthetic and purification steps compared to
conventional ILs, the functionalized ILs do not appear to be viable for industrial
process. In 2008, the idea of mixing ILs and alkanolamines for CO2 capture was put
forward by Camper et al. [53]. In their study, the 1-hexyl-3-methyl-imidazolium bis
(trifluoromethylsulfonyl)imide ([Hmim][NTf2]) solutions containing 50 mol% are
able to rapidly and reversibly capture 1 mol of CO2 per 2 mol monoethanolamine
(MEA) to give an insoluble MEA-carbamate precipitate that helps to drive the
capture reaction. The desirable properties of ILs and high performance of alka-
nolamines for CO2 may be incorporated and energy can be saved during the
regeneration process without affecting the absorption performance. Ahmady et al.
[64] studied the solubility of CO2 in the aqueous mixture of N-methyldiethanolamine
(MDEA) with 1-butyl-3-methyl-imidazolium tetrafluoroborate ([Bmim][BF4]). The
results showed that the presence of a low concentration of [Bmim][BF4] in aqueous
MDEA has no significant effect on the mixture loading capacity, but increased the
initial absorption rate. The CO2 loading decreased with increasing [Bmim][BF4]
concentration in the mixture due to a lack of water at high concentrations of [Bmim]
[BF4]. They also investigated the solubility of CO2 in the aqueous mixture of MDEA
with other two types of ILs, 1-butyl-3-methyl-imidazolium acetate ([Bmim]
[CH3COO]) and 1-butyl-3-methyl-imidazolium dicyanamide ([Bmim][DCA]), and
found that CO2 loading decreased significantly as the ILs concentration increased
[65]. Four functionalized amino acid ILs: tetramethylammonium glycinate ([N1111]
[Gly]), tetraethylammonium glycinate ([N2222][Gly]), tetramethylammonium
Ionic Liquids: Advanced Solvents for CO2 Capture 167

lysinate ([N1111][Lys]) and tetraethylammonium lysinate ([N2222][Lys]) were mixed


with MDEA aqueous solutions to capture CO2 by Zhang et al. [54, 66, 67]. The
results suggested that ILs could greatly enhance both the absorption and absorption
rate of CO2 in MDEA aqueous solutions, because the ILs could capture CO2 through
chemical absorption. Huang et al. [55] found that the chloride ions could signifi-
cantly enhance the capacity and thermal stability of CO2 captured by MEA in
hydroxyl imidazolium based ILs. This phenomenon may be caused by the hydrogen
bonding and electrostatic attraction between the chloride and the cationic species.
Sairi et al. [68] investigated the systems of guanidinium trifluoromethanesulfonate
([gua][OTf]) and MDEA for absorption of CO2., which indicated that the addition of
[gua][OTf] to MDEA solution leads to a slight decrease for CO2 solubility. Yang
et al. [69] used the mixed IL-amine solution (30 wt% MEA + 40 wt% [Bmim]
[BF4] + 30 wt% H2O) to capture CO2, and showed that the energy consumption of
the hybrid system for absorbent regeneration was 37.2 % lower than that of aqueous
MEA solution. The MEA loss per ton of captured CO2 for the mixed solution was
1.16 kg, which is much lower than that of 3.55 kg for the aqueous amine solution.
Our group also developed a series of IL-alkanolamine mixtures to capture CO2 and
obtained some good results. A series of novel ILs, 2-aminoethanol tetrafluoroborate
([MEA][BF4]), 2-[2-hydroxyethyl(methyl)-amino] ethanol tetrafluoroborate
([MDEA][BF4]), 2-[2-hydroxyethyl(methyl)amino] ethanol chloride ([MDEA][Cl]),
2-[2-hydroxyethyl(methyl)amino] ethanol phosphate ([MDEA][PO4]), and 2-
[2-hydroxyethyl(methyl)amino] ethanol sulfate ([MDEA][SO4]), were synthesized
and blended with amines and H2O to capture CO2. Among these absorbents, the
MDEA + [MDEA][Cl] + H2O + piperazine system shows the best performance on
CO2 capture [64]. Two low viscous ILs, 1-(2-hydroxyethyl)-3-methyl-imidazolium
dicyanamide ([C2OHmim][DCA]) and [Bmim][DCA] were also mixed with aqueous
30 wt% MEA for CO2 absorption. The results indicated that the CO2 loading
decreased significantly as the ILs concentration increased [56]. Another three kinds of
ILs [Bmim][BF4], [Bmim][NO3] and [Bmim][Cl] were selected to blend with
MDEA + PZ aqueous solution to capture CO2. The results showed that the influence
of ILs on the CO2 cyclic capacity following the order of MDEA + PZ + [Bmim]
[BF4] > MDEA + PZ + [Bmim][Cl]  MDEA + PZ > MDEA + PZ + [Bmim]
[NO3], indicating that an addition of [Bmim][BF4] decreased the sensible heat [70].
Because of certain amount of water replacing by ILs, the energy consumption of
IL-alkanolamine mixtures would be lower than that of aqueous alkanolamine solu-
tion. For the functionalized ILs, the solubility of CO2 in IL-alkanolamine mixtures
increases as the ILs concentration increased. Therefore, IL-alkanolamine mixtures
may be a potentail solvent to capture CO2 in an industrial scale.

4.3 IL-Organic Solvent Systems

The viscosity of ILs could decrease significantly by adding a small amount of


organic solvent besides water. Therefore, the mixture of ILs and organic solvents
168 X. Zhang et al.

are investigated as other candidates for CO2 capture. Hong et al. [71] measured the
solubility of CO2 in the mixture of 1-ethyl-3- methylimidazolium bis
(trifluoromethanesulfonyl)amide ([Emim][NTf2]) + acetonitrile (CH3CN) at
0.1 MPa and temperature between 290 and 335 K. The results indicated that the
solubility of CO2 decrease about 50 % with the mole fraction of acetonitrile
increasing from 0 to 0.77. Li et al. [57] investigated the absorption of CO2 using the
mixture of (2-hydroxyethyl)-trimethyl-ammonium (S)-2-pyrrolidine-carboxylic
acid salt ([Choline][Pro]) and PEG 200 at the temperature from 308.15 to
353.15 K under the ambient pressure. The results showed that the molar ratio of
CO2 to the IL could exceed 0.5 slightly, which is the theoretical maximum for the
absorption of CO2 chemically, indicating that both chemical and physical absorp-
tion exists. Addition of PEG200 in the IL could improve the rates of absorption and
desorption of CO2. Ahnn et al. [72] measured the solubility of CO2 in the mixtures
of dimethyl carbonate (DMC) + [Hmim][NTf2] at several temperatures between
303.15 K and 333.15 K and at pressures up to about 7 MPa, showing that the CO2
solubility decreased as the IL content increased at the same pressure and temper-
ature. Wang et al. [73] investigated the absorption performance of CO2
by 1-butyl-3-methylimidazolium bis(trifluoromethanesulfonyl)amide ([Bmim]
[NTf2]) + 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) and [Bmim][NTf2] +
1,3,4,6,7,8-hexahydro-1-methyl-2H-pyrimido[1,2-a]pyrimidine (MTBD), and the
results suggested that the CO2 absorption capacity is about 1 mol per mole of IL,
which is superior to traditional ILs, indicating that both chemical and physical
absorption occurred simultaneously. Lei et al. [59, 74] determined the solubility of
CO2 in the mixtures of propanone + 1-ethyl-3-methylimidazolium tetrafluoroborate
([Emim][BF4]) or acetone + [Bmim][BF4] and found that the mixtures could be
applied as promising solvents for capturing CO2 since they combine the advantages
of organic solvents and ILs. Lei et al. [75] also measured the solubility of CO2 in
the mixtures of methanol + 1-octyl-3-methylimidazolium bis(trifluoromethanesul-
fonyl)amide ([Omim][NTf2]) at 273.2, 258.2, 243.2, 228.2 K and pressures up to
3.0 MPa. The process simulation reveals that the volatile loss of methanol
decreases in some degrees by adding [Omim][NTf2] into methanol.
The mixtures of IL and organic solvent show potentials for CO2 capture, but the
stability and loss of the solvent should be considered when being applied for real
industries.

4.4 IL-IL Systems

The mixtures of ILs themselves are also reported to capture CO2, which could
provide an additional degree of tailoring over the intrinsic tunable properties of
“single” ILs, while totally maintaining the IL property. Finotello et al. [49] found
that the selectivity for CO2 with N2 and CH4 in the 90 and 95 mol% mixtures of
[Emim][BF4] in [Emim][NTf2] was higher than in both pure components. Shiflett
et al. [58] measured the solubility of CO2 in the IL mixture containing
Ionic Liquids: Advanced Solvents for CO2 Capture 169

equimolar amounts of 1-ethyl-3-methylimidazolium acetate ([Emim][Ac]),


1-ethyl-3-methylimidazolium trifluoroacetate ([Emim][TFA]) at three temperatures
(298.1, 323.1, and 348.1 K) and pressures up to about 2 MPa. The results indicated
that the mixture could absorb CO2 by chemical and physical interaction. Lei et al.
[76] determined the solubility of CO2 in binary mixtures of [Emim]
[BF4] + [Omim][NTf2] and [Bmim][BF4] + [Omim][NTf2] at high pressure up to
6.0 MPa for physical absorption. The Henry’s law constants decreased with
increasing [Omim][NTf2] content in [Emim][BF4] or [Bmim][BF4], and were
consistent with the COSMO-RS calculation. Pinto et al. [4] investigated two binary
mixtures of [Emim][NTf2] + 1-ethyl-3-methylimidazolium ethylsulfate ([Emim]
[EtSO4]) and [Emim][NTf2] + 1-butyl-3-methylimidazolium ethylsulfate ([Bmim]
[EtSO4]) to absorb CO2 at 298 K and pressures up to 1.6 MPa. Although the
absorption capacity of the mixtures was not higher than that of [Emim][NTf2], the
cost and toxicity of the IL mixtures could be tuned by composition. They also
measured the CO2 solubility in the mixture of [Emim][Ac] + [Emim][EtSO4] at
298.2 and 353.2 K, and at pressures up to 1.7 MPa. The results suggested that the
addition of [Emim][EtSO4] to [Emim][Ac] prevent the solidification of the product
resulting from the chemical reaction between CO2 and [Emim][Ac] [60]. Wang
et al. [77] studied the CO2 absorption performance using the binary ILs of
1-butyl-3-propylamineimidazolium tetrafluoroborate ([NH2p-bim][BF4]) + [Emim]
[BF4] + [Bmim][BF4]. The results showed that when the mole fraction of
[NH2e-mim][BF4] was 0.4, the CO2 absorption performance and cost of the mix-
tures were the best. The CO2 absorption capacity and CO2 absorption rate decreased
with the increase of the absorption temperature.
IL-IL mixtures, in particular of those that combine physical and chemical
absorption of CO2, may be better than simple IL for the development of CO2
capture processes. This method could balance the absorption capacity, the
thermo-physical properties, cost and toxicity of the ILs.

5 Simulation and Assessment of Ionic Liquids Process


for CO2 Separation

System integration and assessment is very indispensable for developing new pro-
cesses. As a new kind of solvents, a systematic assessment of CO2 capture process
with ILs is rarely studied, so a comprehensive simulation and assessment is very
necessary for CO2 separation with ILs in order to quickly compare different
strategies, analyse process sensitivity, estimate the whole operating and equipment
cost, understanding the energy consumption and the conversion efficiency, and then
finally find the optimum route.
Many researchers have performed experiments for CO2 capture with various ILs,
as ILs possess good properties of non-volatile, good stability and relative low
viscosity. However, since the shortage of rigorous thermodynamic data for com-
plicated system, IL-based process simulation was rarely reported [78]. Shiflett et al.
170 X. Zhang et al.

[6] simulated a CO2 capture process using pure [Bmim][Ac]. The whole process
consisted of two parts: absorption and solvent regeneration. Compared with the
commercial MEA-based process, their IL-based process separated about 90 % of
CO2 from the flue gas slip stream and the annual capacity was 47,000 metric tons,
the IL-based process showed a higher recovery of 91.3 % and achieved higher CO2
purity of 98.7 %. Moreover, the IL-based system reduced 16 % energy consump-
tion and the assessment result showed that the investment for the IL process was
11 % lower than the MEA-based process and also 12 % lower in equipment
footprint, respectively [6]. Basha et al. [79] developed a process using three flash
drums as regenerators to recycle the [hmim][Tf2N] for CO2 capture. Through three
different pressure adiabatic flash drums (the pressure is 20, 10, 1 bar respectively),
the CO2 gas, containing some H2 and H2O vapor, can be separated from the IL.
Then the IL can be recycled into the absorbers. The results show that at the absorber
conditions (2.4 m inner diameter, 30 m high), the [hmim][Tf2N] IL could achieve a
CO2 recovery about 97.27 %. Eisinger et al. [80] employed an IL-[P2222][BnIm] to
capture CO2 from the flue gas, which showed a phase change from solid to liquid
upon reaction with CO2. Then the process can make a good use of the heat gen-
erated by this phase change to reduce parasitic power consumption, which is 55 %
lower than that of the MEA process. And assessment indicated that the cost of
electricity is 28 % lower but the capital cost is higher than the conventional MEA
process. Not only the pure ionic liquid, researchers also found that the IL-amine
hybrid solvents showed more effective aspect for CO2 capture, which can combine
both the advantages of IL and the traditional solvents. Huang et al. [7] studied three
ILs ([Bmim][DCA], [Bmim][BF4] and [Bpy][BF4]) with MEA aqueous solution
and compared with the process simulation method. The results showed that com-
pared with the conventional MEA process, the [Bpy][BF4]-MEA process can
reduce 15 % in regeneration heat duty, 12 % in the whole energy penalty and
13.5 % in capture cost. As a consequence, this combined solvent process can be
regarded as energy-saving and cost-efficient carbon capture process.
These previous work opens a door of using IL solvents for CO2 separation,
which can achieve the target of high CO2 recovery and low energy consumption
compared to traditional processes. As is known, it is an important approach to
simulate and assess the separation process using IL-based solvents. However, since
the simulation work is different from the actual condition which much loss of the
energy and substances can cause. Then many assumptions should be established
and different conditions should be considered.

6 Vision

Due to the unique properties of ILs, they have been proposed as one of promising
CO2 technologies from the economic and environmental viewpoints. In this
chapter, the research progress on the CO2 capture of conventional ILs, task-specific
ILs and ILs based mixtures have been reviewed and discussed. The CO2 solubility
Ionic Liquids: Advanced Solvents for CO2 Capture 171

of conventional ILs is lower because of the weak physical interaction between ILs
and CO2, while the solubility of task-specific ILs, especially with amine groups is
much higher due to the chemical reaction with CO2. The addition of other sub-
stances to form ILs mixed solvents could significantly decrease the viscosity of ILs
while the desired properties of ILs for CO2 are remained. It was described and
explained through the experimental and simulation method that the cations, anions
and functional groups have influences on the CO2 absorption. The process simu-
lation of ILs/ILs mixtures for CO2 capture presented that IL system is an
energy-saving and cost efficient way for separating CO2.
Although CO2 capture with ILs/ILs based mixtures has been well developed,
there are still some barriers which need to be solved in future works for the
industrialization of ILs technology. Firstly, the high viscosity and moderate
capacity of ILs is a great obstacle for industrial applications. So designing new
ILs/ILs based mixtures to improve the CO2 gravimetric capacity and selectivity as
well as decreasing the viscosity is extremely important. To this end, it is critical to
deeply study and understand the relationship of structures, properties and absorp-
tion mechanism of ILs, and then directing the design and synthesis of ILs with
certain characters for efficient CO2 separation. Secondly, the high price of ILs
themselves has significantly affected the investment of the new ILs synthesis pro-
cesses. Most ILs are prepared in laboratory scale with complex synthesis and
purification steps, which leads to the high cost of ILs. Thus the improvement of
synthetic method and the production on a large scale to reduce the price of ILs at a
certain extent should be focused on. Thirdly, the impacts on the environment of IL
process are not clear due to the unknown safety and health of ILs, so the toxicity
and degradation of ILs should be studied thoroughly before commercializing the IL
process. Except for these, the effect of other gas components on CO2 capture should
be systematically studied in future work, because the practical separation processes
involving gas mixtures, including CO2 and other gases, like oxygen, water, SO2,
and so on, which has influence on the CO2 absorption with ILs. In addition, a
systematic assessment of the new processes based on a thermodynamic model for
ILs is very scarce. Thus, rigorous thermodynamic of IL complex system and rea-
sonable simulation assumptions should be established, and also different conditions
should be considered in future works, in order to assess the feasibility and rea-
sonability of the IL process before accepted by industrial. Finally, before the lab
research results are transferred to large scaled processes, building a pilot plant test is
helpful to find and clearly understand the problems existed in ILs technology and
hence realize successful industrialization.
Currently, ILs have also been combined with other technologies like absorption
and membrane gas separation to develop new technologies for CO2 capture, in
which ILs based membrane for gas separation is regarded as future research trends
of ILs. Making ILs immobiled to form ILs based membranes could significantly
improve the permeability and selectivity, reduce the energy consumption and the
membrane area, meanwhile, it could overcome the high viscosity and cost of ILs.
Some works about combining ILs with membrane gas separation have been
developed, including supported ionic liquids membranes, poly(ionic liquid)
172 X. Zhang et al.

membranes, ionic liquids based composite membranes, etc. Although ILs based
membranes connecting the advantages of ILs and membranes, it is still a big
challenge to develop novel ILs based membranes with higher permeability and
selectivity, thermal and mechanical stability and thinner membranes in order to
meet the requirement of industrial applications. Moreover, the separation mecha-
nism of membranes is not cleared to date, so the influence of membrane structures,
interaction between ILs and polymers on the membrane performance should be
investigated deeply in order to explain the gas separation mechanism, hence
guiding the designing of membranes and future industrial applications.

Acknowledgments This work was supported by the National Natural Science Fund for
Distinguished Young Scholars (No. 21425625), the National Natural Science Foundation of China
(No. 51574215, 21506219), the International S&T Cooperation Program of China
(No. 2014DFA61670), and the External Cooperation Program of BIC, Chinese Academy of
Sciences (No. l22111KYS820150017).

References

1. Service RF (2004) Choosing a CO2 separation technology. Science 305(5686):963


2. Kenarsari SD, Yang D, Jiang G, Zhang S, Wang J, Russell AG, Wei Q, Fan M (2013) Review
of recent advances in carbon dioxide separation and capture. Rsc Adv 3(45):22739–22773
3. Huang K, Zhang X-M, Xu Y, Wu Y-T, Hu X-B, Xu Y (2014) Protic ionic liquids for the
selective absorption of H2S from CO2: thermodynamic analysis. AIChE J 60(12):4232–4240
4. Pinto AM, Rodriguez H, Colon YJ, Arce A Jr, Arce A, Soto A (2013) Absorption of carbon
dioxide in two binary mixtures of ionic liquids. Ind Eng Chem Res 52(17):5975–5984
5. Sharma P, Park SD, Baek IH, Park KT, Yoon YI, Jeong SK (2012) Effects of anions on
absorption capacity of carbon dioxide in acid functionalized ionic liquids. Fuel Process
Technol 100:55–62
6. Shiflett MB, Drew DW, Cantini RA, Yokozeki A (2010) Carbon dioxide capture using ionic
liquid 1-butyl-3-methylimidazolium acetate. Energy Fuels 24:5781–5789
7. Huang Y, Zhang X, Zhang X, Dong H, Zhang S (2014) Thermodynamic modeling and
assessment of ionic liquid-based CO2 capture processes. Ind Eng Chem Res 53(29):11805–
11817
8. Blanchard LA, Hancu D, Beckman EJ, Brennecke JF (1999) Green processing using ionic
liquids and CO2. Nature 399(6731):28–29
9. Zhang X, Zhang X, Dong H, Zhao Z, Zhang S, Huang Y (2012) Carbon capture with ionic
liquids: overview and progress. Energy Environ Sci 5(5):6668–6681
10. Zhao Y, Zhang X, Zhen Y, Dong H, Zhao G, Zeng S, Tian X, Zhang S (2011) Novel
alcamines ionic liquids based solvents: Preparation, characterization and applications in
carbon dioxide capture. Int J Greenhouse Gas Control 5(2):367–373
11. Wappel D, Gronald G, Kalb R, Draxler J (2010) Ionic liquids for post-combustion CO2
absorption. Int J Greenhouse Gas Control 4(3):486–494
12. Vega LF, Vilaseca O, Llovell F, Andreu JS (2010) Modeling ionic liquids and the solubility
of gases in them: recent advances and perspectives. Fluid Phase Equilib 294(1–2):15–30
13. Kazarian SG, Briscoe BJ, Welton T (2000) Combining ionic liquids and supercritical fluids:
in situ ATR-IR study of CO2 dissolved in two ionic liquids at high pressures. Chem Commun
20:2047–2048
14. Cammarata L, Kazarian SG, Salter PA, Welton T (2001) Molecular states of water in room
temperature ionic liquids. Physical Chemistry Chemical Physics 3(23):5192–5200
Ionic Liquids: Advanced Solvents for CO2 Capture 173

15. Crowhurst L, Mawdsley PR, Perez-Arlandis JM, Salter PA, Welton T (2003) Solvent-solute
interactions in ionic liquids. Phys Chem Chem Phys 5(13):2790–2794
16. Dong K, Zhang S, Wang D, Yao X (2006) Hydrogen bonds in imidazolium ionic liquids.
J Phys Chem A 110(31):9775–9782
17. Yu G, Zhang S (2007) Insight into the cation-anion interaction in 1,1,3,3-tetramethylguanidinium
lactate ionic liquid. Fluid Phase Equilib 255(1):86–92
18. Zhang J, Zhang S, Dong K, Zhang Y, Shen Y, Lv X (2006) Supported absorption of CO2 by
tetrabutylphosphonium amino acid ionic liquids. Chem Eur J 12(15):4021–4026
19. Zhang Y, Zhang S, Lu X, Zhou Q, Fan W, Zhang X (2009) Dual amino-functionalised
phosphonium ionic liquids for CO2 capture. Chem Eur J 15(12):3003–3011
20. Xue Z, Zhang Z, Han J, Chen Y, Mu T (2011) Carbon dioxide capture by a dual amino ionic
liquid with amino-functionalized imidazolium cation and taurine anion. Int J Greenhouse Gas
Control 5(4):628–633
21. Liu X, Zhou G, Zhang S, Yao X (2009) Molecular dynamics simulation of dual
amino-functionalized imidazolium-based ionic liquids. Fluid Phase Equilib 284(1):44–49
22. Wang C, Luo H, D-e Jiang, Li H, Dai S (2010) Carbon dioxide capture by superbase-derived
protic ionic liquids. Angewandte Chemie-International Edition 49(34):5978–5981
23. Zhang J, Jia C, Dong H, Wang J, Zhang X, Zhang S (2013) A novel dual
amino-functionalized cation-tethered ionic liquid for CO2 capture. Ind Eng Chem Res 52
(17):5835–5841
24. Gurkan B, Goodrich BF, Mindrup EM, Ficke LE, Massel M, Seo S, Senftle TP, Wu H,
Glaser MF, Shah JK, Maginn EJ, Brennecke JF, Schneider WF (2010) Molecular design of
high capacity, low viscosity, chemically tunable ionic liquids for CO2 capture. J Phys Chem
Lett 1(24):3494–3499
25. Carlisle TK, Bara JE, Gabriel CJ, Noble RD, Gin DL (2008) Interpretation of CO2 solubility
and selectivity in nitrile-functionalized room-temperature ionic liquids using a group
contribution approach. Ind Eng Chem Res 47(18):7005–7012
26. Blanchard LA, Gu ZY, Brennecke JF (2001) High-pressure phase behavior of ionic
liquid/CO2 systems. J Phys Chem B 105(12):2437–2444
27. Cadena C, Anthony JL, Shah JK, Morrow TI, Brennecke JF, Maginn EJ (2004) Why is CO2
so soluble in imidazolium-based ionic liquids? J Am Chem Soc 126(16):5300–5308
28. Huang XH, Margulis CJ, Li YH, Berne BJ (2005) Why is the partial molar volume of CO2 so
small when dissolved in a room temperature ionic liquid? Structure and dynamics of CO2
dissolved in bmim+pf6−. J Am Chem Soc 127(50):17842–17851
29. Almantariotis D, Gefflaut T, Padua AAH, Coxam JY, Gomes MFC (2010) Effect of
fluorination and size of the alkyl side-chain on the solubility of carbon dioxide in
1-alkyl-3-methylimidazolium bis(trifluoromethylsulfonyl)amide ionic liquids. J Phys Chem B
114(10):3608–3617
30. Aki S, Mellein BR, Saurer EM, Brennecke JF (2004) High-pressure phase behavior of carbon
dioxide with imidazolium-based ionic liquids. J Phys Chem B 108(52):20355–20365
31. Yunus NM, Mutalib MIA, Man Z, Bustam MA, Murugesan T (2012) Solubility of CO2 in
pyridinium based ionic liquids. Chem Eng J 189:94–100
32. Bara JE, Gabriel CJ, Lessmann S, Carlisle TK, Finotello A, Gin DL, Noble RD (2007)
Enhanced CO2 separation selectivity in oligo(ethylene glycol) functionalized
room-temperature ionic liquids. Ind Eng Chem Res 46(16):5380–5386
33. Makino T, Kanakubo M, Umecky T (2014) CO2 solubilities in ammonium bis
(trifluoromethanesulfonyl)amide ionic liquids: Effects of ester and ether groups. J Chem
Eng Data 59(5):1435–1440
34. Anthony JL, Anderson JL, Maginn EJ, Brennecke JF (2005) Anion effects on gas solubility in
ionic liquids. J Phys Chem B 109(13):6366–6374
35. Muldoon MJ, Aki SNVK, Anderson JL, Dixon JK, Brennecke JF (2007) Improving carbon
dioxide solubility in ionic liquids. J Phys Chem B 111(30):9001–9009
36. Zhang X, Liu Z, Wang W (2008) Screening of ionic liquids to capture CO2 by COSMO-RS
and experiments. AIChE J 54(10):2717–2728
174 X. Zhang et al.

37. Bhargava BL, Balasubramanian S (2007) Probing anion-carbon dioxide interactions in room
temperature ionic liquids: gas phase cluster calculations. Chem Phys Lett 444(4–6):242–246
38. Seki T, Grunwaldt J-D, Baiker A (2009) In situ attenuated total reflection infrared
spectroscopy of imidazolium-based room-temperature ionic liquids under “supercritical” CO2.
J Phys Chem B 113(1):114–122
39. Bates ED, Mayton RD, Ntai I, Davis JH (2002) CO2 capture by a task-specific ionic liquid.
J Am Chem Soc 124(6):926–927
40. Gurkan BE, de la Fuente JC, Mindrup EM, Ficke LE, Goodrich BF, Price EA, Schneider WF,
Brennecke JF (2010) Equimolar CO2 absorption by anion-functionalized ionic liquids. J Am
Chem Soc 132(7):2116–2117
41. Zhang SJ, Yuan XL, Chen YH, Zhang XP (2005) Solubilities of CO2 in
1-butyl-3-methylimidazolium hexafluorophosphate and 1,1,3,3-tetramethylguanidium lactate
at elevated pressures. J Chem Eng Data 50(5):1582–1585
42. Yu G, Zhang S, Zhou G, Liu X, Chen X (2007) Structure, interaction and property of
amino-functionalized imidazolium ILs by molecular dynamics simulation and ab initio
calculation. AIChE J 53(12):3210–3221
43. Zhang SJ, Chen YH, Li FW, Lu XM, Dai WB, Mori R (2006) Fixation and conversion of
CO2 using ionic liquids. Catal Today 115(1–4):61–69
44. Wang C, Mahurin SM, Luo H, Baker GA, Li H, Dai S (2010) Reversible and robust
CO2 capture by equimolar task-specific ionic liquid-superbase mixtures. Green Chem 12
(5):870–874
45. Wang C, Luo X, Luo H, D-e Jiang, Li H, Dai S (2011) Tuning the basicity of ionic liquids for
equimolar CO2 capture. Angewandte Chemie-Intl Ed 50(21):4918–4922
46. Wang C, Luo H, Li H, Zhu X, Yu B, Dai S (2012) Tuning the physicochemical properties of
diverse phenolic ionic liquids for equimolar CO2 capture by the substituent on the anion.
Chem Eur J 18(7):2153–2160
47. Luo X, Guo Y, Ding F, Zhao H, Cui G, Li H, Wang C (2014) Significant improvements in
CO2 capture by pyridine-containing anion-functionalized ionic liquids through multiple-site
cooperative interactions. Angewandte Chemie-Intl Ed 53(27):7053–7057
48. Seddon KR, Stark A, Torres MJ (2000) Influence of chloride, water, and organic solvents on
the physical properties of ionic liquids. Pure Appl Chem 72(12):2275–2287
49. Bermejo MD, Montero M, Saez E, Florusse LJ, Kotlewska AJ, Cocero MJ, van Rantwijk F,
Peters CJ (2008) Liquid-vapor equilibrium of the systems butylmethylimidazolium
nitrate-CO2 and hydroxypropylmethylimidazolium nitrate-CO2 at high pressure: Influence
of water on the phase behavior. J Phys Chem B 112(43):13532–13541
50. Ventura SPM, Pauly J, Daridon JL, da Silva JAL, Marrucho IM, Dias AMA, Coutinho JAP
(2008) High pressure solubility data of carbon dioxide in (tri-iso-butyl(methyl)phosphonium
tosylate plus water) systems. J Chem Thermodyn 40(8):1187–1192
51. Wang GN, Hou WL, Xiao F, Geng JA, Wu YT, Zhang ZB (2011) Low-viscosity
triethylbutylammonium acetate as a task-specific ionic liquid for reversible CO2 absorption.
J Chem Eng Data 56(4):1125–1133
52. Seo S, Quiroz-Guzman M, DeSilva MA, Lee TB, Huang Y, Goodrich BF, Schneider WF,
Brennecke JF (2014) Chemically tunable ionic liquids with aprotic heterocyclic anion
(AHA) for CO2 capture. J Phys Chem B 118(21):5740–5751
53. Camper D, Bara JE, Gin DL, Noble RD (2008) Room-temperature ionic liquid-amine
solutions: tunable solvents for efficient and reversible capture of CO2. Ind Eng Chem Res 47
(21):8496–8498
54. Zhang F, Fang CG, Wu YT, Wang YT, Li AM, Zhang ZB (2010) Absorption of CO2 in the
aqueous solutions of functionalized ionic liquids and mdea. Chem Eng J 160(2):691–697
55. Huang Q, Li Y, Jin XB, Zhao D, Chen GZ (2011) Chloride ion enhanced thermal stability of
carbon dioxide captured by monoethanolamine in hydroxyl imidazolium based ionic liquids.
Energy Environ Sci 4(6):2125–2133
Ionic Liquids: Advanced Solvents for CO2 Capture 175

56. Xu F, Gao HS, Dong HF, Wang ZL, Zhang XP, Ren BZ, Zhang SJ (2014) Solubility of CO2
in aqueous mixtures of monoethanolamine and dicyanamide-based ionic liquids. Fluid Phase
Equilib 365:80–87
57. Li XY, Hou MQ, Zhang ZF, Han BX, Yang GY, Wang XL, Zou LZ (2008) Absorption of
CO2 by ionic liquid/polyethylene glycol mixture and the thermodynamic parameters. Green
Chem 10(8):879–884
58. Shiflett MB, Yokozeki A (2009) Phase behavior of carbon dioxide in ionic liquids: [emim]
[acetate], [emim][trifluoroacetate], and [emim][acetate] plus [emim][trifluoroacetate] mix-
tures. J Chem Eng Data 54(1):108–114
59. Lei ZG, Han JL, Zhang BF, Li QS, Zhu JQ, Chen BH (2012) Solubility of CO2 in
binary mixtures of room-temperature ionic liquids at high pressures. J Chem Eng Data 57
(8):2153–2159
60. Pinto AM, Rodriguez H, Arce A, Soto A (2014) Combined physical and chemical absorption
of carbon dioxide in a mixture of ionic liquids. J Chem Thermodyn 77:197–205
61. Goodrich BF, de la Fuente JC, Gurkan BE, Lopez ZK, Price EA, Huang Y, Brennecke JF
(2011) Effect of water and temperature on absorption of CO2 by amine-functionalized
anion-tethered ionic liquids. J Phys Chem B 115(29):9140–9150
62. Romanos GE, Zubeir LF, Likodimos V, Falaras P, Kroon MC, Iiev B, Adamova G,
Schubert TJS (2013) Enhanced CO2 capture in binary mixtures of
1-alkyl-3-methylimidazolium tricyanomethanide ionic liquids with water. J Phys Chem B
117(40):12234–12251
63. Anderson K, Atkins MP, Estager J, Kuah Y, Ng S, Oliferenko AA, Plechkova NV, Puga AV,
Seddon KR, Wassell DF (2015) Carbon dioxide uptake from natural gas by binary ionic
liquid-water mixtures. Green Chem 17(8):4340–4354
64. Ahmady A, Hashim MA, Aroua MK (2010) Experimental investigation on the solubility and
initial rate of absorption of CO2 in aqueous mixtures of methyldiethanolamine with the ionic
liquid 1-butyl-3-methylimidazolium tetrafluoroborate. J Chem Eng Data 55(12):5733–5738
65. Ahmady A, Hashim MA, Aroua MK (2011) Absorption of carbon dioxide in the aqueous
mixtures of methyldiethanolamine with three types of imidazolium-based ionic liquids. Fluid
Phase Equilib 309(1):76–82
66. Gao Y, Zhang F, Huang K, Ma JW, Wu YT, Zhang ZB (2013) Absorption of CO2 in amino
acid ionic liquid (AAIL) activated MDEA solutions. Int J Greenhouse Gas Control 19:379–
386
67. Zhang F, Gao Y, Wu XK, Ma JW, Wu YT, Zhang ZB (2013) Regeneration performance of
amino acid ionic liquid (AAIL) activated mdea solutions for CO2 capture. Chem Eng J
223:371–378
68. Sairi NA, Yusoff R, Alias Y, Aroua MK (2011) Solubilities of CO2 in aqueous
N-methyldiethanolamine and guanidinium trifluoromethanesulfonate ionic liquid systems at
elevated pressures. Fluid Phase Equilib 300(1–2):89–94
69. Yang J, Yu XH, Yan JY, Tu ST (2014) CO2 capture using amine solution mixed with ionic
liquid. Ind Eng Chem Res 53(7):2790–2799
70. Gao JB, Cao LD, Dong HF, Zhang XP, Zhang SJ (2015) Ionic liquids tailored amine aqueous
solution for pre-combustion CO2 capture: role of imidazolium-based ionic liquids. Appl
Energy 154:771–780
71. Hong G, Jacquemin J, Husson P, Gomes MFC, Deetlefs M, Nieuwenhuyzen M, Sheppard O,
Hardacre C (2006) Effect of acetonitrile on the solubility of carbon dioxide in
1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl) amide. Ind Eng Chem Res 45
(24):8180–8188
72. Ahn JY, Lee BC, Lim JS, Yoo KP, Kang JW (2010) High-pressure phase behavior of binary
and ternary mixtures containing ionic liquid [C6mim][Tf2N], dimethyl carbonate and carbon
dioxide. Fluid Phase Equilib 290(1–2):75–79
73. Wang CM, Luo HM, Luo XY, Li HR, Dai S (2010) Equimolar CO2 capture by
imidazolium-based ionic liquids and superbase systems. Green Chem 12(11):2019–2023
176 X. Zhang et al.

74. Zhao YS, Zhang XP, Zeng SJ, Zhou Q, Dong HF, Tian XA, Zhang SJ (2010) Density,
viscosity, and performances of carbon dioxide capture in 16 absorbents of amine plus
ionic liquid + H2O, ionic liquid + H2O, and amine + H2O systems. J Chem Eng Data 55
(9):3513–3519
75. Dai CN, Wei WJ, Lei ZG, Li CX, Chen BH (2015) Absorption of CO2 with methanol and
ionic liquid mixture at low temperatures. Fluid Phase Equilib 391:9–17
76. Lei ZG, Qi XX, Zhu JQ, Li QS, Chen BH (2012) Solubility of CO2 in acetone,
1-butyl-3-methylimidazolium tetrafluoroborate, and their mixtures. J Chem Eng Data 57
(12):3458–3466
77. Wang M, Zhang LQ, Gao LX, Pi KW, Zhang JY, Zheng CG (2013) Improvement of the CO2
absorption performance using ionic liquid [NH2emim][BF4] and [emim][BF4]/[bmim][BF4]
mixtures. Energy Fuels 27(1):461–466
78. Kumar S, Cho JH, Moon I (2014) Ionic liquid-amine blends and CO2BOLs: Prospective
solvents for natural gas sweetening and CO2 capture technology—a review. Int J Greenhouse
Gas Control 20:87–116
79. Basha OM, Keller MJ, Luebke DR, Resnik KP, Morsi BI (2013) Development of a
conceptual process for selective CO2 capture from fuel gas streams using [hmim][Tf2N] ionic
liquid as a physical solvent. Energy Fuels 27(7):3905–3917
80. Eisinger RS, Keller GE II (2014) Process for CO2 capture using ionic liquid that exhibits
phase change. Energy Fuels 28(11):7070–7078
Amine-Blends Screening
and Characterization for CO2
Post-combustion Capture

Abdullah Al Hinai, Nabil El Hadri and Mohammad Abu Zahra

Abstract Amines have been identified as one of the most promising agents to
capture CO2 released from industrial and energy sources to the atmosphere. The
primary aim of this chapter is to present experimental results showing the two
major factors considered in the selection of suitable amine solvents for carbon
capture in post-combustion technology. Ten amines were selected from varying
classes and structure types, and were investigated on the basis of CO2 loading and
heat of absorption. These solvents were selected based on the preliminary experi-
mental results and the effect of blending the best reacting solvents was further
investigated. Details of the experimental procedure and results are discussed as well
as the relevant conclusions.

1 Introduction

CCS (carbon capture and storage) post-combustion technology can be categorized


as one of the most suitable technology process to decrease the CO2 released to the
atmosphere because of the ability to install this technology directly into existing
power plants [1] and this technology can be applied for CO2 separation from highly
diluted stream of flue gas using the chemical absorption technique.
The three basic classes of alkanolamines: Monoethanolamine (MEA)—primary
amine, diethanolamine (DEA)—secondary amine, and methyl diethanolamine
(MDEA)—tertiary amine, are the most widely used [2]. The primary and secondary
alkanolamines have a high rate reaction with CO2 when first there is a formation of
a zwitterion (reaction 1), which then transfers a proton to an amine resulting in the
formation of a carbamate ion (reaction 2). Their maximum CO2 loading is 0.5 mol
of CO2/mole of amine (two molecules of amines is needed to react with one
molecule of CO2). At a high CO2 pressure, the carbamate could be hydrolyzed to

A. Al Hinai  N. El Hadri  M. Abu Zahra (&)


Department of Chemical and Environmental Engineering, Masdar Institute
of Science and Technology, P.O. Box 54224, Masdar City, United Arab Emirates
e-mail: mabuzahra@masdar.ac.ae

© Springer International Publishing AG 2017 177


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_8
178 A. Al Hinai et al.

form a free amine and bicarbonates and then the free amine will again react with
CO2 (reaction 3).

R1 R2 NH þ CO2 ðaqÞ $ R1 R2 NH þ COO ð1Þ

R1 R2 NH þ R1 R2 NH þ COO $ R1 R2 NH þ þ R1 R2 NCOO ð2Þ

R1 R2 NCOO þ H2 O $ R1 R2 NH þ HCO
3 ð3Þ

The maximum CO2 loading which can be obtained from tertiary or hindered
amines is 1 mol of CO2 per mole of amine (one molecule of amine will react with
one molecule of CO2) with the formation of bicarbonates (reaction 4). However,
despite the higher achievable CO2 loading, tertiary amines have low reactivity with
CO2 in comparison with the primary or secondary amines.

R1 R2 R3 N þ CO2 ðaqÞ þ H2 O $ R1 R2 R3 NH þ þ HCO


3 ð4Þ

A typical conventional CO2 post-combustion capture process based on aqueous


MEA solution (30 wt%) has been used as a reference case in many studies. MEA
has been chosen because of its high reaction rate with CO2 and reasonable CO2
absorption capacity and low cost [3]. However, the regeneration energy to regen-
erate the MEA solution after CO2 absorption is high and this hinders industrial
application of the conventional post-combustion process [4, 5]. Recently, great
efforts has been made to find alternative amines with high CO2 absorption capacity,
high absorption rate and low heat of absorption in comparison with MEA. The main
challenge is to find new amine solvents performance in order to reduce the cost of
CO2 capture process.
Several studies have been carried out in order to characterize the CO2-reaction
enthalpy of aqueous amines solution [6–8]. Understanding this property is essential
in order to identify the quantity of heat which is needed in the regeneration column
on the process. The heat of absorption is associated with the quantity of steam
necessary to regenerate the aqueous amine solution. The heat of absorption is very
high for primary and secondary amines and increase the regeneration energy of the
solvent. In opposite, the heat of absorption of the tertiary amines is low and the
energy required to regenerate the solvent is reduced [9]. The determination of the
heat of absorption could be done using the Gibbs-Helmholtz equation from the CO2
solubility data or by direct calorimetric measurements. The Gibbs-Helmholtz
equation was used by Kim et al. for the calculation of the heat of absorption of
MEA and MDEA from the equilibrium constant of each reaction occurred between
the reactions of CO2 with the respective aqueous amine solution [10]. The direct
calorimetric measurements have the ability to give accurate data for the heat of
absorption because it incorporates the heat due to physical dissolution of CO2 in the
aqueous amine solution and the chemical reaction between CO2 and aqueous amine
solution. Different equipment are published in various works. Mathonat et al.
Amine-Blends Screening and Characterization for CO2 … 179

measured the CO2-reaction enthalpy for aqueous solutions of MEA and MDEA
30 wt% at 40, 80 and 120 °C using a flow calorimeter [11].
The objective of this work is to investigate the improvement of thermodynamic
properties such as absorption capacity and heat of absorption by blending amines.

2 Research Methodology and Experiments

2.1 CO2 Loading in Amines

Solubility of CO2 by aqueous amine solutions have been looked into in several
studies in order to find promising molecules with a high CO2 loading [12–17].
Bonenfant et al. studied the effect of different amine structures (amine, diamines
and polyamines) on the absorption and desorption of CO2 at a concentration of
amine of 5 wt%. The work was based on eight amines and the results show that
N-(2-aminoethyl)-1,3-propanediamine (AEPDNH2, polyamines with two primary
and one secondary amine), 2-(2-aminoethylamino)ethanol (AEE, diamine with one
primary and one secondary amine) have the promising characteristics for CO2
removal. Singh et al. also studied different amines (linear, cyclic, polyamines etc.)
in order to establish a relation between the structure of the amine and CO2 loading
and absorption rate [18–20]. The results show the potential of the structure variation
and the authors found 1,6-hexanediamine, N,N-dimethyl to have a good CO2 sol-
ubility in comparison with all amines studied.
The solvent screening setup (S.S.S) which is one of the most available techniques
to obtain aqueous amine solution loaded with CO2 is used for the experiments.
A simple solvent screening setup (S.S.S.) equipment constitutes of six glass reactors
(V = 250 mL ± 0.5) which can be operated independently in the temperature range
from [298.15–423.15 K (±1 K)] and the pressure range of [0–6 bars (±0.01 bar)]
(Fig. 1). A magnetic stirrer [speed max = 1500 rpm (±1 rpm)] is used to ensure a
homogeneous contact between the solution and CO2 by creation of a vortex. An
aqueous amine solution 30 wt% (150 g) was prepared and introduced into the glass
reactors. The tests were carried out at 313.15 K with a mixing speed of 500 rpm. To
simulate the gas flow, a blend of 15 vol.% CO2 and 85 vol.% N2 is initially fed to a
make-up vessel to reach a pressure of 2 bar and allowed into the reactors at 15 L/h
with the aid of a mass flow controller. The pressure inside each reactor was kept at
1 bar during the all CO2 absorption experiment. The reaction of CO2 with the
aqueous amine solution is considered complete when equilibrium is reached i.e.
when the flow of CO2 into the set-up is equal to the flow of CO2 out of the set-up.
The experiments were performed at a partial pressure PCO2 ¼ 15 kPa and tem-
perature T = 40 °C which are representative conditions for the absorber in the
post-combustion capture process. The screening results obtained were further
studied as a basis for initial selection of the aqueous amine solution performance for
CO2 absorption and to understand the impact of amine structure on CO2 loading.
180 A. Al Hinai et al.

Fig. 1 Schematic of the solvents screening setup (S.S.S.)

2.2 Heat of Absorption

The heat required in order to regenerate the aqueous amine solution in the stripper
column could be approximated as the contribution of three types of energy: the heat
of desorption of CO2 from the solution, the sensible heat to elevate the temperature
in the stripper and the latent heat contained in the water vapor leaving the stripper
with the CO2-product [5]. The heat of absorption is associated to the quantity of
steam necessary to regenerate the aqueous amine solution. It is assumed that des-
orption of CO2 from the amine solution in the stripper is the inverse reaction of
absorption of CO2 by the aqueous amine solution that happens in the absorber. In
other words, the heat of absorption CO2 is considered nearly equal to the heat of
desorption of CO2. Moreover, study on the heat of CO2 absorption by an aqueous
amine solution shows that its absorption heat depends strongly on the CO2
absorption capacity, the structure of amine and temperature but not significantly
depend on pressure and amine concentration [21].
To evaluate the heat of absorption of the aqueous amine solution, a flow
micro-calorimeter supplied by Thermal Hazard Technology (UK) controlled by a
URC control software was used (Fig. 2). The equipment can be operated from
298.15 to 353.15 K (±1 K) and flow gas can be adjusted from 0 to 20 ml/min
(±0.1 ml/min).
The micro reaction calorimeter is used to determine the heat of absorption of all
the amine samples under study. This type of calorimeter is selected for this study
due to its numerous advantages over the other methods such as reliability, energy
efficiency, reaction yield and safety. For each experiment, the cell containing
Amine-Blends Screening and Characterization for CO2 … 181

Fig. 2 Schematic of flow Micro Reaction Calorimeter (URC)

sample was placed into the calorimeter at 313.15 K and atmospheric pressure. At
the initial stage of the test, there is a variation of power (mW) with time (s) which
was recorded by the software. Next, CO2 gas flows through a desiccant column in
order to remove moisture before entering into the sample cell at a rate of 0.5 ml/min
(±0.01 ml/min). Due to the exothermic nature of the reaction between CO2 and the
aqueous amine solution, the power signal initially increases and then decreases
before it finally becomes constant when the reaction reaches equilibrium.
The difference between the (cell + sample) weight is calculated before and after
the CO2 absorption to obtain the mass of CO2 absorbed during the test. The CO2
loading is determined by using the mass of CO2 and the mass of aqueous amine
solution as in Eq. (5).

ðm1  m0 Þ=44
CO2 loading ¼ ð5Þ
m0  ðM
C1
1
þMC2
2
Þ

Then, with the software URC, the integral heat Q (in kJ) was determined. Finally,
the heat of absorption DH (−kJ/mole of CO2) is calculated by the division of the
integral heat Q by the mole of CO2 absorbed. The experiments are performed at
PCO2 ¼ 15 kPa and T = 40 °C which are the same conditions as in the absorber in
the post-combustion capture process.
For this study, the heat of absorption limit is taken to be below 70 kJ/mole.

3 Results

This section presents results of both single and blended amines. Ten structures of
stand-alone amines which represents various classes, structure, configuration and
chemical group of amines were investigated. Among this 10 structures, two were
conventional amines: N-methyldiethanolamine (MDEA) and diethanolamine
(DEA). The remaining compounds were selected to investigate the impact of the
amine structure on the CO2 loading. 2MAE is obtained from the substitution of one
182 A. Al Hinai et al.

Table 1 Stand-alone amine structures studied


Name Structure CAS
1 N-methyldiethanolamine (MDEA) 105-59-9

HO
2 Diethanolamine (DEA) NH
OH
111-42-2
3 2-(methylamino)ethanol (2MAE) 109-83-1

4 2-(ethylamino)ethanol (2EAE) 110-73-6

5 2-(butylamino)ethanol (2BAE) 111-75-1

6 2-(dimethylamino)ethanol (2DMAE) 108-01-0

7 1-dimethylamino-2-propanol (1DMA2P) 108-16-7

8 N,N-diethylethanolamine (DEEA) 100-37-8

9 3-dimethylamino-1-propanol (3DMA1P) 3179-63-3

10 N,N,N′,N′-Tetramethyl-1,3-propanediamine 110-95-2
(TMDAP)

hydrogen by methyl or ethyl group in MEA (primary amine) structure. From the
2MAE, the methyl group is changed to ethyl for 2EAE and butyl for 2BAE. The
tertiary amine, 2DMAE, is also obtained from MEA to illustrate the effect of
replacing the 2 methyl groups with 2 ethyl groups (DEEA). The effect of the
increase of the carbon length from 2DMAE to 3DMA1P (2 carbons to 3 carbons) is
also considered by using polyamines such as TMDAP. These structures are listed in
Table 1.
Based on the results of the stand-alone amines, eight amine blends were further
studied under three different concentration weight percentage ratio. The blends
included conventional amines to represent various classes and configuration as well
as chemical groups. The blends and the different concentration weight percent
studied are shown in Table 2.

3.1 Stand Alone Amines

3.1.1 CO2 Loading Results

The values obtained for the CO2 loading of the aqueous amine solution at 30 wt%
at 40 °C are listed in Table 3. The results indicate that TMDAP with two amino
groups in its structure has the highest CO2 loading as 1.16 mol CO2/mole amine.
Among all amines (with one amino group), DEEA and 3DMA1P have the highest
CO2 absorption capacity (  0.80 mol CO2/mole amine). MDEA and DEA have the
Amine-Blends Screening and Characterization for CO2 … 183

Table 2 Blend amine structure studied at 30 wt% total of aqueous amines


Blends with 2EAE Blends with 2MAE Concentration
(wt%)
2EAE + MDEA 2MAE + MDEA 5 + 25
2EAE + 1DMA2P 2MAE + 1DMA2P 5 + 25
2EAE + 2DMAE 2MAE + 2DMAE 5 + 25
2EAE + 3DMA1P 5 + 25
2EAE + TMPDA 2MAE + TMPDA 5 + 25
2EAE + MDEA 2MAE + MDEA 10 + 20
2EAE + 1DMA2P 2MAE + 1DMA2P 10 + 20
2EAE + 2DMAE 2MAE + 2DMAE 10 + 20
2EAE + 3DMA1P 10 + 20
2EAE + TMPDA 2MAE + TMPDA 10 + 20
2EAE + MDEA 2MAE + MDEA 15 + 15
2EAE + 1DMA2P 2MAE + 1DMA2P 15 + 15
2EAE + 2DMAE 2MAE + 2DMAE 15 + 15
2EAE + 3DMA1P 15 + 15
2EAE + TMPDA 2MAE + TMPDA 15 + 15

Table 3 Absorption capacity of CO2 in aqueous amine solution 30 wt% at 40 °C and 15 kPa CO2
Amine solution Concentration Absorption capacity
wt% mol CO2/mol amine
1 N-methyldiethanolamine (MDEA) 30 0.52
2 Diethanolamine (DEA) 30 0.53
3 2-(methylamino)ethanol (2MAE) 30 0.56
4 2-(ethylamino)ethanol (2EAE) 30 0.67
5 2-(butylamino)ethanol (2BAE) 30 0.69
6 2-(dimethylamino)ethanol (2DMAE) 30 0.73
7 1-dimethylamino-2-propanol (1DMA2P) 30 0.72
8 N,N-diethylethanolamine (DEEA) 30 0.90
9 3-dimethylamino-1-propanol (3DMA1P) 30 0.89
10 Tetramethyl-1,3-diaminopropane (TMDAP) 30 1.16

lowest CO2 absorption among all tested amines with values of 0.52 and 0.53 mol
CO2/mole amine, respectively.
The screening of the CO2 absorption capacity results acquired with the S.S.S.
equipment shows that amines which contain an alkyl group (methyl, ethyl or t-butyl)
attached to the nitrogen or near it have a high CO2 loading. It is also observed that
alcohol group near the amino group have a negative effect on the CO2 absorption.
Additionally, the polyamines studied in this work showed the highest CO2 loading
among all amines studied. These results are in accordance with work from Singh et al.
where the impact of the alkyl and number of amino groups in the amine molecule
enhance the absorption capacity of the amine based solvents [15, 16, 22].
184 A. Al Hinai et al.

3.1.2 Heat of Absorption

The heat of absorption was measured at 313.15 K and atmospheric pressure and the
results are presented in Table 4. The experiments for all aqueous amines solutions
were conducted at least 3 times to ensure the consistency of values obtained.
Using the heat of absorption limit as 70 kJ/mole of CO2, there are four amines
with heat of absorption which falls above 70 kJ/mole and they include DEA,
2MAE, 2BAE and DEEA while the other amines are below 70 kJ/mole of CO2,
From these results, we assume that the potential amines for CO2 capture application
will have a CO2 loading up to 0.60 and a heat of absorption below 70 kJ/mole of
CO2. The amines considered are MDEA, 1DMA2P, 3DMA1P, TMDPA, 2DMAE,
2MAE, 2BAE 2EAE.
In order to evaluate the operational validity of the calorimeter equipment used in
this work, the heat of CO2 absorption of MEA, MDEA, DEA and AMP 30 wt% at
313.15 K and atmospheric pressure was measured and compared with the available
literature data [11, 23, 24]. A value of −52.51 kJ/mole of CO2 (aCO2 = 0.74) for
MDEA was determined, and 85.13 kJ/mole of CO2 (aCO2 = 0.59) for MEA. The
results for DEA and AMP also has been found to be −74.24 kJ/mole of CO2 (aCO2
= 0.61) and −80.91 kJ/mole of CO2 (aCO2 = 0.78) respectively [25]. The values
show a good agreement with the literature with 2% uncertainty, which implies that
the calorimeter have been used has a capability and consistency to be worked on.

Table 4 Heat of absorption of CO2 in aqueous amine solution 30 wt% at 40 °C


Amine solution Concentration CO2 Heat of
wt% loading absorption
DH (kJ/mol
of CO2)
1 N-methyldiethanolamine (MDEA) 30 0.74 −52.5
2 Diethanolamine (DEA) 30 0.61 −74.2
3 2-(methylamino)ethanol (2MAE) 30 0.67 −73.8
4 2-(ethylamino)ethanol (2EAE) 30 0.71 −69.0
5 2-(butylamino)ethanol (2BAE) 30 0.73 −74.4
6 2-(dimethylamino)ethanol 30 0.77 −63.3
(2DMAE)
7 1-dimethylamino-2-propanol 30 0.83 −60.7
(1DMA2P)
8 N,N-diethylethanolamine (DEEA) 30 0.83 −73.2
9 3-dimethylamino-1-propanol 30 0.85 −54.6
(3DMA1P)
10 Tetramethyl-1,3-diaminopropane 30 1.32 −59.9
(TMDAP)
Amine-Blends Screening and Characterization for CO2 … 185

3.2 Amine Blends

Tertiary amines have favourable thermodynamics characteristics. However, it has


drawback due to its low reactivity compared to MEA and other conventional amines.
This problem can be overcome by blending them with highly reactive amines that
have good thermodynamics characteristics such as high absorption capacity and low
heat of absorption. Potential candidates for this are 2EAE, 2MAE and 2BAE.
Two compounds 2EAE and 2MAE were selected as the basis for the amine blends
system. For the purpose of clarification in this chapter, system A will consist of 2EAE
blended with MDEA, 1DMA2P, 2DMA2P, 3DMA2P and TMPAD while system B
will consist of 2 MAE blended with MDEA, 1DMA2P, 2DMA2P and TMPAD. In
both systems, the blends concentrations are 5, 10 and 15 wt% of 2EAE or 2MAE.

3.2.1 CO2 Loading Result

For system A, the results shown in Table 5 indicate that the lowest CO2 loading is
0.71 (2EAE 5 wt%/MDEA 25 wt%) while among all the three different

Table 5 Heat of absorption of CO2 in aqueous 2EAE blends amine solution 30 wt% at 40 °C
Blends of System A CO2 loading Heat of absorption
(mol CO2/mol of amine) (kJ/mol of CO2)
2EAE MDEA
5 wt% 25 wt% 0.71 −59.5
10 wt% 20 wt% 0.74 −62.8
15 wt% 15 wt% 0.75 −71.5
2EAE 1DMA2P
5 wt% 25 wt% 0.82 −63.9
10 wt% 20 wt% 0.84 −71.4
15 wt% 15 wt% 0.76 −75.2
2EAE 2DMA2P
5 wt% 25 wt% 0.83 −61.5
10 wt% 20 wt% 0.8 −67.8
15 wt% 15 wt% 0.78 −71.4
2EAE 3DMA2P
5 wt% 25 wt% 0.83 −63.9
10 wt% 20 wt% 0.81 −70.6
15 wt% 15 wt% 0.72 −75.6
2EAE TMDAP
5 wt% 25 wt% 1.25 −62.3
10 wt% 20 wt% 1.14 −67.8
15 wt% 15 wt% 1.02 −69.2
186 A. Al Hinai et al.

concentrations, the 2EAE + TMDPA blend has the highest CO2 loading. 2EAE
5 wt%/TMDPA 25 wt 2EAE 10 wt%/TMDPA 20 wt%—2EAE 15 wt%/TMDPA
15 wt% having 1.25, 1.14 and 1.02 respectively.
Comparison was made between the blend containing 2DMAE (tertiary amine
where the number of carbon between the nitrogen and the alcohol is two) with that
containing 3DMA1P (tertiary amine where the number of carbon between the
nitrogen and the alcohol is three). The increase in the carbon length does not have
significant influence on the absorption of CO2 at the different concentrations except
for the 15% weight concentration of 2EAE where the 2DMAE and 3DMA1P
blends have a CO2 loading which are 0.78 and 0.72, respectively.
Except for the blends 2EAE/TMPDA, all other blends have a CO2 absorption
calculated from calorimeter equipment between 0.7 and 0.9. 2EAE/TMPDA blends
have the highest CO2 loading because it contains a polyamine with two nitrogen
atoms in its structure. However, it can be seen from the results in Table 5 that the
CO2 loading of 2EAE/TMPDA reduces as the concentration weight percentage of
2EAE increases in the blend.
For system B, the result from blending 2MAE with 2DMAE, 3DMA1P,
1DMA2P, MDEA and TMPAD at 40 °C are presented in Table 6. The results show
that the lowest loading obtained is 0.54 from blending (2MAE 5 wt%/MDEA
25 wt%) while the highest loading is obtained from the blend of 2MAE with
TMPAD as 2MAE 5 wt%/TMPAD 25 wt% (aCO2 = 1.35), 2MAE 10 wt%/
TMPAD 20 wt% (aCO2 = 1.22), and for 2MAE 15 wt%/TMPAD 15 wt%
(aCO2 = 1.04).

Table 6 Heat of absorption of CO2 in aqueous 2MAE blends amine solution 30 wt% at 40 °C
Blends of System B CO2 loading Heat of absorption
(mol CO2/mol of amine) (kJ/mol of CO2)
2MAE MDEA
5 wt% 25 wt% 0.54 −58.6
10 wt% 20 wt% 0.56 −61.0
15 wt% 15 wt% 0.57 −64.6
2MAE 1DMA2P
5 wt% 25 wt% 0.68 −62.4
10 wt% 20 wt% 0.65 −65.3
15 wt% 15 wt% 0.63 −69.5
2MAE 2DMA2P
5 wt% 25 wt% 0.69 −65.7
10 wt% 20 wt% 0.65 −67.3
15 wt% 15 wt% 0.63 −70.2
2MAE TMDAP
5 wt% 25 wt% 1.35 −60.4
10 wt% 20 wt% 1.22 −64.5
15 wt% 15 wt% 1.04 −66.5
Amine-Blends Screening and Characterization for CO2 … 187

Comparing 1DMA2P with 2DMAE which are MEA derivatives where one
hydrogen has been substituted with methyl group for 1DMA2P and one hydrogen has
been substituted with ethyl group for 2DMAE, the results show that there are no
significate difference in terms of loadings. The loading for 2MAE 5 wt% + 1DMA2P
25 wt% and 2MAE 5 wt% + 2DMA2P were identical (aCO2 = 0.56). Also for
2MAE 10 wt% + 1DMA2P 20 wt% and 2MAE 10 wt% + 2DMA2P 20 wt%, the
CO2 loading is 0.63.

3.2.2 Heat of Absorption

Similar to the stand alone amines experiments, values for the heat of absorption for
this blends were obtained at 313.15 K and atmospheric pressure. These results are
presented in Tables 5 and 6.
The results from system A show in general that the blends have a much lower heat
of absorption than the stand alone amines with the highest heat of absorption being
−75.25 kJ/mole of CO2 for 2EAE 15 wt%/3DMAIP 15 wt% while the lowest is
−59.53 kJ/mole of CO2 for 2EAE 5 wt%/MDEA 25 wt%. It is obvious from this
study that as the amine concentration of 2EAE increases in the blends 30 wt% total
(from 5 to 15 wt%), the heat of absorption increases. For 2EAE 5 wt%, the heat of
absorption is between 59.53 and 63.91 kJ/mole of CO2 while for the 2EAE 15 wt%,
the heat of absorption is between 69.15–75.25 kJ/mole of CO2 which is the highest
value.
The blends which have a good heat of absorption with a value below 70 kJ/mole
of CO2 are 2EAE 5 wt%/MDEA 25 wt%—2EAE 10 wt%/MDEA 20 wt%—2EAE
5 wt%/1DMA2P 25 wt%—2EAE 5 wt%/3DMA1P 25 wt%—2EAE 10 wt%/
2DMAE 20 wt%—2EAE 5 wt%/2DMAE 25 wt%—2EAE 5 wt%/TMDPA 25 wt%—
2EAE 10 wt%/TMDPA 20 wt%—2EAE 15 wt%/TMDPA 15 wt%.
Our results show similar values of CO2 loading and heat of absorption for the
2EAE-1DMA2P AND 3DMA1P blends. For 2EAE 5, 10 and 15 wt%, the CO2
loading for 1DMA2P/3DMA1P are 0.82/0.83, 0.84/0.81 and 0.76/0.72 respectively
while the heat of absorption are 63.91/63.85, 71.39/70.64 and 75.16/75.25,
respectively.
For the three groups of amine concentration of 2EAE (5 and 10%), the
2EAE/MDEA have the lowest heat of absorption (−59.53 and −62.82 kJ/mole of
CO2), respectively. These blends also have the lowest CO2 loading of 0.71, 0.74
and 0.75 for 2EAE 5 wt%/MDEA 25 wt%, 2EAE 10 wt%/MDEA 20 wt% and
2EAE 15 wt%/MDEA 15 wt% respectively. Out of all the studied blends, this
blend can be considered for CO2 capture application.
This study shows that the blend with 2EAE 5 wt% having the highest CO2
loading of 1.25 has the third-lowest heat of absorption of 62.29 kJ/mole of CO2 and
from this, we assume that the most promising amine blend for CO2 capture
application will be 2EAE 5 wt%/TMDPA 25 wt%. This blend concentration fulfills
the two major concerns in carbon capture which is a high CO2 loading and low heat
of absorption.
188 A. Al Hinai et al.

On the other hand, results from system B show in general that blends have a
higher heat of absorption than the stand alone amines and system A. The highest
heat of absorption is −84.86 kJ/mole of CO2 for 2MAE 15 wt%/MDEA 15 wt%
while the lowest is −65.13 kJ/mole of CO2 for 2MAE 5 wt%/1DMA2P 25 wt%.
Also, the system B illustrates similar to system A, that for all the blends, a higher
concentration of 2MAE from 5 to 15 wt%, will have a higher heat of absorption.
The blends which have a good heat of absorption, i.e. with a value below
70 kJ/mole of CO2, are 2MAE 5 wt%/MDEA 25 wt%, 2MAE 5 wt%/1DMA2P
25 wt%, 2MAE 5 wt%/2DMA2P 25 wt%, 2MAE 5 wt%/TMDPA 25 wt%, 2MAE
10 wt%/TMDPA 20 wt%.
Studying the five blends with low heat of absorption, results show that three of
the five blends have a CO2 loading less than 0.70-2MAE 5 wt%/MDEA 25 wt%
(0.54 mol CO2/mol of amine), 2MAE 5 wt%/1DMA2P 25 wt% (0.68 mol
CO2/mol of amine) and 2MAE 5 wt%/2DMA2P 25 wt% (0.69 mol CO2/mol of
amine). 2MAE 5 wt%/TMDPA 25 wt% and 2MAE 10 wt%/TMDPA 20 wt% have
a high CO2 loading of 1.35 and 1.22 mol CO2/mol of amine respectively and thus
have the potential to be used in CO2 capture application.
In general, for both systems, TMDPA 25 wt% is a good blend combination with
5 wt% of 2EAE or 2MAE.

4 Conclusions

Following the results from this study, it can be concluded that CO2 loading is the
highest and in the range of 1.04–1.45 by blending either 2EAE or 2MAE with
TMDAP. Increasing the concentration of 2EAE or 2MAE in the blend causes
increase in the heat of absorption. Blends containing 5 wt% of 2MAE or 2EAE in
general shows the lowest heat of absorption but due to the low CO2 loading
observed, further investigation is not advised to be carried out on MDEA,
1DMA2P, 2DMA2P and 3DMA2P. It is important to follow up this work by
investigating the kinetics of blends of 2MAE or 2EAE with TMDAP.

References

1. Wang M, Lawal A, Stephenson P, Sidders J, Ramshaw C (2011) Post-combustion CO2


capture with chemical absorption: a state-of-the-art review. Chem Eng Res Des 89:1609–1624
2. Samanta A, Zhao A, Shimizu GKH, Sarkar P, Gupta R (2012) Post-combustion CO2 capture
using solid sorbents: a review. Ind Eng Chem Res 51:1438–1463
3. Jou F-Y, Mather AE, Otto FD (1995) The solubility of CO2 in a 30 mass percent
monoethanolamine solution. Can J Chem Eng 73:140–147
4. Supplement
5. Quang DV, Rabindran AV, El Hadri N, Abu-Zahra MR (2013) Reduction in the regeneration
energy of CO2 capture process by impregnating amine solvent onto precipitated silica. Eur Sci J 9
Amine-Blends Screening and Characterization for CO2 … 189

6. Rodier L, Ballerat-Busserolles K, Coxam J-Y (2010) Enthalpy of absorption and limit of


solubility of CO2 in aqueous solutions of 2-amino-2-hydroxymethyl-1,3-propanediol, 2-[2-
(dimethyl-amino)ethoxy] ethanol, and 3-dimethyl-amino-1-propanol at T = (313.15 and
353.15) K and pressures up to 2 MPa. J Chem Thermodyn 42:773–780
7. Arshad MW, von Solms N, Thomsen K, Svendsen HF (2013) Heat of absorption of CO2 in
aqueous solutions of DEEA, MAPA and their mixture. Energy Procedia 37:1532–1542
8. Arcis H, Ballerat-Busserolles K, Rodier L, Coxam J-Y (2012) Enthalpy of solution of carbon
dioxide in aqueous solutions of triethanolamine at temperatures of 322.5 K and 372.9 K and
pressures up to 5 MPa. J Chem Eng Data 57:3587–3597
9. Carson JK, Marsh KN, Mather AE (2000) Enthalpy of solution of carbon dioxide in (water
monoethanolamine, or diethanolamine, orN-methyldiethanolamine) and (water + mono-
ethanolamine + N-methyldiethanolamine) atT = 298.15 K. J Chem Thermodyn 32:
1285–1296
10. Kim I, Hoff KA, Hessen ET, Haug-Warberg T, Svendsen HF (2009) Enthalpy of absorption
of CO2 with alkanolamine solutions predicted from reaction equilibrium constants. Chem Eng
Sci 64:2027–2038
11. Mathonat C, Majer V, Mather AE, Grolier JPE (1997) Enthalpies of absorption and solubility
of CO2 in aqueous solutions of methyldiethanolamine. Fluid Phase Equilib 140:171–182
12. Chowdhury FA, Okabe H, Yamada H, Onoda M, Fujioka Y (2011) Synthesis and selection of
hindered new amine absorbents for CO2 capture. Energy Procedia 4:201–208
13. Chowdhury FA, Yamada H, Higashii T, Goto K, Onoda M (2013) CO2 capture by tertiary
amine absorbents: a performance comparison study. Ind Eng Chem Res 52:8323–8331
14. Ma’mun S, Jakobsen JP, Svendsen HF, Juliussen O (2006) Experimental and modeling study
of the solubility of carbon dioxide in aqueous 30 mass% 2-((2aminoethyl)amino)ethanol
solution. Ind Eng Chem Res 45:2505–2512
15. Singh P, Niederer JPM, Versteeg GF (2009) Structure and activity relationships for
amine-based CO2 absorbents-II. Chem Eng Res Des 87:135–144
16. Singh P, Brilman DWF, Groeneveld MJ (2011) Evaluation of CO2 solubility in potential
aqueous amine-based solvents at low CO2 partial pressure. Int J Greenhouse Gas Control
5:61–68
17. Singh P, Niederer JPM, Versteeg GF (2007) Structure and activity relationships for amine
based CO2 absorbents—I. Int J Greenhouse Gas Control 1:5–10
18. Arcis H, Rodier L, Ballerat-Busserolles K, Coxam J-Y (2008) Enthalpy of solution of CO2 in
aqueous solutions of methyldiethanolamine at T = 322.5 K and pressure up to 5 MPa. J Chem
Thermodyn 40:1022–1029
19. Arcis H, Ballerat-Busserolles K, Rodier L, Coxam J-Y (2011) Enthalpy of solution of carbon
dioxide in aqueous solutions of monoethanolamine at temperatures of 322.5 K and 372.9 K
and pressures up to 5 MPa. J Chem Eng Data 56:3351–3362
20. Kim I, Svendsen HF (2007) Heat of Absorption of carbon dioxide (CO2) in mono-
ethanolamine (MEA) and 2-(aminoethyl)ethanolamine (AEEA) solutions. Ind Eng Chem Res
46(2007/08/01):5803–5809
21. Kim I, Svendsen HF (2011) Comparative study of the heats of absorption of post-combustion
CO2 absorbents. Int J Greenhouse Gas Control 5:390–395
22. Singh P, Versteeg GF (2008) Structure and activity relationships for CO2 regeneration from
aqueous amine-based absorbents. Process Saf Environ Prot 86:347–359
23. Sodiq A, Rayer AV, Olanrewaju AA, Abu Zahra MRM (2014) Reaction kinetics of carbon
dioxide (CO2) absorption in sodium salts of taurine and proline using a stopped-flow
technique. Int J Chem Kinet 46:730–745
24. Caplow M (1968) Kinetics of carbamate formation and breakdown. J Am Chem Soc 90
(1968/11/01):6795–6803
25. Chowdhury FA, Okabe H, Yamada H, Onoda M, Fujioka Y (2011) Synthesis and selection of
hindered new amine absorbents for CO2 capture. Energy Procedia 4:201–208
Post-combustion Carbon Dioxide
Capture with Aqueous (Piperazine +
2-Amino-2-Methyl-1-Propanol) Blended
Solvent: Performance Evaluation
and Analysis of Energy Requirements

Sukanta K. Dash

Abstract Post-combustion CO2 capture (PCC) and its sequestration has been
found to be a viable option for reducing CO2 in the earth’s atmosphere. There are
many technological options for separation of CO2 from a post combustion gas
stream. However, regenerative chemical absorption process is considered to be a
near-term feasible solution for this. In regenerative chemical absorption, the key
component is the solvent, which plays a major role in the process efficiency and
economics. There are many conventional and newer commercial solvents with
patented technologies available for this process. In this chapter, the suitability of
aqueous AMP along with PZ as an energy efficient mixed solvent for the PCC
process have been presented by critically analyzing the absorption rate, equilibrium
thermodynamics, reaction kinetics as well as regeneration energy requirement.
Energy analysis from bench scale and pilot scale studies, and modelling and sim-
ulation work have been investigated and compared with the bench marked solvent
MEA. The role of important solvent properties for this application, i.e., density,
viscosity, physical gas solubility, reaction mechanism and kinetics, equilibrium
solubility and heat of absorption are found to be suitable for the CO2 capture by
AMP + PZ solvent. Besides, it is also found that the negative impact such as,
corrosion, thermal and oxidative degradation, possible amine and nitrosamine
emission from the capture plant have less impact to the environment. Heat energy
requirements of this process are found to be in the range of 2.9–3.7 GJ/tCO2 for
different conditions such as, %CO2 capture, etc., and from different study. This
energy requirement is about 20% less than that of the bench marked MEA solvent.
All this performance indicators show that the AMP + PZ blended solvent is a
competitive energy efficient alternative one for CO2 capture by chemical
absorption.

S.K. Dash (&)


Department of Chemical Engineering, Pandit Deendayal Petroleum University,
Gandhinagar 382007, Gujarat, India
e-mail: sukantakdash@gmail.com

© Springer International Publishing AG 2017 191


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_9
192 S.K. Dash

Nomenclature

AMP Amino-2-methyl-1-propanol
AMPH+ Protonated AMP
CCS Carbon capture and storage
CCTs Clean coal technologies
CH4 Methane
CO2‒ 3 Carbonate
CO2 Carbon dioxide
DAE Differential and algebraic equations
DRS Data regression system
ECO2 Enhancement factor
eNRTL model Electrolyte non-random two-liquid model
GC Gas chromatography
HCO‒3 Bicarbonate
He Helium
H+PZCOO‒ Protonated PZ carbamate
k Reaction rate constant
L/G Liquid to gas ratio
MDEA Methyldiethanolamine
MEA Monoethanolamine
N2 Nitrogen
N 2O Nitrous oxide
NOx Nitrogen oxide
O2 Oxygen
pamine Partial pressure of amine
PCC Post-combustion CO2 capture
pCO2 Partial pressure of CO2
pH 2 O Partial pressure of H2O
PZ Piperazine
PZCOO− PZ carbamate
PZ(COO−)2 PZ dicarbamate
PZH+ Protonated PZ
R Ideal gas constant
RA Rate of absorption
SOx Sulfur oxide
T Absolute temperature
VLE Vapor-liquid equilibrium
aCO2 CO2 loading (mol CO2/mol amine)
apCO2;lean CO2 loading at regenerator outlet pressure
apCO2;rich CO2 loading at absorber inlet pressure
DHabs Heat of absorption
c Activity coefficient
u Fugacity coefficient
dni Mol change of species with mol change of CO2 in reaction
dCO2
Post-combustion Carbon Dioxide Capture with Aqueous … 193

1 Introduction

Fossil fuels based power plants are the largest contributor to anthropogenic CO2
emission in this planet. At present coal combustion generates about 40% of the
world’s electrical energy and coal is likely to maintain this major role for the
coming decades. However, a central issue in this will be the increasing use of
Carbon Capture and Storage (CCS) and Clean Coal Technology (CCT) initiatives
to mitigate the anthropogenic greenhouse gas concentration in the atmosphere. Both
will be important pathways in maintaining the use of coal as a power generating
fossil fuel in the future.
Out of the amine based chemical absorption processes, the MEA based one is
considered the bench mark process for CO2 capture. This process has been practised
for several years now for CO2 removal from natural gas streams and for synthesis
gas processing. It is easier to separate CO2 from a high pressure gas stream but
difficult to separate it from the flue gas stream of a power plant, which is essentially
at atmospheric pressure. Besides, large volumetric flow rates of flue gas from coal
combustion necessitate the requirement of a large diameter absorption column.
Again, apart from CO2 and N2, the flue gas often contains SOx, NOx and significant
amount of O2. Presence of these components put up further resistance for imple-
mentation of chemical process for CO2 capture from atmospheric flue gas streams,
as these components can degrade the amine solvent and lower its capacity for CO2
intake. Hence, the major R&D thrust for CO2 capture is on the development of high
reactivity, high equilibrium capacity, and higher degradation resistance solvents
with lower regeneration energy requirement to make the process economically
viable.
Advanced solvents can enhance the technology for providing lower energy
consumption with equivalent or better mass transfer rates and lower
degradation/corrosion impact than those of the bench marked MEA process. The
energy requirement can be reduced by using activated solvents based on tertiary
amines or sterically hindered amines (SHA). One most commercially used tertiary
amine is MDEA and SHA is AMP. As the reaction rates of these amines with CO2
are relatively lower, activators/promoters are necessary to maintain equivalent rates
of reaction and mass transfer in the newer solvent. Diamines such as, PZ shows
very high CO2 absorption capacity and reaction rate as single solvent for CO2
capture as well as when blended with other amines such as MDEA or AMP [1, 2].
Use of mildly hindered amines such as AMP may also lead to a lower heat of
absorption. When the amine is not too greatly hindered, it will provide equivalent
mass transfer rates at high loading because the reactive amine is not significantly
depleted by reaction with CO2. AMP is characterized by a moderate CO2 absorption
rate and a high CO2 stoichiometric loading capacity. But, since AMP is the hin-
dered form of MEA, it absorbs CO2 at a slower rate than MEA, and the low pCO2 of
flue gas prevents AMP from realizing the one mole per mole CO2 loading which is
stoichiometrically possible for non-carbamate forming amines. While, on the other
hand, AMP based solvents offer good absorption and regeneration efficiencies
194 S.K. Dash

towards CO2 as compared to those of MEA [3]. Again, from a bench scale absorber
pilot plant study by Hairul et al. [4] shows that the mass transfer performance of
AMP + PZ blended solvent has superior mass transfer characteristics over the
single AMP solvent. In view of these facts (AMP + PZ) solvent has been chosen
for the present work. On going through the resent literature, it is also found that
there are lot of R&D focus on this energy efficient solvent.

2 Energy Requirement

The objective of the energy efficient solvent is to reduce both the Capex and Opex
in a CO2 capture plant attached to a fossil fuel based power plant. The solvent
should have appropriate equilibrium solubility and fast reaction kinetics.
Furthermore, stability, corrosion and environmental aspects are important.
A complete survey of reported energy demand of the PCC process for different
patented processes and solvents can be found in Budzianowski [5]. In addition to
heat energy, a CO2 capture plant may require electrical energy to drive machineries.
The total electrical energy requirement of a 400 MW coal fired power plant with
MEA based solvent for CO2 capture is *32.5 MW which is for the inlet CO2
blower to the absorber, recompression of the stripped CO2 and miscellaneous
electrical energy needed for the solvent pumps etc., [6]. Overall this electrical
energy is a small fraction of the total regeneration energy requirement for the
process. As the scope of this chapter is to focus energy efficient solvent for CO2
capture by absorption process, the regeneration energy demand, of few patented
process are discussed here.
Kansai Electric and Mitsubishi Heavy Industries (MHI), commercially employ
PCC solvents based on aq. solution of a sterically hindered amine known as
“KS-1TM”. The organization claims that the regeneration energy requirement for the
KS-1TM solvent is *3 GJ/tCO2 with the use of their process and proprietary
equipments [7]. This energy demand is about 20% lower than that of MEA solvent,
which is *3.7 GJ/tCO2 for CO2 capture at coal-fired power plants [8]. MHI also
claims that the KS-1TM process have lower degradation and higher corrosion
resistance as compared to the MEA process. Siemens and E.ON, use another sol-
vent based on aq. amino acid salt known as “Siemens AAS” [9]. They claim that
‘Siemens AAS’ solvent demands 2.7 GJ/tCO2 for regeneration which operated in a
full-scale capture plant. It is realized that in many cases the reported numbers of
energy demand are just claimed by the vendors or developers. These are just
indicatives, applicable in the specified conditions and should not be compared with
each other.
Since the chemical absorption of CO2 still remains an energy intensive process
in spite of the developments achieved so far, the R&D focus is on advanced and
newer energy efficient solvents and processes for CO2 capture.
The Chapter begins with a discussion on the regeneration energy demand of
commercial solvents claimed by their developers and from literature review in
Post-combustion Carbon Dioxide Capture with Aqueous … 195

Sect. 2. A brief review of physico-chemical properties, chemical and aqueous phase


thermodynamics, amine-CO2 reaction kinetics, as well as, a summary of the pre-
vious work in respect of CO2 + AMP + H2O, CO2 + PZ + H2O and
CO2 + AMP + PZ + H2O systems are presented in Sects. (3–5). Experimental and
theoretical studies of ternary systems such as vapour-liquid equilibrium (VLE) data
from the equilibrium cell measurements and the electrolyte non-random two-liquid
(eNRTL) modelling of VLE to predict the equilibrium solubility and speciation of
these solvents are presented in Sects. 4 and 5. VLE data of CO2 in aq. (AMP + PZ)
in a wide range of compositions, temperatures and CO2 partial pressures, ðpCO2 Þ
have also been discussed along with the estimation of heats of absorption of CO2 in
these solvents. The absorption rate measurements of CO2 in this blended solvent
using wetted wall contactor in our laboratory and results of absorption measure-
ments from other literature sources are presented as well. The experimental results
and modeling work of CO2 absorption in aq. (AMP + PZ) and the
physico-chemical properties necessary for the rate model along with simulation
study of absorber-regenerator have been described. Amine degradation and corro-
sion have been explained in Sect. 6. Energy requirement analysis obtained from a
parametric study and bench scale and pilot tests information from the literature are
described in Sect. 7. Finally general conclusions and recommendations in this
energy efficient process have been discussed.

3 Physico-Chemical Properties of AMP + PZ Solvent

Knowledge of physico-chemical properties such as viscosity and density of the


solvent is needed for the selection, operation and energy requirement analysis of
process equipments such as, the lean-rich heat exchanger, pumps, and for the
hydrodynamic calculations of absorber and stripper. Again, these data are required
for estimating the liquid phase diffusivity by modified Stokes-Einstein equation.
Viscosity, density and physical gas solubility (Henry’s constant) of the mixed
solvent AMP + PZ at different temperatures is also required in mass transfer and
kinetics rate modelling in the absorber and regenerator as these quantities are
influencing the liquid phase mass transfer coefficients. The viscosity and density of
aq. (AMP + PZ) has been reported at different temperatures and relative concen-
tration of AMP and PZ. The reader may refer Paul and Mandal [10], Sun et al. [11],
Samanta and Bandyopadhyay [12], Dash et al. [13] for a detail discussion about the
experimental work as well as development of correlations for viscosity, density of
unloaded solvents and physical gas solubility. It is reported by these authors that the
density and viscosity decreases with increasing temperature and decreasing PZ
concentration in the solution. Viscosity and density data for the CO2 loaded
solution could be useful for plant design but these data are scarce as it is difficult to
maintain the CO2 loading in the solvent and temperature simultaneously while
196 S.K. Dash

measuring these properties at different temperatures. Any future research on this


topic is highly desirable.

4 Thermodynamics of Quaternary Systems:


CO2 + AMP + PZ + H2O

Thermodynamic study includes the understanding of CO2 equilibrium solubility,


pH of loaded solution, heat of absorption, distribution of chemical species in the
loaded solvent (speciation) and amine vaporization losses. The CO2 equilibrium
solubility in liquid phase has contributions of both physical solubility and solubility
by chemical reaction. The VLE data at different temperatures, solvent concentra-
tions, and CO2 partial pressures have significant role in the design and optimization
of absorber-regenerator as it is necessary to calculate the driving force for these
equilibrium governed processes. Normally, absorber model is represented mathe-
matically by a set of differential and algebraic equations (DAE). Information on
physical and chemical equilibrium acts as boundary conditions for the DAE. For a
detail discussion on this, the reader may refer Dash et al. [13] for the development
of DAE for the CO2 + AMP + PZ + H2O system. The equilibrium solubility of
CO2 in the liquid phase also determines the solvent recirculation rate in the
absorber-regenerator set up. Besides, it also fixes the CO2 concentration in the
outgoing gas stream from the absorber and regenerator. Appropriate mathematical
model is necessary to represent the VLE of CO2 in the solvents. Among the
different models used to represent VLE of CO2 in amine solvents, the frequently
used model considers the activity coefficient-fugacity coefficient model for the
phase equilibrium, known as the c-u approach. One such popular model is elec-
trolyte non-random two-liquid (eNRTL) model used by Dash et al. [14–16] to
model VLE of CO2 in aqueous PZ, aqueous AMP and aqueous (AMP + PZ)
solvents.

4.1 Experimental Techniques for the Measurement of VLE

The various experimental techniques used to measure the vapor (CO2) and liquid
(solvent) equilibrium are classified in two categories, (i) static method and (ii) dy-
namic method. In the static method, the total pressure ðpCO2 þ pH2 O þ pamine Þ is
measured with a pressure measuring device. The mixing of vapor and liquid phases
is carried out by agitating the liquid phase by a shaft mounted with impellers or by
using a magnetic stirrer. The pCO2 is then obtained by subtracting the water vapour
pressure from the total pressure thus neglecting the amine vapour pressure. It is
known that, this method is highly productive to generate equilibrium data at high
temperature and pressure for CO2-amine-H2O systems. In the dynamic method the
Post-combustion Carbon Dioxide Capture with Aqueous … 197

vapour composition can be analyzed by using gas chromatography (GC), infrared


analyzer, and mass spectrometer. To maintain a relative high pressure in the system,
N2, CH4, or He is usually used as the inert gas. Liquid samples can be analyzed by
titration. However, for mixed or blended amine solvents, liquid analysis by titration
is difficult. A complete discussion of the different experimental methods for VLE
measurement of CO2 in AMP + PZ solvent and its modelling, the reader may refer
Bruder et al. [17], Dash et al. [14, 15], Tong et al. [18], Hartono et al. [19],
Haghtalab et al. [20] and Halim et al. [21].

4.2 VLE of CO2 in Aqueous PZ

Dash et al. [14, 15, 22] have obtained the VLE data of CO2 in aq. PZ, aq. AMP and
aq. (AMP + PZ) using the static method explained earlier. Incorporating the
eNRTL equation for VLE modelling, they determined the interaction parameters by
data regression using the Data Regression System (DRS) available in Aspen Plus®
simulation software. As VLE of CO2 in single amine such as PZ and AMP are
necessary to formulate a model to represent the VLE of CO2 in blended amines
such as AMP + PZ, it is worthwhile to report here some of the results of VLE of
CO2 in aq. PZ and aq. AMP. Figure 1 shows a typical VLE plot of CO2 over
0.6 mol.dm3 PZ at various temperatures. This VLE is presented by plotting the
pCO2 , kPa) verses CO2 loading in liquid phase (aCO2 mol CO2/mol amine). The
model results have been obtained using the interaction parameters reported by Dash
et al. [14].
The behaviour of the equilibrium curves can be explained by observing the trend
in the experimental data and model predictions. This trend is, pCO2 increases with

Fig. 1 Equilibrium pCO2 over


aqueous 0.6 mol/dem3 aq. PZ
100 100
at 313 and 343 K. T = 343 K
Experimental result from
CO 2 Partial Pressure [kPa]

reference [23, 24] 10 10

1 1

Expt. Data: [PZ]=0.6 M


0.1 0.1
-Bishnoi, T= 313 K
-Bishnoi; T= 343 K
-Derks; T= 343 K
0.01 0.01
-Derks; T= 313 K
T = 313 K -eNRTL model
-eNRTL model
1E-3 1E-3
0.0 0.2 0.4 0.6 0.8 1.0 1.2

Loading (mol CO2/mol PZ)


198 S.K. Dash

increase in CO2 loading in liquid phase (expressed in mol CO2/mol PZ) and
temperature.
One important information obtained from the thermodynamic modelling are the
solution chemistry and liquid phase speciation. After estimating the c values of the
species in liquid phase, the concentration of the species can be calculated with
respect to the extent of reaction with CO2. Normally, the extent of reaction is
expressed in terms of liquid phase CO2 loading and can be predicted up to the
stoichiometric limit of loading of CO2 in the solvent which is 2 mol of CO2/mol of
PZ theoretically. Figure 2 shows the reaction stoichiometry i.e., the relation among
change of moles of any species in the liquid phase with respect to change in number
of moles of CO2 verses CO2 loading. Figure 3 shows the speciation of
CO2 + PZ + H2O system predicted using eNRTL model [14]. As usual PZ is

Fig. 2 Model predicted -


6 + + - HCO 3
reaction stoichiometry in PZH H PZCOO
3.2 mol/l PZ solution at
318 K 4
CO 2
2

δni
0
δ CO 2
2-
-2 CO 3
PZ
-
PZ(COO )2
-4 -
PZCOO [PZ] = 3.2 M
T = 318 k
-6
0.0 0.2 0.4 0.6 0.8 1.0
α CO2 (mol CO 2 / mol PZ)

Fig. 3 Liquid phase 1.0


speciation (xi) in CO2 loaded [PZ] = 0.8 M
Fraction of PZ species in liquid phase

0.8 mol/l aqueous PZ at PZ T = 328 K


328 K 0.8
+ -
+
H PZCOO
PZH
0.6

-
HCO3
0.4
-
PZCOO
-
PZ(COO)2 CO2
0.2
2-
CO3
0.0

0.2 0.4 0.6 0.8 1.0 1.2 1.4


Loading mol (CO2/mol PZ)
Post-combustion Carbon Dioxide Capture with Aqueous … 199

consumed gradually with CO2 loading forming protonated PZ carbamate and PZ


dicarbamate at lower loading. At higher loading, PZ carbamate and PZ dicarbamate
are converted to protonated PZ carbamate and at this concentration most of the PZ
exists in the protonated form (PZH+). This indicates that the activity of PZ is more
in these conditions of CO2 loading. Availability of PZ activity also supports the
action of PZ as a good rate promoter for CO2 reaction.
The heat of absorption is an essential data for predicting the performance of the
solvent as it is a major component of the total energy requirement for solvent
regeneration. The heat of absorption can be estimated from the VLE data using the
well known Gibbs- Helmholtz equation given in Eq. (1):

DHabs d ln pCO2
¼ ð1Þ
R dð1=TÞ

Figure 4 compares experimental heat of absorption of CO2 in 2.4 m (mol/kg


water) aq. PZ at 313 K taking data from Hillard [25] and our own results using the
model predicted equilibrium pCO2 . As expected, heat of absorption of CO2 in aq. PZ
is comparable to that of MEA, which is *80 kJ/mol CO2. Amine volatility is
important information needed for choosing the solvent for CO2 capture application.
Figure 5 compares the experimental and model results of PZ volatility expressed in
partial pressure of PZ over the CO2 loaded solvent.

4.3 VLE of CO2 in Aqueous AMP

Equilibrium concentration of CO2 in aq. AMP and the pCO2 over the solution is also
an important data for the thermodynamic modelling of CO2 + AMP + H2O and
CO2 + AMP + PZ + H2O systems. Many authors have reported this important
VLE data and a good review on this topic can be found in Bougie and Iliuta [26].

Fig. 4 −ΔHabs (kJ/mol CO2) 100


of a 2.4 mol/(kg water) [PZ] = 2.4 m,
aqueous PZ at 313 K T=313 K
80
-ΔHabs (kJ/ mol CO 2)

60

40

20 - model predicted
Experimental results
- Hilliard (2008)
0
0.0 0.2 0.4 0.6 0.8 1.0
αCO2 (mol CO2 / mol PZ)
200 S.K. Dash

Fig. 5 Amine volatility for a 10


-1

2 mol/(kg water) PZ solution


[PZ = 2 m]
at 313 and 343 K loaded with 343 K
CO2 -2
10

pPZ , [kPa]
-3
10

- model predicted 313 K


-4
10 Experimental results
- Hilliard (313 K)
- Hilliard (343 K)
-5
10
0.2 0.4 0.6 0.8 1.0
αCO2 (mol CO2/ mol PZ)

Here, we compare few literature data with our own modelled results. Figure 6
represents the experimental solubility of CO2 in 3 mol/dm3 aq. AMP from
Tontiwachwuthikul et al. [27] and our own work [12]. Figure 7 represents solubility
of CO2 in 5 mol.dm−3AMP from Teng and Mather [28]. Figures 8 and 9 present
the solubility data CO2 over aq. AMP and VLE modelling results at different
temperatures considering the model parameters from Dash [29].
In this case also, the equilibrium pCO2 is low at lower CO2 loading because at
this condition CO2 is almost consumed in the solvent by chemical reaction. But
pCO2 increases at high CO2 loading which can be due to the dominance of physical
absorption. A detail discussion of other properties i.e., model predicted activity
coefficients, liquid phase speciation, reaction stoichiometry, heat of absorption and
amine volatility, can be found in Dash et al. [15] and Tong et al. [28]. The −ΔHabs
of CO2 into aqueous AMP solvent is estimated from the eNRTL model using

Fig. 6 Solubility of CO2 1000


over 3 mol/l AMP at different
temperatures [AMP] =3 M
100
CO2 partial pressure [kPa]

10
Experimental results
Tontlwachwuthikul
1
- 293 K
- 313 K
- 333 K
0.1
- 353 K
- lines
model predicted
0.01
0.2 0.4 0.6 0.8 1.0 1.2
Loading (mol CO2/ mol AMP)
Post-combustion Carbon Dioxide Capture with Aqueous … 201

Fig. 7 Solubility of CO2 10000


over 5 mol/(kg water) AMP at
323 K
1000

Y1 Axis Title
100

10
Experimental
Teng & Mater
5. 0m
1 T = 50 C
- Model

0.1
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3
X Axis Title

Fig. 8 Equilibrium pCO2 over Experimental results


22 mass% aqueous AMP This work,
3
10 T = 298 K
solution
T = 308 K
T = 318 K
2
10 T = 328 K
- (lines)
pCO2, [kPa]

model predicted
1
10

0
10
[AMP] = 22 mass %
-1
10

0.2 0.4 0.6 0.8 1.0 1.2


α CO2 (mol CO 2 /mol AMP)

Fig. 9 Equilibrium pCO2 over Experimental results


30 mass% aqueous AMP T = 298 K
solution 3
T = 308 K
10
T = 318 K
T = 328 K
10
2
- (Lines)
model predicted
pCO2, [kPa]

1
10

0
10

-1
10 [AMP]= 30 mass %

-2
10
0.2 0.4 0.6 0.8 1.0
αCO2 (mol CO2/ mol AMP)
202 S.K. Dash

Eq. (1). It is found to be about 65 kJ/mol CO2 by taking an average value between
313 and 373 K.

4.4 VLE of CO2 in Aqueous (AMP + PZ)

Since this chapter discusses the suitability of AMP + PZ solvent for PCC process,
we focus more on the VLE of CO2 in AMP + PZ solvent and the energy
requirement. The solubility of CO2 in aqueous 2 and 3 mol/dm3 AMP + 0.5, 1.0
and 1.5 mol/dm3 PZ and 25 and 20 wt%, AMP + 5 and 10 wt% PZ have been
presented by Yang et al. [30] and Tong et al. [18], respectively. Both the workers
modelled the VLE using Kent- Eisenberg approach. Tong et al. [18] compared the
results with the solubility of CO2 in 30 wt% AMP and reported that the stoichio-
metric loading of this blend is comparable to that of 30 wt% MEA. The solubility of
CO2 in aq. (3 mol/dm3 AMP + 1.5 mol/dm3 PZ) has been presented by Bruder
et al. [17] after screening some other compositions of AMP and PZ. They found
that this composition has shown the highest cyclic capacity and can be used
commercially without the problem of solid precipitation upon CO2 loading. Dash
et al. [16, 22] presented experimental results of VLE of CO2 in aqueous 30 wt%
(AMP + PZ: 22 + 8, 25 + 5, and 28 + 2), 40 wt% (AMP + PZ: 40 + 0, 32 + 8,
35 + 5, 32 + 8) and 50 wt% (AMP + PZ: 50 + 0, 42 + 8, 45 + 5, 48 + 2) total
amine. The VLE of CO2 in these solvents have been represented by modelling it
using eNRTL equation for the ternary system AMP + H2O + CO2 and
PZ + H2O + CO2 and also by regressing the VLE data of the quaternary system
AMP + PZ + H2O + CO2. Figure 10 compares typical CO2 solubility data at dif-
ferent temperature and the smoothed curves obtained using e-NRTL model of this
work.

Fig. 10 Equilibrium pCO2 Experimental results, this work


over aqueous (25 wt% 10
3
[AMP + PZ] = (25 + 5) mass %
AMP + 5 wt% PZ) at 298– T: 298 K
328 K 2 T: 308 K
10
T: 318 K
T: 328 K
pCO2 /kPa

1
10 (Lines): Model

0
10

-1
10

-2
10
0.2 0.4 0.6 0.8 1.0
αCO2, mole CO 2/mole amines
Post-combustion Carbon Dioxide Capture with Aqueous … 203

It is observed that more PZ in AMP + PZ solvent leads to increase in CO2


loading as the VLE curves exhibit. From these VLE curves the solvent capacity can
be calculated using Eq. (2). The assumptions in this calculation
 are (a) the liquid
phase CO2 loading in the rich solvent outlet stream apCO2 ;rich from the absorber is
in equilibrium with the pCO2 in the inlet gasstream and (b) the liquid phase CO2
loading in the lean solvent stream apCO2 ;lean entering to the absorber is in equi-
librium with the pCO2 in the outlet gas stream.
     
mole CO2 mole ðAMP + PZÞ
capacity ¼ apCO2 ;rich  apCO ;lean ½AMP þ PZ 
kg solvent 2 kg solvent
ð2Þ

The solvent capacity of aq. (35 wt% AMP + 5 wt% PZ) and (25 wt%
AMP + 5 wt% PZ) are 0.902 and 0.623, respectively, at 313 K and at the lean-rich
pCO2 of 1 and 10 kPa respectively, as reported by Dash et al. [16]. They also
reported that increasing PZ concentration in the aqueous AMP solvent and
increasing the total amine concentration in the solvent both increased the solvent
capacity. Higher capacity can be achieved if the pCO2 is more in the flue gas rather
than decreasing the lean loading of the solvent at the outlet of the regenerator. This
behaviour also observed in case of aq. PZ, for example the cyclic capacity of
4.5 mol/dm3 PZ solvent is 40% more than the capacity of 3.2 mol/dm3 PZ when
analyzed at 1 kPa lean pCO2 and 10 kPa rich pCO2 .
The eNRTL model can predict the liquid phase concentration of various species
formed by CO2 reaction such as protonated AMP (AMP+), Protonated PZ,
PZCOO−, PZ (COO−)2, H+PZCOO−, carbonate (CO32−) and bicarbonate (HCO3−)
ions. This speciation is presented in Figs. 11 and 12 for 22 wt% AMP + 8 wt% PZ
and 35 wt% AMP + 5 wt% PZ, respectively, at 313 K. It is evident that there are

Fig. 11 Model predicted 4.0


speciation in CO2 loaded
aqueous (22 wt% AMP + 8 3.5
-3

wt% PZ) at 318 K


Constration of species, mole.dm

-
3.0 HCO3
AMP AMPH
+

2.5
PZ
2.0
-
PZCOO +
1.5 - PZH CO2
PZ(COO )2 2-
CO 3
1.0 + -
H PZCOO
0.5 -
OH
0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
αCO2 , mole CO2/ mole amines
204 S.K. Dash

Fig. 12 Liquid phase PZ 0.6


species of a CO2 loaded AMPH
+
AMP
aqueous (35 wt% AMP + 5

-3
0.5 2-

Concentration of species, kmol.m


wt% PZ) at 313 K CO3

0.4 -
PZ HCO3
+ -
- H PZCOO
0.3 PZ(COO )2

0.2
-
PZCOO +
0.1 - PZH
OH
CO2
0.0
0.0 0.2 0.4 0.6 0.8 1.0
αCO 2(mol CO2/ mol amine)

differences between the speciation of aq. (AMP + PZ) and aq. AMP or aq. PZ. Both
AMPH+ and PZH+ are the predominant products in the carbonated solution of
AMP + PZ favouring rapid CO2 absorption. The fast formation of AMPH+ and its
concentration delays the protonation of PZ towards higher CO2 loading making
more PZ activity and reactive PZ available in the solution. This characteristics
favours fast kinetics of PZ-CO2 reaction and thus make the solvent suitable for CO2
capture.
As discussed earlier, the heat of absorption of CO2 in AMP + PZ solvent is an
important information for the estimation of the energy demand in the regenerator.
Using a differential reaction calorimeter, Xie et al. [31] measured the heat of CO2
absorption in 3 mol/dm3 AMP + 0.5 mol/dm3 PZ, 4 mol/dm3 AMP + 1 mol/dm3
PZ and 3 mol/dm3 AMP + 1.5 mol/dm3 PZ at 313–373 K. It is reported by them
that heat of absorption is high and it slowly decreases in the low CO2 loading
region. This is due to the fact that at the low CO2 loading region, PZ is converted to
PZ carbamate which contributes about 30.07 kJ/mol energy as heat of reaction.
Their finding revealed that ΔHabs of the three different blends of AMP + PZ are
comparable and at low loading this value is about 75 kJ/mol. Dash et al. [14]
reported the eNRTL model predicted average ΔHabs between 313 and 343 K. It is
about 70 kJ/mol of CO2 for aqueous (AMP + PZ) solvent which is about 7% lower
than the value found out experimentally by Xie et al. [31]. The low heat of
absorption value qualifies this AMP + PZ solvent as an energy efficient one.
The information on amine volatility expressed as vapour pressure of AMP at the
absorber outlet temperature is needed to quantify the amount of amine loss with the
outgoing flue gas. As excessive volatility of the solvent should be avoided to reduce
amine loss to environment, a water wash packed section should be provided at the
top portion of the column. Nguyen et al. [32] presented the amine volatility of 5 m
AMP solvent at 313–333 K as data on volatility is important at the conditions of the
outlet gas from the absorber. They reported that AMP has a slight higher volatility
Post-combustion Carbon Dioxide Capture with Aqueous … 205

of 112 ppm at this condition of temperature and pCO2 as AMP does not form
non-volatile carbamates in the carbonated solution. Dash et al. [16] showed that the
amine volatility of a mixed solvent of aqueous (AMP + PZ) loaded with CO2 is less
than that of CO2 loaded single amine solvent i.e., aq. AMP or aq. PZ. Here, addition
of PZ in AMP shows a positive effect of lowering the volatility of both the amines
due to highly non ideal phase of the carbonated solution. For a detailed discussion
of non-ideality and activity coefficients, amine volatility, heat of absorption, and
other properties of CO2-(AMP + PZ) system Dash et al. [16] may be referred.
Khakharia et al. [33] reported amine emission to environment for a AMP + PZ
CO2 capture process. Their experimental study reviled that the amount of amine
emission to environment depends on lean solvent temperature. Amine emission is
reduced with increasing lean solvent temperature and aerosol emissions are found
only when lean solvent have high pH value. The measured emission of AMP is in
the range of 1500–3000 mg/Nm3 and PZ in the range of 200–400 mg/Nm3. This
emission can be controlled by providing a water wash section in the absorber that
fall within the environmental regulations.

5 Absorption of CO2 into AMP + PZ Solvent

One of the most essential data for CO2 absorption in amine solvents is the kinetic
data. Both high equilibrium capacity and high absorption rate are desirable for the
CO2 capture solvent. For AMP and AMP + PZ solvents, knowledge about reaction
mechanism and kinetic constants are important. The CO2 absorption rates into
various compositions of aq. (AMP + PZ) have been measured by various workers
[9, 13, 34] at different temperatures and at different CO2 partial pressures using
wetted wall column. For the purpose of calculation of kinetic parameters using the
coupled mass transfer kinetics model, they considered the following reactions of
CO2 with aq. (AMP + PZ) represented in reactions R1–R11.

K1 ;k21
CO2 þ AMP þ H2 O $ AMPH þ þ HCO
3 ðR:1Þ

K2 ;k22
CO2 þ PZ þ H2 O $ PZCOO þ H3 O þ ðR:2Þ

K3 ;k23
CO2 þ AMP þ PZ $ PZCOO þ AMPH þ ðR:3Þ

K4 ;k24
CO2 þ PZCOO þ H2 O $ PZ(COO Þ2 þ H3 O þ ðR:4Þ

K5 ;k25
CO2 þ AMP þ PZCOO $ PZ(COO Þ2 þ AMPH þ ðR:5Þ
206 S.K. Dash

K6 ;k26
CO2 þ 2H2 O $ HCO
3 þ H3 O
þ
ðR:6Þ

K7
HCO
3 þ H2 O $ CO3 þ H3 O
2 þ
ðR:7Þ

K8
PZ þ H3 O þ $ PZH þ þ H2 O ðR:8Þ

K9
PZCOO þ H3 O þ $ H þ PZCOO þ H2 O ðR:9Þ
K10
AMPH þ þ H2 O $ AMP þ H3 O þ ðR:10Þ

K11
2H2 O $ H3 O þ þ OH ðR:11Þ

Samanta and Bandyopadhyay [34] reported the values of kinetic parameters (k23
and k25) for the CO2 reaction with aq. total 30 wt% (AMP + PZ) in the range of
298–313 K. These values are presented in Eqs. (R.3) and (R.4).
  
7:2  104 1 1
k23 ¼ 3:516  10 exp 
4
 ð3Þ
R T 298
  
7:6  104 1 1
k25 ¼ 1:836  10 exp 
4
 ð4Þ
R T 298

Dash et al. [2] worked with more concentrated solvents having total 40 wt%
(AMP + PZ) and total 50 wt% (AMP + PZ) keeping PZ concentration within the
range of 2–8 wt%. They reported the absorption measurements at 303–333 K and
pCO2 in the range of 4.9–15.1 kPa. In that work, for the first time they incorporated
the eNRTL model to estimate the initial liquid phase concentrations for the coupled
mass transfer- kinetics model to predict the kinetic parameters. The kinetic
parameters obtained for total 50 mass% (AMP + PZ) are given by Eqs. (R.5) and
(R.6).
  
65225 1 1
k23 ¼ 3:516  104 exp   ð5Þ
R T 298
  
70550 1 1
k25 ¼ 1:836  104 exp   ð6Þ
R T 298

The Arrhenius plots of these kinetic parameters are represented in Fig. 13. The
effect of temperature on specific rate of absorption (RA) of CO2 into aq. (45 wt%
AMP + 5 wt% PZ) at different pCO2 have been presented in Fig. 5.13. It is evident
that the rate of CO2 absorption increases with the increase of temperature and pCO2 .
Considering the experimental results of CO2 absorption rate in aq. (45 wt%
AMP + 5 wt% PZ) solvent from Dash et al. [2], we have analysed the variation of
Post-combustion Carbon Dioxide Capture with Aqueous … 207

8
Fig. 13 Arrhenius plot for 10
the reaction rate constant, eq. (5.55) 6
10
Kpzcoo–(k23) and kPZ(COO–) eq. (5.56)
2 for CO2– 10
7

k2PZ(COO )2(m6.kmol-2.s-1)
kPZCOO (m6.kmol-2.s-1)
(AMP + PZ + H2O) 50 wt%
(AMP + PZ)
6
10
5
10

-
5
10

4 4
10 10
2.9 3.0 3.1 3.2 3.3
1000/T(K)

rate of absorption with temperature and the trend of enhancement factor with CO2
loading. Figure 14 shows the variations of the measured rates of absorption with
pCO2 at different temperatures. It shows that the rates of absorption increase with
pCO2 . Similarly, Fig. 15 shows the variation of the measured enhancement factor

Fig. 14 Effect of temperature 35


on specific rate of absorption
(RA) of CO2 into aq. (45 wt% 30
AMP +5 wt% PZ) at different pCO2=15 kPa
RAx106(kmol.m-2.s-1)

CO2 partial pressure


25
pCO2=10 kPa

20

pCO2= 5 kPa
15

10
320 324 328 332 336
T(K)

Fig. 15 Effect of partial


pressure on specific rate of 35
CO2 absorption into aqueous
323 K
(45 wt% AMP + 5 wt% PZ) 30 328 K
at different temperatures 333 K
RAx10 (kmol.m .s )
-1
-2

25

20
6

15

10
4 6 8 10 12 14 16
pCO2(kPa)
208 S.K. Dash

Fig. 16 Effect of CO2 420


loading on the enhancement 323 K
of CO2 absorption rate into 410
328 K
aqueous solution of 45 wt% 333 K
400
AMP + 5 wt% PZ at different
temperatures 390

ECO2
380

370

360

350
0.02 0.03 0.04 0.05
αCO2

ðECO2 Þ with CO2 loading. It shows that the enhancement factor ðECO2 Þ sharply
decreases with increase in the loading of CO2 until a loading of about 0.035.
Thereafter, enhancement factor remains essentially constant (Fig. 16).
All these findings indicate that rapid reaction occur at low CO2 loading corre-
sponding to low pCO2 indicating the suitability of the AMP + PZ solvent for CO2
capture from flue gas.

6 Amine Degradation and Corrosion

Usually CO2 capture plants operate at near ambient to high temperature. The
regenerator operates at elevated temperature of about 380–400 K. At this temper-
ature the solvent may degrade producing components of its homologous series.
Another difficulty in flue gas treating is the fact that it contains O2 which may result
in oxidizing the solvent species and accelerating the production of degradation
products. Thus, high a temperature and presence of O2 environment both lead to
higher solvent degradation. These amine degradation products are sometimes
responsible for corrosion in the stripper, reboiler and the piping. Again, corrosion
may be there due to carbonated solution and it may be vary localized where tem-
perature is high and concentration of O2 is relatively more. Corrosion and corrosion
inhibition of aqueous AMP has been studied by Veawab et al. [35, 36] using static
weight loss method. They reported that aq. solutions of AMP were more corrosion
resistant to carbon steel as compared to aq. MEA solution under similar environ-
ment of both pure CO2 and a mixture of CO2 + N2 + O2 having about 10% O2 in it.
Freeman et al. [37] reported that both AMP and PZ are more stable than other
alkanolamines. They also reported that both AMP and PZ are slower in oxidative
degradation than MEA and much more resistant to thermal degradation than MEA.
Although AMP is relatively stable, it is susceptible to oxazolidinone formation in
concentrated O2 environment. Due to presence of NOx in flue gas, it may act as a
source of nitrosamine formation in the degraded products. A complete degradation
Post-combustion Carbon Dioxide Capture with Aqueous … 209

study of AMP and PZ based solvents, the degradation mechanism and effect of
degradation products, can be found in Wang and Jens [38, 39], Mazari et al. [40]. In
his report, Wang and Jens [39] mentioned that mononitrosopiperazine may be
present in the degraded product of AMP + PZ solvent which may emit to the
environment but can be avoided by adding water wash section. All these study
reveals that overall, with respect to corrosion, thermal and oxidative degradation,
AMP + PZ solvent yet have an advantage over the bench marked MEA.

7 Energy Demand Analysis from Pilot Plant Study


and Process Simulation

As discussed earlier in the Sect. 5.2, there are many vendors developing com-
mercial solvents for PCC processes. Few examples are described in the energy
analysis section and a detailed review of the energy requirements for the com-
mercial solvents developed by power generation companies can be found from
Budzianowski [5]. A comparison of energy requirement for MEA and AMP + PZ
process is described here. Mangalapally and Hasse [41] performed pilot-plant study
of CO2 absorption using both MEA and AMP + PZ blended solvents. They found
that for the gas fired power plant case (pCO2 = 5.5 kPa in their flue gas), the energy
demand for MEA in the pilot-plant was around 7.2 GJ/t CO2. For coal fired power
plant case (pCO2 = 11.20 kPa), it was around 5.5 GJ/t CO2 to reach 90% removal,
which is relatively high compared to the standard energy demand of around 4 GJ/t
CO2. Notz et al. [42] conducted systematic pilot-plant study of CO2 absorption
from natural gas-fueled power plant using MEA absorption process. They reported
that for a loading range of 0.096–0.36 mol CO2/mol MEA, the specific energy
demand was on the lower side of 3.7 GJ/ton of CO2 and on the higher side of 10.2
GJ/ton of CO2.
In order to compare the performance of different solvents, it is necessary to
perform pilot plant study and process simulation on a consistent basis and perform a
process analysis of the system. Concentrated aqueous blends of AMP and PZ have
been proposed to have improved rate kinetics and batter capacity solvent for CO2
capture with lower energy requirement. The regeneration energy studies have been
presented. The simplified PCC process using amine absorption is presented in
Fig. 17.
In this process, the treaded flue gas (cooled and desulfurized) is fed to an
absorption column where gas phase CO2 reacts with the aq. amine solvent and
transferred to solvent phase. This CO2-rich solvent is sent to the regenerator column
where heat is used to strip the CO2 from the solvent.
The process flow sheet described in Fig. 17 is a simple one which has been
practised for several years. But, for flue gas treating, alternative process configu-
rations should be considered to lower the energy requirement. Performance of
various alternative process configurations have been studied by Karim et al. [43,
210 S.K. Dash

CO2OUT
GASOUT

LEANIN

ABSORBER RICHIN STRIPPER

GASIN
LRHX

RICHOUT
LEANOUT

Fig. 17 Simplified flow sheet of absorption-regeneration unit

44]. Based on their simulation study they suggested a number of process alterna-
tives and conditions to be followed. One such consideration is temperature profile
of absorber and regenerator. Since CO2 absorption in aqueous amines is an
exothermic process, it increases the temperature of the liquid phase and reduces the
driving force. On the other hand increased temperature also increases the rate of
reaction. Hence, there should be a trade-off between this two effects and the opti-
mum temperature profile should be adopted. One configuration in view of energy
requirement is the split flow process. In this process a fraction of the partially lean
and partially rich solvent from the middle of the regenerator and absorber column
respectively, can be withdrawn and reconnected at the bottom of the respective
columns. Cousins et al. [45] studied these alternatives i.e., inter stage cooling for
temperature control and the split flow process and found that there could be increase
in capture efficiency. The other alternative arrangements are rich split and vapour
recompression in the regenerator and heat integration. In rich split configuration,
vapour liberated from the hot rich solvent transfers heat to the cold rich solvent
entering to the regenerator. In vapour recompression, striping steam is removed
from a suitable location of the regenerator, compressed and reintroduced into the
regenerator to supply additional heat. The recompression of the vapour is done by
applying mechanical energy which is utilized to provide additional stripping steam.
Vapour recompression is often suggested for reducing regeneration energy
requirement but it is achieved by providing additional energy for compression. The
extensions of vapour recompression are multi-pressure stripping and matrix strip-
ping reported by Cousins et al. [45]. Heat integration aims at utilizing some of the
waste heat from the PCC process and hence reducing the heat loss.
The pinch analysis which analyzes approach to equilibrium in the absorber and
stripper determines the rich and lean loadings and the capacity of the solvent which
also helps energy minimization and efficient plant operation. The rich loading
specifically determines the minimum vapor rate at the top of the stripper. There are
Post-combustion Carbon Dioxide Capture with Aqueous … 211

important trade-offs of absorber and stripper packing height, solvent rate, and steam
rate. Any change in contactor, solvent, process configuration etc., should be fol-
lowed by a careful optimization of the rich and lean loadings. Cousins et al. [46]
and Karimi et al. [43] compared five alternative stripper configurations with respect
to energy consumption and capital investment. Process simulation tools UniSim
and ProTreat were used by them to investigate the alternative stripper configura-
tions. Of the five alternative stripper configurations considered by Karimi et al. [43],
the vapor compression stripper configuration was found to be the best by them. The
vapour recompression technique can be adopted for this AMP + PZ process to gain
some energy advantage.

8 Regeneration Energy Requirement for AMP + PZ


Solvent

Prior to commercialization of CO2 capture process, several pilot plant studies are
necessary to gain experience in this process. Mangalapally and Hasse [41] pre-
sented pilot plant study of CO2 capture process. They used AMP + PZ mixed
solvent with an objective of commercialization of the AMP + PZ process. They
reported that AMP + PZ is a suitable solvent for CO2 capture from fossil fuel based
power plants. Based on their pilot plant study, they also found, CESAR-1 solvent
28 wt% AMP + 17 wt% PZ has an expected reduction of about 20% in the
regeneration energy demand as compared to that of MEA solvent. Also, AMP + PZ
solvent can reduce about 45% in the solvent circulation rate over those of the MEA
process. Artanto et al. [46] conducted pilot plant study of CO2 capture from flue gas
using 25 wt% AMP + 5 wt% PZ solvent and compared its performance with that of
30 wt% MEA solvent. They reported that AMP + PZ solvent shows better per-
formance over MEA when the pilot plant study is conducted for the same %CO2
removal and at the same L/G ratio. Their modelling work also revealed that the
regeneration energy could be lower by using concentrated solvent since the cyclic
capacity of the solvent would be higher. They found that the condenser heat duty
and the sensible heat duty are the major contributions to the total energy demand.
The energy efficiency of the aq. (18 wt% AMP + 17 wt% PZ) as a solvent for
CO2 absorption by chemical process has been studied by Dash et al. [47].
A parametric study of this process based on regenerative CO2 absorption with
AMP + PZ solvent has been done by them using rate based model. They compared
their simulation results with the reported pilot plant data for CO2 capture using aq.
AMP and aq. (AMP + PZ) [41] and reported good agreements. From the modeling
and simulation work of Dash et al. [47], It has been found that the optimum
regeneration temperature of AMP + PZ solvent is about 393–398 K at the bottom
and 381–384 K at the top of the regenerator column. As the reboiler heat duty
represents the major portion of the total operating cost in a PCC plant, they ana-
lyzed the reboiler heat duty by splitting it into three components viz., energy to be
212 S.K. Dash

supplied (i) for providing the heat of absorption or heat of desorption of CO2, (ii) to
generate additional stripping steam, and (iii) for rich solvent heating and to provide
the sensible heat. By making an energy balance as described by Artanto et al. [46],
they estimated all these quantities.
These findings of the individual energy components are illustrated in Figs. 18
and 19 with respect to various %CO2 capture, rich loading at the absorber and lean
loading at the regenerator, respectively. To have a sensitivity analysis on the lean
loading and the reboilor duty, it is found that with the increase of %CO2 capture,
the reboiler duty increases as well. Due to higher sensible heat requirement, there is
a steep increase in reboilor duty once the capture approaches 95%. It is estimated
that the total energy demand is 3.9 GJ/ton of CO2 at an L/G of 2.75 and a lean
loading of 0.129. However, the energy demand is 3.75 GJ/ton-CO2 at an L/G of 2.9
and a lean loading of 0.155 for the same conditions. Hence, the energy demand is
less when the lean loading is lowered to achieve the desired %CO2 capture rather
than by increasing the L/G ratio.

Fig. 18 Effect of % CO2 4600


capture on reboiler duty 0.50
4400 Rich loading
(kJ/kg CO2) for a constant 0.45

Loading (mol CO2/mol amine)


L/G with a temperature
Reboiler duty (kJ/kg-CO 2)

4200 0.40
approach of 5 K in the
(18 wt.% AMP + 17.5 wt.% PZ)
lean-rich heat exchanger. 4000 Column ht.= 10 m 0.35
(Reproduced from Dash et al.
L/G= 2.9 0.30
[47] with permission) 3800
0.25
3600
0.20
Lean loading
3400 0.15

3200 0.10
0.84 0.86 0.88 0.90 0.92 0.94 0.96
%-CO2 capture

Fig. 19 Effect of CO2 4200 3.0


capture on reboiler duty L/G
(kJ/kg CO2) for a constant 2.5
lean loading with a 4100
Reboiler duty (kJ/kg-CO 2)

temperature approach of 5 K
(18 wt.% AMP + 17.5 wt%. PZ) 2.0
in the lean-rich heat 4000 Lean loading = 0.12
exchanger. Reproduced from
Column ht.= 10 m 1.5
Dash et al. [47] with
permission 3900
1.0

3800
0.5
Rich loading
3700 0.0
84 86 88 90 92 94 96
%-CO2 capture
Post-combustion Carbon Dioxide Capture with Aqueous … 213

It is found that for 90% CO2 removal, about 1.56 GJ/tCO2 is required towards
heat of desorption of CO2. The heat of desorption for AMP + PZ solvent is
independent of CO2 loading up to a value of about 0.5 mol CO2/mol amine. About
1.45 GJ/tCO2 goes towards generation of stripping steam which again increases
with the requirement of increase in % CO2 capture. Only about 0.64 GJ/tCO2 is
needed to meet the sensible heat requirement, which decreases to some extent with
the increase in stripping steam. The total energy requirement for the AMP + PZ
solvent is *3.65 GJ/tCO2 and *3.2 for 90 and 85% CO2 capture respectively.
This range is nearly equal to the value reported by Mangalapally and Hasse [41],
but slightly lower than the value obtained by Artanto et al. [46] from their pilot
plant study. The reason could be Artanto et al. [46] used lower concentration of
25 wt% AMP + 5 wt% PZ solvent in their pilot plant study and found the energy
demand to be *4 GJ/tCO2. This value could be still lower if they would have used
higher amine concentration such as, 20% AMP + 15% PZ. On the contrary Khan
et al. [3] reported that the regeneration energy demand increases with the increase in
PZ concentration as PZ has a heat of absorption *80 kJ/mol of CO2. They also
found that the energy demand for the 30 wt% total AMP + PZ solvent is in the
range of 3.4–4.2 GJ/t CO2 whereas for 30 wt% MEA it is 3.6–4.5 GJ/tCO2 when
both the solvents tested in the same bench scale pilot plant. Spek et al. [48] con-
ducted simulation study of a full scale CO2 capture plant and reported a regener-
ation heat duty of 2.9 GJ/tCO2 for AMP + PZ solvent as compared to 3.6 GJ/tCO2
for MEA solvent. They also verified that all the desirable properties of AMP + PZ
solvent are technical performance indicators for the capture plant.
As PZ has a solubility limit in aq. AMP, the final solvent composition should be
selected based on solvent circulation and the energy requirement as well as the
maximum solid (PZ) solubility so that, there are no precipitate formation in the
colder parts of the plant and equipments. This higher solubility limit could be 15 wt
% of PZ as reported by Bruder et al. [17], Dash et al. [47]. Based on all these
findings, we suggest that the optimum solvent concentration should be 20–25 wt%
AMP + 10–15 wt% PZ) for an energy efficient CO2 capture process.

9 Conclusions

In this study, analysis of various important properties of CO2 capture solvent


AMP + PZ has been presented. From the critical analysis of the present study and
comprehensive literature review on this CO2 capture process, it is confirmed that
this solvent has all the desirable properties such as moderate density and viscosity,
relatively fast kinetics, high equilibrium solubility and better regeneration charac-
teristics, all of which lead to energy efficient CO2 capture. Besides, energy effi-
ciency is also related to the other essential properties of this AMP + PZ solvent
system such as low or moderate corrosiveness, higher degradation resistance, lower
undesirable environmental emissions and relatively lower regeneration energy
requirement. It is evident from the forgoing discussion that this solvent is suitable
214 S.K. Dash

for CO2 capture from low partial pressures gas streams. From the various pilot plant
tests and the simulation data, the regeneration energy requirement been found to be
*2.9–3.7 GJ/tCO2 for a solvent composition of aq. (18–25 wt% AMP + 10–15 wt
% PZ) for *85–90% CO2 capture. This energy requirement is about 20–30% less
than the energy requirement of the MEA process.
Although pilot plant study and simulation work provide good information on the
heat duty, it is desirable to check the energy demand of this PCC process with large‐
scale pilot plant test and demonstration plant test. The amount of actual energy
saving in various process configurations such as vapour recompression should be
confirmed by tests in a demonstration plant with efficient random or structured
packing in the column with applicable column stage efficiencies. A comprehensive
study of combination of one model for the power plant including coal combustion,
steam cycle etc., with the another one for CO2 capture should be useful for esti-
mation of total energy demand as well as for cost optimization utilizing the low
pressure steam of the power to supply the required energy for the regenerator.

References

1. Dash SK, Bandyopadhyay SS (2016) Studies on the effect of addition of piperazine and
sulfolane into aqueous solution of N-methyldiethanolamine for CO2 capture and VLE
modelling using enrtl equation. Int J Greenhouse Gas Control 44:227–237
2. Dash SK, Bandyopadhyay SS (2013) Carbon dioxide capture: absorption of carbon dioxide in
piperazine activated concentrated aqueous 2-amino-2-methyl-1-propanol. J Clean Energy
Technol 1(3):184–188
3. Khan AA, Haldera GN, Saha AK (2016) Experimental investigation of sorption character-
istics of capturing carbon dioxide into piperazine activated aqueous
2-amino-2-methyl-1-propanol solution in a packed column. Int J Greenhouse Gas Control
44:217–226
4. Hairul NAH, Shariff AM, Bustam MA (2016) Mass transfer performance of
2-amino-2-methyl-1-propanol and piperazine promoted 2-amino-2-methyl-1-propanol
blended solvent in high pressure CO2 absorption. Int J Greenhouse Gas Control 49:121–127
5. Budzianowski WM (2015) Single solvents, solvent blends, and advanced solvent systems in
CO2 capture by absorption: a review. Int. J. Global Warming 7(2)
6. Svendsen, H (2010) CO2 capture by solvents; possibilities and challenges. In: Proceedings of
post-combustion CO2 capture workshop. Tufts European Center Talloires, France, 11–13 July
2010
7. Kishimoto S, Hirata T, Iijima M, Ohishi T, Higaki K, Mitchell R (2009) Current status of
MHI’s CO2 recovery technology and optimization of CO2 recovery plant with a PC fired
power plant. Energy Procedia 1(1):1091–1098
8. Knudsen JN, Jensen JN, Vilhelmsen PJ, Biede O (2009) Experience with CO2 capture from
coal flue gas in pilot-scale: testing of different amine solvents. Energy Procedia 1:783–790
9. Jockenhoevel T, Schneider R, Rode H (2010) Validation of a second-generation
post-combustion capture technology—results from POSTCAP pilot plant operation. In:
Powergen Europe, 8–10 June 2010
10. Paul S, Mandal B (2006) Density and viscosity of aqueous solutions of
(N-Methyldiethanolamine + Piperazine) and (2-Amino-2-methyl-1-propanol + Piperazine)
from (288 to 333) K. J Chem Eng Data 51:808–1810
Post-combustion Carbon Dioxide Capture with Aqueous … 215

11. Sun WC, Yong CB, Li MH (2005) Kinetics of absorption of carbon dioxide into mixed
aqueous solutions of 2-amino-2-methyl-1-propanol and piperazine. Chem Eng Sci 60:503–
516
12. Samanta A, Bandhyopadhyay SS (2006) Density and viscosity of aqueous solutions of
piperazine and (2-Amino-2-methyl-1-propanol + Piperazine) from 298 to 333 K. J Chem Eng
Data 51:467–470
13. Dash SK, Samanta A, Samanta AN, Bandyopadhyay SS (2011) Absorption of carbon dioxide
in piperazine activated concentrated aqueous 2-amino-2-methyl-1-propanol solvent. Chem
Eng Sci 66:3223–3233
14. Dash SK, Samanta A, Samanta AN, Bandyopadhyay SS (2011) Vapour liquid equilibria of
carbon dioxide in dilute and concentrated aqueous solutions of piperazine at low to high
pressure. Fluid Phase Equilib 300:145–154
15. Dash SK, Samanta AN, Bandyopadhyay SS (2011) (Vapour + liquid) equilibria (VLE) of
CO2 in aqueous solutions of 2-amino-2-methyl-1-propanol: New data and modelling using
enrtl-equation. J Chem Thermodyn 43:1278–1285
16. Dash SK, Samanta AN, Bandyopadhyay SS (2012) Experimental and theoretical investigation
of solubility of carbon dioxide in concentrated aqueous solution of
2-amino-2-methyl-1-propanol and piperazine. J Chem Thermodyn 51:120–125
17. Bruder P, Grimstvedt A, Mejdell T, Svendsen HF (2011) CO2 capture into aqueous solutions
of piperazine activated 2-amino-2-methyl-1-peopanol. Chem Eng Sci 66:6193–6198
18. Tong D, Maitland GC, Trusler M, Fennell PS (2013) Solubility of carbon dioxide in aqueous
blends of 2-amino-2-methyl-1-propanol and piperazine. Chem Eng Sci. http://dx.doi.org/10.
1016.ces.2013.05.034
19. Hartono A, Saeed M, Ciftja AF, Svendsen HF (2013) Binary and ternary VLE of the
2-amino-2-methyl-1-propanol (AMP)/piperazine (Pz)/water system. Chem Eng Sci 91:151–
161
20. Haghtalab A, Eghbali H, Shojaeian A (2014) Experiment and modelling solubility of CO2 in
aqueous solutions of Diisopropanolamine + 2-amino-2-methyl-1-propanol + Piperazine at
high pressures. J Chem Thermodyn 71:71–83
21. Halim HNA, Shariff AM, Bustam MA (2015) High pressure CO2 absorption from natural gas
using piperazine promoted 2-amino-2-methyl-1-propanol in a packed absorption column. Sep
Purif Technol 152:87–93
22. Dash SK, Samanta AN, Bandyopadhyay SS (2011) Solubility of carbon dioxide in aqueous
solution of 2-amino-2-methyl-1-propanol and piperazine. Fluid Phase Equilib 307:166–174
23. Bishnoi S, Rochelle GT (2000) Absorption of carbon dioxide into aqueous piperazine:
reaction kinetics, mass transfer and solubility. Chem Eng Sci 55:5531–5543
24. Derks PWJ, Dijkstra HBS, Hogendoorn JA, Versteeg GF (2005) Solubility of carbon dioxide
in aqueous piperazine solutions. AIChE J 51:2311–2327
25. Hillard M (2008) A predictive thermodynamics model for an aqueous blend of potassium
carbonate, piperazine, and monoethanolamine for carbon dioxide. Ph.D. Thesis, The
University of Texas at Austin
26. Bougie F, Iliuta MC (2012) Strerically hindered amine-based absorbents for the removal of
CO2 from gas streams. J Chem Eng Eng Data 57:635–669
27. Tontiwachwuthikul P, Meisen A, choon JL (1991) Solubility of carbon dioxide in
2-amino-2-methyl-1-propanol solutions. J Chem Eng Data 36:130–133
28. Teng TT, Mather AE (1989) Solubility of H2S, CO2 and their mixtures in an AMP solution.
Can J Chem Eng 67:846–850
29. Dash SK (2012) Carbon dioxide capture by Absorption in piperazine activated
2-Amino-2-methyl-1-propanol solvent. Ph.D. Thesis. Indian Institute of Technology,
Kharagpur
30. Yang ZY, Soriano AN, Caparanga AR, Li MH (2010) Equilibrium solubility of carbon
dioxide in (2-amino-2-methyl-1-peopanol + piperazine + water). J Chem Thermodyn
42:659–665
216 S.K. Dash

31. Xie Q, Aroonwilas A, Veawab A (2013) Measurement of Heat of CO2 Absorption into
2-Amino-2-methyl-1-propanol (AMP)/piperazine (PZ) blends using differential reaction
calorimeter. Energy Procedia 37:826–833
32. Nguyen T, Hilliard M, Rochelle GT (2010) Amine Volatility in CO2 capture. Int J
Greenhouse Gas Control 4:707–715
33. Khakharia P, Brachert L, Mertensc J, Anderlohr C, Huizinga A, Fernandez ES, Schallert B,
Schaber K, Vlugt TJH, Goetheer E (2016) nderstanding aerosol based emissions in a post
combustion CO2 capture process: Parameter testing and mechanisms. Int J Greenhouse Gas
Control 34:63–74
34. Samanta A, Bandyopadhyay SS (2009) Absorption of carbon dioxide into aqueous solutions
of piperazine activated 2-amino-2-methyl-1-propanol. Chem Eng Sci 64:1185–1194
35. Veawab A, Tontiwachwuthikul P, Bhole SD (1996) Corrosivity in
2-amino-2-methyl-1-peopanol (AMP)-CO2 system. Chem Eng Commun 144:65–71
36. Veawab A, Tontiwachwuthikul P, Bhole SD (1997) Studies of corrosion and corrosion
control in a CO2-2-amino-2-methyl-1-peopanol (AMP) environment. Ind Eng Chem Res
36:264–269
37. Freeman SA, Dugas R, Wagener DV, Nguyen T, Rochelle GT (2010) Carbon dioxide capture
with concentrated, aqueous piperazine. Int J Greenhouse Gas Control 4:119–124
38. Wang T, Jens K (2012) Oxidative degradation of aqueous 2-Amino-2-methyl-1-propanol
solvent for post combustion CO2 capture. Ind Eng Chem Res 51:6529–6536
39. Wang T, Jens K-J (2014) Oxidative degradation of aqueous PZ solution and AMP/PZ for
post-combustion carbon dioxide capture. Int J Greenhouse Gas Control 24:98–105
40. Mazari SA, Ali B, Jan BM, Saeed IM (2014) Degradation study of piperazine, its blends and
structural analogs for CO2 capture: a review. Int J Greenhouse Gas Control 31:214–228
41. Mangalapally HP, Hasse H (2011) Pilot plant study of two new solvents for post combustion
carbon dioxide capture by reactive absorption and comparison to monoethanolamine. Chem
Eng Sci 66:5512–5522
42. Notz R, Mangalapally HP, Hasse H (2012) Post combustion CO2 capture by reactive
absorption: pilot plant description and results of systematic studies with MEA. Int J
Greenhouse Gas Control 6:84–112
43. Karimi M, Hillestad M, Svendsen HF (2011) Capital costs and energy considerations of
different alternative stripper configurations for post combustion CO2 capture. Chem Eng Res
Des 89:1229–1236
44. Karimi M, Hillestad M, Svendsen HF (2012) Investigation of the dynamic behavior of
different stripper configurations for post-combustion CO2 capture. Int J Greenhouse Gas
Control 7:230–239
45. Cousins A, Wardhaugh L, Feron P (2011) Preliminary analysis of process flow sheet
modifications for energy efficient CO2 capture from flue gases using chemical absorption.
Chem Eng Res Des 89:1237–1251
46. Artanto Y, Jansen J, Pearson P, Puxty G, Cottrell A, Meuleman E, Feron P (2014) Pilot-scale
evaluation of AMP/PZ to capture CO2 from flue gas of an Australian brown coal-fired power
station. Int J Greenhouse Gas Control 20:189–195
47. Dash SK, Samanta AN, Bandyopadhyay SS (2014) Simulation and parametric study of the
post combustion CO2 capture using aqueous 2-amino-2-methyl-1-propanol and piperazine.
Int J Greenhouse Gas Control 21:130–139
48. van der Spek M, Arendsen R, Ramirez A, Faaij A (2016) Model development and process
simulation of postcombustion carbon capture technology with aqueous AMP/PZ solvent. Int J
Greenhouse Gas Control 47:176–199
Energy Efficient Absorbents for Industry
Promising Carbon Dioxide Capture

Y.S. Yu, T.T. Zhang and Z.X. Zhang

Abstract There are growing concerns that carbon dioxide (CO2) emissions are
contributing to global climate change. CO2 capture and sequestration system is an
effective way to alleviate this phenomenon. Chemical absorption of CO2 is a mature
and efficient way to capture CO2 from industrial flue gas. Amines are the most
commonly discussed solvent, such as NH3 as inorganic solvent and mono-
ethanolamine (MEA) as the typical alkanolamine. MEA aqueous solutions have
been widely analyzed and achieved good performance. However, the energy cost
for the amine regeneration remains great which barricades its widely industrial
application. Several energy efficient absorbents are currently discussed from the
perspective of the reaction kinetics, desorption efficiency and sensible heat con-
sumption. These discussions show that the absorbents with higher reaction rate and
mass transfer coefficient and advanced process can reduce the sensible heat (less
consumption of absorbent) and reaction heat, which thus make them as the energy
efficient absorbents.

Nomenclature
C Concentration of the amine, kmol/m3
D, Dm Diffusion coefficient, m2/s
G Inert gas flow rate, kmol/m2/s
H Henry’s constant, MPa m3/kmol
HS Henry’s constant of the solvent, MPa m3/kmol
h Van Krevelen coefficient, m3/kmol
I Ionic strength of the solution, kmol/m3
k Reaction rate constant, m3/kmol/s
km Reaction kinetics, dimensionless

Y.S. Yu  T.T. Zhang  Z.X. Zhang


School of Chemical Engineering and Technology, Xi’an Jiaotong University,
No. 28 Xianning West Road, 710049 Xi’an, People’s Republic of China
T.T. Zhang  Z.X. Zhang (&)
State Key Laboratory of Multiphase Flow in Power Engineering,
Xi’an Jiaotong University, 710049 Xi’an, People’s Republic of China
e-mail: zhangzx@mail.xjtu.edu.cn

© Springer International Publishing AG 2017 217


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_10
218 Y.S. Yu et al.

KGaV Overall mass transfer coefficient, kmol/m2/s/MPa


MSN Molecular synergy number, dimensionless
P Operating pressure, MPa
Re Reynolds number, dimensionless
R Gas constant, 8.314 J/mol/K
T Temperature, K
t Time, s
yA,G Gas phase concentration, mol/mol
YA,G Mole ratio value
Z Column height, m
a Conversion rate, dimensionless

1 Introduction

With the increasing amount of greenhouse gas emission in the atmosphere, the
average temperature of the world rises during the past century, which has a serious
impact on the environment of the earth. Therefore, it is becoming urgent to take
effective measures to reduce carbon dioxide (CO2) emissions from its intensive
emission source, such as coal fired power plants and cement plants, since CO2 is
recognized as a typical greenhouse gas [1, 2]. Among all the methods to capture
CO2, chemical absorption is considered to be an effective way to cover the large
reduction of the greenhouse gas emission. During the chemical absorption, amines
(alkanolamines) are the typical absorbents to capture CO2. However, there are still
some problems for amines absorption of CO2, such as low CO2 absorption capacity
and high energy consumption for CO2 desorption [3]. In order to solve these
problems, improvements on the absorbent are proposed as one of the efficient
methods. After reviewing the recent improvements on the absorbents for CO2
capturing, several kinds of absorbents are analyzed below, including mixed aqueous
solutions, non-aqueous solutions and alternative amines. These categories are
determined as the energy efficient absorbents, which are quite important for
improving the performance of amine absorption of CO2. The provided detail
information of these absorbents is believed to be helpful for the theoretical research
and industrial application.

2 Blended Aqueous Energy Efficient Solutions

The absorption of acid gases in mixed amines has outstanding advantages over the
use of single amines. The addition of a small amount of primary amine to con-
ventional tertiary amines can increase the rate of absorption of CO2 to a large extent
without appreciably affecting the stripping characteristics [4, 5]. By varying the
Energy Efficient Absorbents for Industry … 219

relative concentrations of the amines, an optimum absorption system can, in


principle, be designed for a specific application. Blends of primary and tertiary
amines, such as MDEA + MEA + H2O, have been suggested for CO2 removal [5].
Compared to MDEA, AMP does not only have the same high CO2 loading capacity
but also has a higher reaction rate constant for the reaction with CO2 [5–7]. Thus,
AMP + MEA + H2O may be an attractive new solvent in addition to
MDEA + MEA + H2O for the acid gases treating process [8].
Due to the increasing importance of blended-amine systems in acid-gas treating
processes, it is necessary to understand the kinetic phenomena in mixed amine
systems. The reaction kinetics normally determines the mass transfer coefficient and
further determines the solution consumption amount, which has great impacts on
the sensible heat consumption in CO2 desorption.

2.1 Reaction Types of CO2 in Aqueous Solutions

In aqueous amines solutions, the reactions between amines and CO2 mainly consist
of two parts. The first reaction to be considered is the chemically process between
H2O and CO2,

CO2 + H2 O $ HCO
3 + H
þ
ð1Þ

This reaction rate is very slow and may be usually neglected.


The second reaction is the bicarbonate formation:

CO2 + OH $ HCO


3 ð2Þ

This reaction rate is fast and can enhance mass transfer even when the con-
centration of hydroxyl ion is low.
The second part is the reactions of CO2 with alkanolamines. Here, primary and
secondary amines reacting with CO2 have the same reaction process, but the tertiary
amines are different. The carbamate formation reaction occurs when CO2 reacts
with primary and secondary alkanolamines,

CO2 + 2R1 R2 NH $ R1 R2 NCOO + R1 R2 NH2þ ð3Þ

where R1 is an alkyl and R2 is H for primary amines and an alkyl for secondary
amines.
The zwitterion mechanism is generally accepted as the reaction mechanism for
the carbamate formation between CO2 with primary and secondary alkanolamines.
The zwitterion mechanism has been used successfully in aqueous alkanolamine
solutions as well as in some organic and viscous solutions [9, 10]. Besides the
primary and secondary alkanolamines, the zwitterion mechanism was also found to
220 Y.S. Yu et al.

be suitable for modeling the absorption of CO2 into aqueous AMP solutions and
into 1-propanol-AMP solvents [11, 12].
The reaction steps successively involve the formation of a zwitterion,

CO2 + R1 R2 NH $ R1 R2 NH þ COO ð4Þ

R1 R2 NH þ COO + B $ R1 R2 NCOO + BH þ ð5Þ

where B is a base that could be an amines, OH−, or H2O, and the corresponding
reactions are as follows,

R1 R2 NH þ COO + R1 R2 NH $ R1 R2 NCOO + R1 R2 NH2þ ð6Þ

R1 R2 NH þ COO + OH $ R1 R2 NCOO + H2 O ð7Þ

R1 R2 NH þ COO + H2 O $ R1 R2 NCOO + H3 O þ ð8Þ

For the reaction of CO2 with tertiary alkanolamines(R3N) in aqueous solution, as


no N–H occurs in the amine structure, tertiary amines cannot react directly with
CO2 [7, 13]. The following reaction mechanism was proposed [14],

CO2 + R3 N + H2 O $ R3 NH þ + HCO
3 ð9Þ

which is a based-catalyzed hydration of CO2.


In multi-amine solutions, with the different mixing rules, the reaction process
could be a combination of all the involved possible reactions.

2.2 Energy Efficient Amine Solvents

2.2.1 Binary Amine Solutions

The binary amine solutions are the most discussed mixtures for CO2 absorption,
such as MEA, DEA based amines solutions mixed with MDEA, AMP. In these
binary amine solutions, MEA and DEA were usually used as promoters to enhance
the absorption rate of CO2 while AMP and MDEA were used to reduce the energy
consumption for solvent regeneration [15–18].
In the researches above, the basic Henry constant H is defined as
 
H
log10 ¼ hI ð10Þ
HS
Energy Efficient Absorbents for Industry … 221

The diffusion coefficient D is simply offered as the function of the temperature,


which is

D ¼ expð13:275  2198:3=T  0:078142CÞ ð11Þ

By utilizing the parameters above, the fundamental absorption performance can


be assessed correctly with the help of reaction kinetics and viscosity parameters [6].
The performance of blended solutions was ever examined at a total alkanolamine
concentration of 3.0 kmol/m3, a mixing ratio of 1:1 mol/mol and three different
CO2 loadings (0, 0.25 and 0.4 mol/mol) [19]. The results show apparently that the
absorption performances of blended solutions are generally between the perfor-
mances of their separate component. When CO2 loading increased, the CO2 con-
centration of the blended solutions shifted upward, indicating a lower CO2
absorption efficiency. For MEA-MDEA and DEA-MDEA blend solutions, the
performance of blended solutions approached the performance of the rate promoters
(MEA and DEA) at a low CO2 concentration. With the increasing of CO2 loading,
the performance of blended MEA-MDEA and DEA-MDEA solutions moves to the
performance of MDEA. These results show that at low CO2 loading, the rate
promoters (MEA and DEA) played a dominant role because they reacted with CO2
faster than MDEA to form very stable carbamate compounds. As CO2 loading was
increased, more MEA or DEA absorbed CO2 to form carbamate, leading to the
increase in the ratio of unreacted MDEA to unreacted promoter. Subsequently, the
unreacted MDEA amount became the dominant factor for the CO2 absorption rate.
Besides MEA and DEA, piperazine (PZ), diglycolamine (DGA), aminoethy-
lethanolamine (AEEA), 3-methylaminopropylamine (MAPA), diethylenetriamine
(DETA), triethylenetetramine (TETA), and tetraethylenepentamine (TEPA) are also
used as activators to improve CO2 absorption performance of aqueous MDEA
solutions [20–24]. Take the last four alkyl amines with multiple amine groups for
example, their blends with MDEA have higher CO2 absorption capacities compared
with that of aqueous solutions of MEA [30% (w/w)] and MDEA [30% (w/w)].
MDEA/TEPA shows the highest CO2 loading amount of 0.753 mol/mol at 313 K.
Also, these MDEA blends show high cyclic capacities (0.241−0.330 mol/mol),
which are about 3 times higher than that of MEA. All MDEA blends show higher
absorption fluxes than MDEA. Some blends (MDEA/MAPA) showed about 8
times higher overall mass transfer coefficient than that of MDEA and even higher
than that of MEA. In addition, the absorption heats of the MDEA blends are about
30% higher than that of MDEA and about 30% lower than that of MEA [22].
However, scarce studies are performed for triethanolamine (TEA) blended amines.
As known to all, TEA is a kind of tertiary amine, which has advantages in absorption
capacity and desorption efficiency compared with MEA and DEA. Thus, TEA
blended amine is considered as an energy efficient solvent for CO2 capture.
Its drawback is the low absorption rate. Based on the principle of balancing
the absorption rate and desorption energy, MEA, PZ, AEEA, AEP (N-aminoethyl
piperazine) with high absorption rates, are here used to mix with TEA respectively to
222 Y.S. Yu et al.

produce new binary solutions. This will probably offer an alternative energy efficient
solvent for CO2 capture.
In order to assess the TEA blended amines, absorption and desorption experi-
ments are performed. The experiment is set up in Figs. 1 and 2.
As shown in Fig. 1, flue gas is firstly sent to the bottom of the absorption column
at a constant flow rate. After the countercurrent absorption process, the rich solution
gathered from the bottom is pumped to the top of the absorption column. The CO2
volume fraction in exhaust gas is monitored by an infrared gas analyzer and
recorded every minute till the outlet concentration becomes constant. The liquid
flow rate is 0.575 L/min. The absorption column diameter is 80 mm and the
packing is plastic Raschig rings and the packing height is 280 mm. The accuracy of
the infrared gas analyzer is 0.1%.
Desorption apparatus is shown in Fig. 2. After absorption, the produced rich
solution is desorbed in a conical flask atmospherically in an oil bath. CO2 generated
after desorption is dried and then measured by Drainage method. At the same time,
a certain amount of absorbed rich liquid is desorbed by strong acid and then
measured by Drainage method. The results are compared to calculate the desorption

1. Simulated flue gas 2. Pressure release valve 3. Flow meter


4. Packed column 5. Pump 6. Infrared gas analyzer

Fig. 1 CO2 absorption setup

Fig. 2 CO2 desorption setup

1. Thermostatic oil bath 2. Drying equipment


3. Wild-mouth bottle 4. Cylinder
Energy Efficient Absorbents for Industry … 223

ratio. In general, the CO2 fraction in industrial flue gas is 10–13%. Here, it is set as
13% and the flue gas is made up of N2 and CO2 only, which is defined as the
simulated flue gas.
By keeping the total amine concentration as 3 mol/L, the absorption performances
of the mixed solutions such as 3 mol/L TEA, 2 mol/L TEA+1 mol/L MEA, 2 mol/L
TEA+1 mol/L AEP, 2 mol/L TEA+1 mol/L AEEA and 2 mol/L TEA+1 mol/L PZ
are discussed and compared. The absorption rate of the mixed solutions changes with
time in Fig. 3 and the CO2 loading varies with time in Fig. 4.

Fig. 3 Absorption rates of


different mixed solutions

Fig. 4 Solution loadings of


different mixed solutions
224 Y.S. Yu et al.

As shown in Fig. 3, the absorption rates of the TEA blended amines show great
differences. At the beginning, the absorption rate of the blended TEA amines are 3
to 5 times of TEA solution, which verifies that adding additional amine does greatly
improve TEA absorption rate. This is due to the fact that the primary amine groups,
secondary amine groups have a better absorption property than tertiary amine
groups. Among these additives, AEP is the most effective one, which helps TEA to
obtain a high absorption rate for a long time. For the other discussed amines, the
absorption rate follows the order of PZ > AEEA > MEA. TEA blended with MEA
solution firstly absorbs CO2 at a high rate. In 33 min, the absorption rate becomes
lower than TEA solution. 50 min later, the absorption rate remains almost constant.
The initial absorption rate of TEA blended with AEEA is better than that blended
with MEA and PZ. Adding PZ into TEA solution achieves a great absorption rate
initially, which however decreases rapidly versus time. Even though the absorption
rate of the blended TEA amines decreases versus time, it is still higher than single
MEA and single TEA solutions.
Figure 4 shows that, with the increasing absorption time, the loading of TEA
blended amines increases accordingly and shows a significant difference. At the first
15 min, the loadings of the four kinds of TEA blended amines are quite close,
which are much higher than that of the 3 mol/L TEA solution. Thereafter, with the
decrease of the absorption rate, TEA and MEA blended amines tends to saturate
and thus no further increases in the loading. However, the loadings of the other
TEA blended amines still increase, corresponding to the high absorption rates. The
loading of TEA blended with AEP is the highest one due to fast absorption
occurring. The loading of the TEA blended with AEEA solution is inferior to TEA
blended with AEP. More importantly, after adding the AEP, the loading of the TEA
solutions has been improved significantly.
After the absorption experiment, a portion of the rich solution was heated to
120° in the oil bath until CO2 is totally desorbed. Then, the 50 mL rich solution and
50 mL lean solution are desorbed by using a strong acid. The corresponding pro-
duct is measured by Drainage method. The desorption results are offered in Table 1.
As can be seen in Table 1, among the TEA blended amine solutions, the TEA
blended with AEP shows the 12% higher desorption efficiency than the TEA
solutions. However, the other TEA blended amines shows lower desorption effi-
ciency comparing with TEA solution. This is because the AEP itself having tertiary

Table 1 Desorption of TEA based solutions


Composition Desorption amount in Desorption amount in Desorption
50 mL rich solution/mL 50 mL poor efficiency/%
solution/mL
3 mol/L TEA 414 120 71.01
TEA/MEA = 2:1 910 370 59.34
TEA/AEP = 2:1 1280 260 79.69
TEA/AEEA = 2:1 1083 380 64.91
TEA/PZ = 2:1 820 350 57.32
Energy Efficient Absorbents for Industry … 225

amine groups and thus it is easier to be desorbed than the rest. Therefore, con-
sidering absorption rate and desorption efficiency, TEA + AEP blended amine is
taken as a promising energy efficient solvent for CO2 capture. It is necessary to
determine the optimal composition for TEA + AEP blended amine.
By keeping the total amine concentration as 3 mol/L and changing the TEA and
AEP concentration, the absorption performance of the TEA + AEP blended amine
solutions such as 3 mol/L AEP, 1 mol/L TEA+2 mol/L AEP, 1.5 mol/L
TEA+1.5 mol/L AEP, 2 mol/L TEA+1 mol/L AEP and 2.5 mol/L TEA
+0.5 mol/L AEP are compared. The absorption rates results are presented in Fig. 5
and the CO2 loadings results are offered in Fig. 6.
As clearly observed in Figs. 5 and 6, with the increasing amount of AEP from
0.5 mol/L to 3 mol/L, the absorption rate of the TEA + AEP blended amine has
shown a decrease tendency. CO2 loading reaches a maximum at 1.5 mol/L.
For the desorption of CO2 + TEA + AEP, the rich solution is heated in the oil
bath at 120 °C. Simultaneously, the 50 mL rich solution and 50 mL lean solution
are desorbed by using a strong acid. The results are summarized in Table 2.
As clearly observed in Table 2, TEA + AEP blended amine with TEA/AEP ratio
of 1:1 shows the largest desorption efficiency of 82.26%, which shows great pro-
mise to reduce the energy consumption for CO2 capture.
All the results show that binary amines solutions can be used as alternatives to
MEA, achieving high CO2 loading and high reaction kinetics with low heat of
reaction. These would allow reducing the solvent flow rate and solvent regeneration
energy in the CO2 capture process, which strongly proves that binary amines are the

Fig. 5 Absorption rates of the TEA + AEP blended amine


226 Y.S. Yu et al.

Fig. 6 Solution loadings of the TEA + AEP blended amine

Table 2 Desorption of TEA + AEP blended amine


Composition Desorption amount in Desorption amount in Desorption
50 mLrich solution/mL 50 mL poor solution/mL efficiency/%
TEA/AEP = 2:1 1280 260 79.69
TEA/AEP = 1:1 1240 220 82.26
TEA/AEP = 5:1 1060 240 77.36
TEA/AEP = 1:2 1090 354 67.52
TEA/AEP = 0:3 950 331 65.16

energy efficient absorbents. However, the high viscosity and high corrosiveness of
amines with four and five amino groups such as TETA and TEPA have to be
considered seriously.

2.2.2 Ternary and More Energy Efficient Amine Solutions

Upon the binary amine mixing solutions, few works were done on ternary or more
kinds of amine solutions [25–29]. Adding AMP into a mixture of MDEA and DEA
is an attempt to produce ternary amine solution for CO2 capture. For ternary or
more complicated mixture amine solutions, experimental method is difficult to
know the complex physical and chemical properties due to high uncertainties under
multi-variable condition. Molecular dynamics (MD), as a mature computational
Energy Efficient Absorbents for Industry … 227

technology, could be used instead as an alternative method after validated models


are developed properly.
Considering the integration of field synergy theory and molecular dynamics
theory, a synergy molecular dynamics (SMD) model was developed to describe
multiple amine mixtures. The SMD model is able to provide the transport properties
of complicated mixture amines for CO2 capture. In the SMD model, the molecular
synergy number (MSN) was used to quantify the interactions between diffusion and
molecule motion, which are usually hard to be obtained by experiments. In the
SMD model, the route is to deduce the transport properties by the key parameter of
total energy. Here, the total energy of the amines and CO2 system is considered as
the sum of potential energy and kinetic energy. The potential energy is comprised
of four parts, i.e., bond stretching, angle bending, out of plane bending, and dihedral
motion potentials. Moreover, Lennard-Jones potentials are determined to charac-
terize the van der Waals interactions and electrostatic interactions. The kinetic
energies are calculated by the classical velocity computational method. After
potential and kinetic energies are obtained correctly, the diffusivity is achieved by
Green Kubo equation. The transfer characteristic of viscosity is thus induced by its
correlation to pressure tensors.
In order to achieve the optimization index for mixing amines, the molecular
synergy number (MSN) is developed by referencing the synergy number ever used
in the multi-field synergy analysis, which is given by

Dm 2km
MSN = ð12Þ
Dm0 15Re4=3

More detailed information is given in Yu’s work [29]. The MSN is able to help
with selection of the energy efficient solvent since the smaller MSN corresponding
to the less energy consumption for desorption.
The sensible heat consumption in the desorption process is affected by the CO2
loading and mass transfer coefficient. The higher CO2 loading and mass transfer
coefficient will normally requires less solution flux and thus produce less sensible
heat consumption in desorption. According to the mass transfer coefficient results of
the blended amines, the sensible heat consumption of the blended amines are
assessed accordingly, which helps to determine which blended amines are energy
efficient or not to some extent.
The sensible heat consumption amount for CO2-AMP desorption is set as the
baseline, which is represented by E0. The ratio of the sensible heat consumption
amount of CO2 and blended amines over CO2-AMP is presented by E/E0.
Here, MEA-DEA-AMP, MEA-DEA-TEA, MDEA-DEA-AMP and MDEA-
DEA-TEA with different weight fraction ratios are selected as the discussed ternary
systems as shown in Figs. 7, 8 and 9. With the analysis of molecular synergy number
(MSN), it is clearly seen that that the mixture of MDEA, DEA and TEA with a MSN
being 3.89 (the smallest one) is the best synergy one of the discussed combinations.
228 Y.S. Yu et al.

Fig. 7 Field synergy effects and sensible heat consumption of ternary amines mixtures

Fig. 8 Field synergy effects and sensible heat consumption of quaternary amine mixtures

This mixture achieves the best synergy between diffusion and molecule motion and
less drag force with mass transfer coefficient being 33.33% higher than that of
AMP. For the quaternary amines comprising of MEA-MDEA-DEA-AMP and
MEA-MDEA-DEA-TEA and quintuple amine solution of MEA-MDEA-
DEA-AMP-TEA, the mass transfer coefficient are all higher than that of refer-
enced AMP system. This proves that mixed amines improve the CO2 absorption
Energy Efficient Absorbents for Industry … 229

Fig. 9 Field synergy effects and sensible heat consumption of quintuple amine mixtures

performance from the mass transfer view of point. Also, the overall synergy in
ternary amine systems has been found better than that in quaternary and quintuple
amine systems. All the results could provide a benchmark for the amine mixing
solution design with better CO2 capture.
As shown in Fig. 7, the MEA-DEA-TEA is an energy efficient solvent since its
sensible heat consumption amount is 40% lower than that of the AMP baseline.
This is due to the fact that higher mass transfer coefficient of MEA-DEA-TEA
produces much less solvent consumption amount by absorbing equal CO2. For the
quaternary amine, the similar results are found that the sensible heat consumption
for desorption of CO2-MEA-MDEA-DEA-TEA is 16–27% lower than that of the
AMP baseline in Fig. 8. For the quintuple amine MEA-MDEA-DEA-AMP-TEA
and CO2 system, the sensible heat consumption amount decreases to 68–79% as
shown in Fig. 9. All the results prove that blended amines are energy efficient
solvents for CO2 capture. The sensible heat consumption can be qualitatively
assessed by the MSN since they have similar variation tendency versus the amine
weight fraction as shown in Figs. 7, 8 and 9.
According to the discussion above, the mixed amine solutions have the potential
to perform better than the single alkanolamines as long as mixing species and
weight ratios are well-designed. The higher mass transfer coefficient of the mixed
amine solutions offers the less consumption of the absorbent amount, which could
reduce the sensible heat consumption in the desorption process. Hence, the mixed
amine solutions are energy efficient absorbents for CO2 capture.
230 Y.S. Yu et al.

3 Non-aqueous Energy Efficient Solutions

Apart from aqueous alkanolamine solutions for CO2 removal from gas produced by
burning fossil fuels, non-aqueous systems comprising methanol solutions of alka-
nolamines have been employed commercially for the absorption of acid gases (CO2,
H2S, COS) for the high solubility and capacity, low corrosiveness, and low energy
consumption during generation of the used liquor. Methanol is widely used as a
physical solvent for CO2 removal from gas streams, and the solubility of CO2 in
aqueous monoethanolamine can be enhanced by the presence of methanol, which
was 25% higher in a MEA and methanol mixture than in an aqueous MEA solution
of equivalent MEA concentration. Mixed solvents (chemical and physical) are
expected to have a higher capacity for the acid gases (CO2) over a wide range of
partial pressures than the separate physical or chemical solvent. Meanwhile,
non-aqueous solvent such as methanol has lower boiling point and smaller latent
heat which leads to lower operating temperature in the regeneration process and
thus requires a lower regeneration energy amount compared with aqueous envi-
ronment. Hence, the non-aqueous alkanolamines have potentials to improve the
performance of CO2 absorption.

3.1 Reaction Mechanism

The Eqs. (3), (4) and (5) show the overall reaction between CO2 and alkanolamines
of primary and secondary in aqueous solution [30–33]. In case of non-aqueous
solvents, only amine is considered as the base in the proton removal step.

CO2 + R1 R2 NH $ R1 R2 NH þ COO ð13Þ

R1 R2 NH þ COO + R1 R2 NH $ R1 R2 NCOO þ R1 R2 NH2þ ð14Þ

For the tertiary amines, since they do not form carbamates and act as homo-
geneous base catalysts for carbon dioxide hydrolysis, the reaction Eq. (4) cannot be
used to explain the reaction mechanism between CO2 and TEA/MDEA in
non-aqueous solution [9, 34]. It was proposed that in non-aqueous solutions of
tertiary amine the dissolved carbon dioxide will react with solvated tertiary amine to
form anion pair as follows:

R3 NH  HSOL þ CO2 ! R3 NH þ CO2 SOL ð15Þ

where HSOL represents the solvent.


Experiment works on reaction kinetics of CO2 absorption in non-aqueous
alkanolamines, such as monoethanolamine (MEA) and triethanolamine (TEA) in
methanol, n-propanol and ethylene glycol with different weight mixing ratio were
done at 298 K. The reaction orders and rate constants were achieved, and the
Energy Efficient Absorbents for Industry … 231

Table 3 The reaction orders and rate constants of CO2 in the reaction of CO2 with non-aqueous
MEA at 298 K [34]
Solvent n K1 K2/K3
water 1.53 6312 (7740) 1.73 (1.28)
methanol 1.79 (1.62) 5631 (8330) 7.41 (1.28)
ethanol 1.85 (1.62) 5479 9.65 (2.91)
n-propanol 1.85 (1.90) 5361 11.23 (3.66)
n-butanol 1.86 5408 11.88
Ethylene glycol 1.82 5731 7.24
Propylene glycol 1.85 5536 10.14
Propylene carbonate 1.83 5621 8.79

experimental results of MEA in non-aqueous solutions absorbing CO2 are shown in


Table 3[34–37].
The reaction orders of the solvents depended on the polarity of the solvent,
which have some relationship with the dielectric constant of the solvent. For the
reaction rate constant, linear relationship with the solubility parameter of the solvent
was obtained in several experiments. The magnitude of the rate constants may be a
function of the degree to which the solvent is able to stabilize the zwitterionic
intermediate. Thus, the reaction of CO2 with MEA in polar solvents was identified
as the reaction scheme above such as the zwitterionic mechanism.
Besides experiment, as computational calculation developed, ab initio calcula-
tions were also employed to investigate the reaction mechanism of alkanolamines
absorbing CO2 in gas phase and in aqueous and non-aqueous solvation effects [38,
39]. With different calculating methods and basis sets, the reaction energies were a
little different from each other. The reaction pathways are almost consistent with the
zwitterion mechanism, which are two steps reaction processes with zwitterion as
intermediate, followed by a proton transfer process with another base (amine, H2O
molecule). According to different bases, some of the second proton transfer pro-
cesses are almost energy-free and one-step and three-molecular reaction process
was also proposed. The reaction energy barriers show that the highest energy barrier
appears in the water solvation effect.
All the results above prove that non-aqueous solvents show great mass transfer
coefficient from the reaction kinetics data and ab initio calculation data, which
produces less solvent consumption in absorbing equal amount of CO2 compared
with the MEA baseline. Thus, the non-aqueous solvent is another kind of energy
efficient solvent in CO2 capture, which has great potential in industrial application.

3.2 Absorption Performance of the MEA-Methanol System

To compare the absorption performances between MEA aqueous and non-aqueous


solutions, studies on CO2 absorption into three solvent solutions, 5 M MEA-methanol
232 Y.S. Yu et al.

solution, 5 M MEA in 1:1 water-methanol (volume ratio) and 5 M MEA aqueous


solutions, were investigated in a packed column, offering overall mass transfer
coefficient (KGav) and mass flux of CO2 absorption over ranges of methanol com-
position, CO2 loading, liquid flow rate and inert gas flow rate. In the film theory, by
combing mass flux and material balance equations at steady-state condition, the
overall mass transfer coefficient is defined as Eq. (16) [40, 41],
!     
G dYA;G G 1 + YA;G dYA;G
K G av ¼ ¼ ð16Þ
PyA;G dZ P YCO2 dZ

The results show that the mass transfer performance of 5 M MEA in methanol
was higher than those of 5 M MEA in 1:1 water-methanol and 5 M MEA aqueous
solutions, respectively, as the methanol has higher physical solubility and physical
diffusivity of CO2 than water. The higher methanol composition leads to the higher
physical solubility and physical diffusivity of CO2. For the CO2 loading, as CO2
loading increases in the range of 0.05–0.3 mol/mol, the mass transfer performance
in terms of KGav decreases due to that the amount of active amine decreases as CO2
loading increases. Increasing liquid flow rate can lead to greater degree of wetted
packing surface and increasing of mass transfer performance. However, this result
is controversial since the bubbles will be produced at high liquid flow rates, which
directly affects the active surface area between CO2 and absorbent. Thus, the
optimum operating flow rate differs from the actual operating conditions. As the
CO2 absorption process is liquid film control, the inert gas flow rate has no sig-
nificant effect on mass transfer performance. Above all, the mass transfer perfor-
mance was higher in MEA methanol solutions than aqueous solutions.

3.3 Multi-stage Energy Efficient Process for CO2-


MEA-Methanol Regeneration System

As methanol can enhance the mass transfer of MEA absorbing CO2 process and
show the properties different from water, a new regeneration operation process
specially designed for MEA-methanol system was proposed. It’s a multi-stage
regeneration process which can be operated below 373 K [33].
For the MEA-methanol regeneration system, the main reaction considered and
reaction rate constant at 343–363 K are shown as follows.

MEAH þ + MEACOO ! 2MEA + CO2 ð17Þ

28642:44
k = exp(  84:478 þ Þ ð18Þ
T
Energy Efficient Absorbents for Industry … 233

In the operation process, to demonstrate the CO2 desorption performance in the


absence of water, three cases using N2, ethanol vapor and steam as the purge gases
were analyzed for the same amount of lean solution. The results show that the
operating conditions (temperature, pressure), lean solvent loadings, gas/liquid ratio,
packing and internal all influence the regenerating performance. A parametric
analysis found that the energy consumption could be reduced by 7−24%.
A minimum energy consumption of 2.28 GJ/t was identified in the non-aqueous
process. It was clearly found that increasing the temperature, pressure and lean
solvent loading could reduce the energy consumption by 21, 22 and 20%,
respectively. Regarding the effect of the packing, it is noted that a random packing
was found to consume less energy than a structured packing. The CMR-2 packing
produced the lowest energy consumption regardless of the desorption conditions.
The internals placed in the stripper intensified the desorption process, with energy
consumption being reduced by about 23%. The non-aqueous solvent could improve
the desorption efficiency by 10% compared to that obtained in a typical aqueous
solution desorption.
Apart from the typical non-aqueous solution of MEA in methanol, other dis-
cussed non-aqueous solutions for CO2 capture include MDEA/diisopropanolamine
(DIPA)/triethanolamine (TEA) with polar organic solvents, MEA/triethylene glycol
(TEG), AMP and some AMP–alkanolamine blends (IPMEA, TBMEA) in
non-aqueous solvents (triethylene glycol (TEG), diethylene glycol (DEG), ethylene
glycol and 1-propanol), 2-(2-aminoethylamine)ethanol (AEEA)+benzylalcohol and
so on [36, 42–44]. These non-aqueous capture systems were also under discussion
for their absorbing characteristics, which show the good performance being oper-
ated at low stripping temperature (under 373 K), which have the potential to reduce
the regeneration energies. However, more works needed to be done to determine the
optimum operating conditions, and to avoid the corrosion and degradations of the
proposed solutions.
Above all, compared with traditional aqueous alkanolamine solutions, the
non-aqueous solutions could do a better work on CO2 absorption. For the regen-
eration process, they can lower the regeneration temperature, which could greatly
lead to the reduction of energy cost in the regeneration process. Therefore,
non-aqueous solutions are energy efficient absorbents to capture CO2.

4 Alternative Energy Efficient Solvent

In order to find out the alternatives to traditional alkanolamines, several other new
solvents or special mixtures are developed in capturing CO2. The developed ones
offer the large capacity and low energy consumption and thus show a wide
application potential. Recently, tetramethylammonium hydroxide (TAMH) has
been already identified as an alternative solvent to capture CO2 for the advantages
of using its CO2 absorption product for photoresist and etch residue removal.
Hence, it is interesting to do research on using TAMH to solve large CO2 emission
234 Y.S. Yu et al.

and utilization problem to some extent. However, the limitation of using TAMH is
that some tetramethylammonium carbamates are easy to be precipitated which
barricades the residue removal. To avoid the shortcoming, several physical solvents
are proposed to dissolve the tetramethylammonium carbamates. Here, the typical
physical solvent is selected as the tetramethylene sulfone (TMS), which shows great
solubility for amine and CO2. Additionally, ethylene glycol (EG) is considered as
another typical solvent for physical absorption of CO2 [45–47]. More importantly,
TMS and EG show great ability to provide good solubility for tetramethylammo-
nium carbamates, which affords to solve the precipitation problem for TAMH
absorption of CO2. Thus, TAMH-TMS-EG aqueous solution was suggested as a
new solvent mixture for CO2 capture [48].

4.1 Reaction Mechanism

As TAMH reacts with CO2 to form carbonate and TMS-EG provides the strong
dissolution of tetramethylammonium carbonate, it is difficult for the tetramethy-
lammonium carbonate to exist in the TAMH-TMS-ETG-CO2 system. Therefore,
the main reaction type is reasonably summarized as the formation of carbonate
between TAMH and CO2, shown as follows.

TAMH + CO2 ! TAM þ HCO


3 ð19Þ

Based on the reaction equation, the reaction kinetics is accurately determined as

da
¼ 8:975  1035  e200590=RT  ð1aÞ0:91 ð20Þ
dt

The precise reaction kinetics determines the reaction rate of CO2-


TAMH-TMS-EG system, which is greater than the reaction rates of typical aqueous
CO2-MEA and CO2-MDEA [49]. Thus, greater reaction rate of CO2-
TAMH-TMS-EG provides the strong base for increasing the mass transfer coeffi-
cient, which suggests that TAMH-TMS-EG is suitable for CO2 absorption.

4.2 Absorption Performance and Sensible Heat


Consumption

Since the TAMH-TMS-EG is a new solvent, its CO2 absorption performance is


determined precisely. CO2 solubility in the TAMH-TMS-EG solutions, as the basic
mass transfer parameter, is carefully tested and discussed. It is concluded that the
2–25 kPa partial pressure of CO2 produces the corresponding CO2 loading of
0.15–0.453 mol/molTAMH respectively at 313.15 K. This CO2 loading is
Energy Efficient Absorbents for Industry … 235

comparable to that in the typical MEA system, which proves that the solubility of
CO2 in TAMH-TMS-EG is high enough for CO2 absorption. At 373.15 K, the
partial pressure of CO2 varies from 100 to 675 kPa, which offers the CO2 loading of
0.18–0.5 mol/molTAMH. This average CO2 partial pressure is a little less than the
partial pressure of CO2 in MDEA-TMS aqueous system and suggests that it is
supposed to strip CO2 more easily in TAMH-TMS-EG solution. This is quite useful
to achieve the low cost CO2 capture process.
During the CO2 absorption operation assessment, CO2 loading is the key
parameter to achieve the continuous run. CO2 fraction, solvent composition
(TAMH fraction and TMS fraction) and pressure normally influence CO2 loading in
the solvent. Therefore, the effects of these factors were discussed in detail here. In
order to explain the energy efficient advantage of the TAMH-TMS-EG solutions,
the sensible heat consumption amount of CO2-MEA is set as the baseline. The ratio
of the sensible heat consumption amount of CO2-TAMH-TMS-EG desorption over
CO2-AMP desorption is presented by E/E0.
CO2 fraction in flue gas is normally hard to keep constant with the dynamic
operating conditions in industrial process. Here, the CO2 mol fraction is set by
5–25%, which are the common states in the industrial flue gas. The results in
Fig. 10 provide that the sensible heat consumption ratio E/E0 apparently decreases
as the CO2 mol fraction increases at 303.15, 308.15 and 313.15 K, respectively.
This offers that the sensible heat consumption of CO2-TAMH-TMS-EG desorption

Fig. 10 Effects of CO2 fraction on sensible heat consumption


236 Y.S. Yu et al.

is 25–45% lower than that of the CO2-MEA desorption in the typical 10–15%
CO2 mol fraction in the power plant, which strongly supports that TAMH-TMS-EG
is an energy efficient solvent. This is attributed to the fact that TAMH-TMS-EG
absorbs more CO2 than MEA under equal solution flux conditions.
Taking the solvent composition into account, the weight fractions of TAMH and
TMS is supposed to influence CO2 loading, which are depicted in Figs. 11 and 12.
It is shown that the sensible heat consumption ratio E/E0 decreases from 0.85 to
0.45 as TAMH weight fraction increases from 5 to 25%. This interesting conclusion
is due to the fact that TAMH is involved in the chemical reaction with CO2, which
results in that the redundant TAMH naturally absorbs more CO2. This will certainly
consume much less solvent compared with MEA case and thus produce less sen-
sible heat consumption amount. In this sense, the sensible heat consumption ratio
E/E0 is decreased by 15% when the temperature increases from 303.15 to 313.15 K.
This is partially due to the exothermic heat occurring in the TAMH absorption of
CO2 process.
Another composite TMS, one of the physical solvents, also has great influence
on the sensible heat consumption ratio E/E0. Its effects on sensible heat con-
sumption ratio E/E0 is provided in Fig. 12. As expected, sensible heat consumption
ratio E/E0 increases from 0.36 to 0.43 as TMS weight fraction increases from 55 to
75%. The meaningful result is attributed to the fact that the absorption capacity of
the physical solvent TMS is not so strong as that of chemical solvent of TAMH.

Fig. 11 Effects of TAMH fraction on sensible heat consumption


Energy Efficient Absorbents for Industry … 237

Fig. 12 Effects of TMS fraction on sensible heat consumption

This is further validated by that the CO2 loading is averagely decreased by 24% as
temperature ascends from 303.15 to 313.15 K.
After the chemical reaction and dissolution are considered to affect the
TAMH-TMS-ETG absorption of CO2, diffusion is believed to be another important
factor that plays great role in the CO2 absorption. Thus, the mass transfer resistance
able to achieve the goal of the further solvent composition optimization is studied as
follows.
According to the previously correct findings, the Hatta number is here used as a
variable to determine the solvent chemical property. Also, a dimensionless solu-
bility is employed to indicate the ratio of the CO2 concentration between liquid
phase and gas phase at equilibrium state. It is quite clear that the Hatta number is
below 3 and agrees with the literature results [49] for CO2 absorption in the sol-
vents. The most interesting one is that the minimum mass transfer resistance cor-
responds to the TMS weight fraction from 40 to 80%. The TMS concentration
range produces the higher Hatta number accordingly. Additionally, this TMS
weight fraction range offers the higher CO2 solubility above 1. All the results may
help well with the future TAMH-TMS-ETG solvent optimization as it is used in
industry to mitigate the CO2 emission.
Since the energy consumption for desorption of the solutions and CO2 deter-
mines the CO2 capture cost, energy consumption comparisons with typical
MEA-CO2-H2O and MEA-CO2-TMS are made to determine the overall CO2
capture performance of TAMH-TMS-EG. It is suggested that the average CO2
238 Y.S. Yu et al.

loading is 25% higher in TAMH-TMS-EG than that in MEA-CO2-H2O and


MEA-CO2-TMS. Moreover, the energy consumption is determined as 1.11–1.34
GJ/t for TAMH-TMS-EG absorption of CO2. This energy cost is much less than the
typical 3.0–4.5 GJ/t in MEA solution absorption of CO2. It is believed that the
molecules of TMS have a higher dielectric constant (about 82.56 at 303.15 K)
compared with water. Thus, TMS becomes a more active hydrogen atom donor and
accept or rather than keep intermolecular relations. All in all, the TAMH-TMS-EG
was proven as an energy efficient solvent candidate for effective CO2 capture.

5 Conclusions

In order to reduce the energy consumption, energy efficient solvents are summa-
rized and analyzed in details. Reaction kinetics, desorption efficiency and sensible
heat consumption are discussed since these parameters determine the energy con-
sumption amount directly. The proper composition and operating conditions for
binary amines, ternary amines and quaternary amines, non-aqueous amine and
alternative solvents being the energy efficient solvents are determined. The higher
CO2 loading and mass transfer coefficient of these solvents help to obtain much less
sensible heat consumption compared with typical MEA baseline, which strongly
support their energy efficient characteristic and shows great economical advantages
in the industrial application for CO2 capture.

Acknowledgments Financial support of National Natural Science Foundation of China (no.


51276141) is gratefully acknowledged. This work is also supported by the Natural Science Basic
Research Plan in Shaanxi Province of China (No. 2015JQ5192) and “Fundamental Research
Funds for the Central Universities”.

References

1. Mathias PM, Zheng F, Heldebrant DJ, Zwoster A, Whyatt G, Freeman CM et al (2015)


Measuring the absorption rate of CO2 in nonaqueous CO2-binding organic liquid solvents
with a wetted-wall apparatus. Chem Sus Chem 8:3617–3625
2. Walters MS, Edgar TF, Rochelle GT (2016) Regulatory control of amine scrubbing for CO2
capture from power plants. Ind Eng Chem Res 55:4646–4657
3. Singto S, Supap T, Idem R, Tontiwachwuthikul P, Tantayanon S, Al-Marri MJ et al (2016)
Synthesis of new amines for enhanced carbon dioxide (CO2) capture performance: The effect
of chemical structure on equilibrium solubility, cyclic capacity, kinetics of absorption and
regeneration, and heats of absorption and regeneration. Sep Purif Technol 167:97–107
4. Abu-Zahra MR, Abbas Z, Singh P, Feron P (2013) Carbon dioxide post-combustion capture:
solvent technologies overview, status and future directions. Materials and processes for
energy: communicating current research and technological developments. Formatex Research
Center, Badajoz, pp 923–34
5. Chakravarty T, Phukan U, Weilund R (1985) Reaction of acid gases with mixtures of amines.
Chem Eng Prog (U.S) 81:32–36
Energy Efficient Absorbents for Industry … 239

6. Budzianowski WM (2015) Single solvents, solvent blends, and advanced solvent systems in
CO2 capture by absorption: a review. Int J Glob Warming 7:184–225
7. Crooks JE, Donnellan JP (1990) Kinetics of the reaction between carbon dioxide and tertiary
amines. J Organic Chem 55:1372–1374
8. Li M-H, Chang B-C (1994) Solubilities of carbon dioxide in Water + Monoethanol
amine + 2-Amino-2-methyl-1-propanol. J Chem Eng Data 39:448–452
9. Versteeg G, Van Dijck L, Van Swaaij W (1996) On the kinetics between CO2 and
alkanolamines both in aqueous and non-aqueous solutions. An overview. Chem Eng
Commun 144:113–158
10. Versteeg G, Van Swaaij W (1988) On the kinetics between CO2 and alkanolamines both in
aqueous and non-aqueous solutions—I. Primary and secondary amines. Chem Eng Sci
43:573–585
11. Bosch H, Versteeg G, Van Swaaij W (1990) Kinetics of the reaction of CO2 with the sterically
hindered amine 2-amino-2-methylpropanol at 298 K. Chem Eng Sci 45:1167–1173
12. Xu S, Wang Y-W, Otto FD, Mather AE (1996) Kinetics of the reaction of carbon dioxide with
2-amino-2-methyl-1-propanol solutions. Chem Eng Sci 51:841–850
13. Rangwala H, Morrell B, Mather A, Otto F (1992) Absorption of CO2 into aqueous tertiary
amine/MEA solutions. Can J Chem Eng 70:482–490
14. Donaldson TL, Nguyen YN (1980) Carbon dioxide reaction kinetics and transport in aqueous
amine membranes. Ind Eng Chem Fundam 19:260–266
15. Dash SK, Samanta AN, Bandyopadhyay SS (2014) Simulation and parametric study of post
combustion CO2 capture process using (AMP + PZ) blended solvent. Int J Greenhouse Gas
Control 21:130–139
16. Li H, Le Moullec Y, Lu J, Chen J, Marcos JCV, Chen G (2014) Solubility and energy
analysis for CO2 absorption in piperazine derivatives and their mixtures. Int J Greenhouse Gas
Control 31:25–32
17. Smith K, Lee A, Mumford K, Li S, Thanumurthy N, Temple N et al (2015) Pilot plant results
for a precipitating potassium carbonate solvent absorption process promoted with glycine for
enhanced CO2 capture. Fuel Process Technol 135:60–65
18. Xiao J, Li C-W, Li M-H (2000) Kinetics of absorption of carbon dioxide into aqueous
solutions of 2-amino-2-methyl-1-propanol+ monoethanolamine. Chem Eng Sci 55:161–175
19. Aroonwilas A, Veawab A (2004) Characterization and comparison of the CO2 absorption
performance into single and blended alkanolamines in a packed column. Ind Eng Chem Res
43:2228–2237
20. Al-Juaied M, Rochelle GT (2006) Absorption of CO2 in aqueous diglycolamine. Ind Eng
Chem Res 45:2473–2482
21. Bishnoi S, Rochelle GT (2002) Absorption of carbon dioxide in aqueous piperazine/methyl
diethanolamine. AIChE J 48:2788–2799
22. Choi SY, Nam SC, Yoon YI, Park KT, Park S-J (2014) Carbon Dioxide Absorption into
Aqueous Blends of Methyldiethanolamine (MDEA) and alkyl amines containing multiple
amino groups. Ind Eng Chem Res 53:14451–14461
23. Zhang X, Zhang C-F, Qin S-J, Zheng Z-S (2001) A kinetics study on the absorption of carbon
dioxide into a mixed aqueous solution of methyldiethanolamine and piperazine. Ind Eng
Chem Res 40:3785–3791
24. Zoghi AT, Feyzi F, Zarrinpashneh S (2012) Experimental investigation on the effect of
addition of amine activators to aqueous solutions of N-methyldiethanolamine on the rate of
carbon dioxide absorption. Int J Greenhouse Gas Control 7:12–19
25. Adeosun A, Abu-Zahra MR (2013) Evaluation of amine-blend solvent systems for CO2
post-combustion capture applications. Energy Procedia. 37:211–218
26. Huang Y, Zhang X, Zhang X, Dong H, Zhang S (2014) Thermodynamic modeling and
assessment of ionic liquid-based CO2 capture processes. Ind Eng Chem Res 53:11805–11817
27. Mumford KA, Smith KH, Anderson CJ, Shen S, Tao W, Suryaputradinata YA et al (2011)
Post-combustion capture of CO2: results from the solvent absorption capture plant at
Hazelwood power station using potassium carbonate solvent. Energy Fuels 26:138–146
240 Y.S. Yu et al.

28. Thee H, Nicholas NJ, Smith KH, da Silva G, Kentish SE, Stevens GW (2014) A kinetic study
of CO2 capture with potassium carbonate solutions promoted with various amino acids:
Glycine, sarcosine and proline. Int J Greenhouse Gas Control 20:212–222
29. Yu Y, Lu H, Wang G, Zhang Z, Rudolph V (2013) Characterizing the Transport Properties of
Multiamine Solutions for CO2 Capture by Molecular Dynamics Simulation. J Chem Eng Data
58:1429–1439
30. Beyad Y, Puxty G, Wei S, Yang N, Xu D, Maeder M, et al (2014) An SO2 tolerant process for
CO2 capture|NOVA. The University of Newcastle’s Digital Repository
31. Budzianowski WM (2011) Mitigating NH3 vaporization from an aqueous ammonia process
for CO2 capture. Int J Chem Reactor Eng 9:1–27
32. Cousins A, Wardhaugh LT, Feron PH (2011) Preliminary analysis of process flow sheet
modifications for energy efficient CO2 capture from flue gases using chemical absorption.
Chem Eng Res Des 89:1237–1251
33. Yu Y, Lu H, Zhang T, Zhang Z, Wang G, Rudolph V (2013) Determining the performance of
an efficient nonaqueous CO2 capture process at desorption temperatures below 373 K. Ind
Eng Chem Res 52:12622–12634
34. Park S-W, Choi B-S, Lee J-W (2006) Chemical absorption of carbon dioxide with
triethanolamine in non-aqueous solutions. Korean J Chem Eng 23:138–143
35. Chen S, Chen S, Zhang Y, Qin L, Guo C, Chen J (2016) Species distribution of CO2
absorption/desorption in aqueous and non-aqueous N-ethylmonoethanolamine solutions. Int J
Greenhouse Gas Control 47:151–158
36. Guo C, Chen S, Zhang Y, Wang G (2014) Solubility of CO2 in Nonaqueous Absorption
System of 2-(2-Aminoethylamine) ethanol+ Benzyl Alcohol. J Chem Eng Data 59:1796–1801
37. Park S-W, Lee J-W, Choi B-S, Lee J-W (2005) Kinetics of absorption of carbon dioxide in
monoethanolamine solutions of polar organic solvents. J Ind Eng Chem 11:202–209
38. Da Silva EF, Svendsen HF (2007) Computational chemistry study of reactions, equilibrium
and kinetics of chemical CO2 absorption. Int J Greenhouse Gas Control 1:151–157
39. Zhang T, Zhang Z (2014) Computational study of CO2 absorption in Aqueous and
non-aqueous solutions using MEA. Energy Procedia. 63:1347–1353
40. Naami A, Edali M, Sema T, Idem R, Tontiwachwuthikul P (2012) Mass transfer performance
of CO2 absorption into aqueous solutions of 4-diethylamino-2-butanol, monoethanolamine,
and N-methyldiethanolamine. Ind Eng Chem Res 51:6470–6479
41. Sema T, Naami A, Fu K, Edali M, Liu H, Shi H et al (2012) Comprehensive mass transfer and
reaction kinetics studies of CO2 absorption into aqueous solutions of blended MDEA–MEA.
Chem Eng J 209:501–512
42. Li J, You C, Chen L, Ye Y, Qi Z, Sundmacher K (2012) Dynamics of CO2 absorption and
desorption processes in alkanolamine with cosolvent polyethylene glycol. Ind Eng Chem Res
51:12081–12088
43. Tan J, Shao H, Xu J, Du L, Luo G (2011) Mixture absorption system of monoethanolamine−
triethylene glycol for CO2 capture. Ind Eng Chem Res 50:3966–3976
44. Zheng C, Tan J, Wang Y, Luo G (2012) CO2 Solubility in a mixture absorption system of
2-Amino-2-methyl-1-propanol with Glycol. Ind Eng Chem Res 51:11236–11244
45. Galvão AC, Francesconi AZ (2010) Solubility of methane and carbon dioxide in ethylene
glycol at pressures up to 14 MPa and temperatures ranging from (303 to 423) K. J Chem
Thermodyn 42:684–688
46. Gui X, Tang Z, Fei W (2011) Solubility of CO2 in alcohols, glycols, ethers, and ketones at
high pressures from (288.15 to 318.15) K. J Chem Eng Data 56:2420–2429
47. W-m Qian, Y-g Li, Mather AE (1995) Correlation and prediction of the solubility of CO2 and
H2S in an aqueous solution of methyldiethanolamine and sulfolane. Ind Eng Chem Res
34:2545–2550
48. Yu YS, Zhang TT, Wu XM, Mu DL, Zhang ZX, Wang GGX (2015) Exploiting an alternative
CO2 absorption process by efficient solvent mixture. Ind Eng Chem Res 54:6165–6174
49. Derks P, Versteeg G (2009) Kinetics of absorption of carbon dioxide in aqueous ammonia
solutions. Energy Procedia. 1:1139–1146
The Absorption Kinetics of CO2 into Ionic
Liquid—CO2 Binding Organic Liquid
and Hybrid Solvents

Ozge Yuksel Orhan, Cyril Sunday Ume and Erdogan Alper

Abstract Carbon dioxide (CO2) capture is a global concern because of its effect on
climate change especially as regards to global warming. Among greenhouse gases,
CO2 is the most abundant with high concentration released from post combustion
processes into the atmosphere. For instance, the volume of CO2 emission from
thermal power plants, petroleum refineries, petrochemical plants, hydrogen and
cement factories has become one of the top important global concerns nowadays. In
order to capture post-combustion CO2 and securely store it way or to produce useful
products from it, it requires separation of CO2 from flue gas stream. Industrially, it is
generally accepted that the most appropriate method that can be applied commer-
cially to capture CO2 involves absorbing it with a reversible reaction from gas
streams into aqueous amine especially monoethanolamine. Although, CO2-aqueous
amine process is accepted as a mature technology but its absorption/desorption
systems are the subject of several studies as the process is energy-intensive among
other issues. In view of the shortfalls of CO2-aqueous amine systems and the greater
societal concern to control the amount of CO2 released to the environment from
industrial sources to abate its effects, research for alternative viable solvent systems
becomes of high interest to researchers as well as industrialists. Hence, this chapter is
mainly to focus on highlighting and discussing relevant advanced solvent systems
for CO2 capture. Among the novel solvents or technology worthy of discussion here
include use of organic solvents consisting of an amidine or a guanidine and a linear
alcohol, such as 1-hexanol, instead of aqueous amines. In this case, CO2 loaded
solvent could be regenerated at 90–100 °C which is much lower than the boiling
point of the solvent and as a result, sufficient drop in energy requirements could be

O.Y. Orhan (&)  E. Alper


Department of Chemical Engineering, Hacettepe University, Ankara, Turkey
e-mail: oyuksel@hacettepe.edu.tr
E. Alper
e-mail: ealper@hacettepe.edu.tr
C.S. Ume
Department of Chemical and Petroleum Engineering, Federal University
Ndufu-Alike Ikwo (FUNAI), Abakaliki, Nigeria
e-mail: umesoncy@gmail.com

© Springer International Publishing AG 2017 241


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_11
242 O.Y. Orhan et al.

achieved. In order to be applicable, CO2 binding organic liquid systems


(“CO2BOLs”) will react with CO2 at sufficient rates. In case, the reaction rate is not
sufficient it can be upgraded with piperazine derivatives. Another potential novel
method is the capture of CO2 by special liquids which have negligible vapor pressure
and high thermal stability known as ionic liquids (ILs). The ILs also have favorable
CO2 solubility and a wide liquid state temperature with tunable physicochemical
characteristics. Further advanced approach can be to develop a blended solvent
which can also react with CO2 more efficiently. A blended system is a hybrid that
possesses combined benefits of the amines mixture components thereby providing a
better alternative than using single component. Other novel solvents for post com-
bustion CO2 capture are noted and discussed here.

Nomenclature—Symbols and Acronyms


AMP 2-amino-2-methylpropanol
[BF4−] Tetrafluoroborate
[BR−] Bromide
CO2 Carbon dioxide
CO2BOLs Carbon dioxide binding organic liquids
CO + H2 Synthetic gas
CH4 Methane
[CI−] Chloride
CFCs Chlorofluorocarbons
[(CF3SO2)2 N−] Bis (trifluoromethylsulfonyl) imide
[CF3SO3−] Triflate
[CH3CO2−] Acetate
[CF3CO2−] Trifluoroacetate
[(CN)2 N−] Dicyanamide
CS2 Carbon disulphide
DBU 1,8-diazabicyclo[5,4,0] undec-7-ene
DEA Diethanolamine
GHG Greenhouse gas
GHGs Greenhouse gases
HFCs Hydrofluorocarbon
H 2S Hydrogen sulphide
ILs Ionic liquids
[I−] Iodide
ko Observed reaction rate constant
MDEA Methyldiethanolamine
MEA Monoethanolamine
N2 Nitrogen
N 2O Nitrogen (i) oxide
NO Nitrogen (ii) oxide
The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 243

NO2 Nitrogen (iv) oxide


[NO3−] Nitrate
PEG 200 Polyethylene glycol 200
[PF6−] Hexafluorophosphate
ROH Alkanol
RTILs Room temperature ionic liquids
SHAs Sterically hindered amines
SO2 Sulphur (iv) oxide
SO3 Sulphur (vi) oxide

1 Introduction

The removal of carbon dioxide (CO2) from process or flue gas streams is an
important step in many industrial processes for a number of reasons especially as it
raises environmental concern. CO2 capture from flue gas streams of process
industries is part of environmental issues of major concern in the world today as it is
the most important greenhouse gas (GHG) released in abundant quantity to the
atmosphere which contributes to global warming. The perceived consequences of
global warming have created alarming environmental worry over the reduction of
GHGs emission from industrial sources. Though the patent for application of
amines to industrially capture CO2 has been granted to Bottoms as early as 1930
and notwithstanding the improvements made on the technology till date, the process
still has a great setback due to high energy input and cost. In view of the global
outcry for effective control of GHGs emission to mitigate its global warming effect
there is still high need to search and obtain alternative CO2 capture system that will
be efficient and cost effective as CO2 is one of the main target gases. Therefore, this
book chapter provides a judicious practical application of this core subject. High
emphasis is placed on the aspects of chemical kinetics in relation to its practical
application in analyzing and solving real problems that form the foundation for the
practice of a chemical engineer and other related fields.
Section two of this chapter discusses briefly the term global warming, green-
house gases and sources of CO2 emission. Section three highlights CO2 capture
systems with concise explanation of pre-combustion, oxy-combustion and post
combustion CO2 capture processes. In addition, solvent scrubbing techniques or
CO2 capture technology were presented in section four amines—CO2 capture
process were explicitly discussed with additional information on sterically hindered
amines and blends of amines. Application of CO2 binding organic liquids, ionic
liquids and hybrid solvents for CO2 capture are treated. Finally, reaction mecha-
nisms of amines are explicitly discussed and kinetic equations of the various pro-
cesses and model are formulated.
244 O.Y. Orhan et al.

2 Global Warming and Sources of CO2 Emission

Global warming is real and it is fast affecting our planet with several evidences
around us. The major factor causing increased global warming comes from carbon
dioxide emission. Global warming doesn’t just mean that the earth gets hotter but
the whole climate is changing. It should be noted that naturally the average surface
temperature of the earth is a function of solar energy obtained from sun as a primary
source. Energy from the sun is sourced as heat and light to earth during the day
time. Some of the sun’s rays get ‘trapped’ in the atmosphere making the atmosphere
serve as heat store, this heat warms the earth at night which makes our planet warm
enough to live on. Some of the solar energy gets reflected back into space and the
earth also radiates heat into space, which cools it down. Thus, when the difference
in amount of energy absorbed from sun is much greater to the one reflected or
radiated into space from the earth surface, then the earth heats up and the climate
will change leading to global warming.

2.1 Greenhouse Gases

The gases in earth’s atmosphere that generate greenhouse effect are known as
greenhouse gases. Greenhouse effect is attributed to some gases present in the
atmosphere which causes absorption and emission of infrared radiation leading to
warming of the atmosphere. Examples of such gases include methane (CH4), car-
bon dioxide (CO2), nitrous oxide (N2O) and halogens namely chlorofluorocarbons
(CFCs) and hydrofluorocarbons (HFCs). The contribution of all the gases to the
overall greenhouse effect is based on its emission volume as well as their individual
greenhouse potentials.

2.2 Sources of CO2 Emission

The sources of CO2 emission comes from different categories ranging from low
concentration to high concentration by volume or mass of CO2 into the environ-
ment. High concentration emission sources are of greater interest to CO2 capture
researchers as it is a worry to the society. Among the high concentration sources of
carbon dioxide and other greenhouse gases released into the atmosphere include
emissions obtained from fossil fuels combustion, burning of agricultural areas or
forest. Other concentrated sources come from human activities such as burning of
coal, natural gas, and oil for generation of electricity.
The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 245

3 CO2 Capture Systems

It should be noted that the established ways or processes of capturing CO2 are termed
CO2 capture systems. The three basic processes in use involve pre-combustion, oxy-
and post-combustion capture. Fossil fuels combustion takes place as a set of reac-
tions between oxygen and hydrocarbons with CO2 as one of the products. The
exhaust gas that is produced resembles air but with a higher concentration of CO2.
The total amount of exhaust gas produced is very large and to store all of it is not an
option therefore, CO2 must be separated from other exhaust gas components.

3.1 Pre-combustion Process

Using pre-combustion carbon capture process, combustible fuel is first transformed


into synthetic gas (CO + H2) by gasification or partial oxidation. The synthetic gas
is furthermore reformed with steam to produce hydrogen and carbon dioxide. The
resultant product stream is then separated into a CO2 gas stream, and a stream of
hydrogen. Due to the high CO2 concentration, it is removed normally using
physical solvents.

3.2 Oxy-Combustion Capture

The application of oxy-combustion process to capture CO2 involves generation of


high concentrated CO2 flue gas stream by using pure oxygen instead of air mixture
for combustion of the primary fuel. The combustion products are mainly water
vapour and CO2. Cooling and compressing the gas stream will remove water
fraction of the product stream while further purification may be needed to remove
air pollutants and non-condensed gases such as nitrogen from the flue gas before the
CO2 is sent to storage.

3.3 Post-combustion Capture

Post-combustion capture refers to CO2 capture from flue gas streams resulting from
fuel combustion using aerial oxygen. Depending on the fuel used, the flue gas
contains mainly N2, steam, CO2, NO2, NO, SO2, SO3. The absorption using chemical
or physical solvent is in fact most well developed technology for post combustion
CO2 removal. The absorption processes are carried out by dissolving CO2 present in
a flue gas stream into organic solvents or simultaneous absorption and reaction into
aqueous base solutions (designated chemical absorption). The process selection
246 O.Y. Orhan et al.

depends mainly on the inlet CO2 partial pressure, degree of removal and also the
energy requirement. The energy requirements are dependent mainly on circulation
rate, temperature difference between the absorber and the regenerator and degree of
heat exchange and solvent chemistry.

4 Solvent Scrubbing Technology for CO2 Capture

There are some proved solvents that are applicable for solvent scrubbing technol-
ogy for CO2 capture. In this section, the focus will be on amines, CO2 binding
organic liquids (CO2BOLs), ionic liquids and hybrid solvents.

4.1 Amines

Amine technology is the most mature technology used to industrially capture CO2.
The basic process covering the application involves acid gas absorption from a gas
stream into an aqueous solution of an (alkanol) amine, which was patented as early
as 1930 by Bottoms [1]. The conventional amine solvents mostly used for CO2
capture are single amines.
In the absorption process, CO2 reacts with amines in aqueous form. Thus, the
chemical solvents do react with CO2 forming non-volatile ionic species; which
creates two main advantages over the physical solvents. It increases reaction
kinetics of CO2 and hence, enhances its solubility in water. Figure 1 illustrates the

Fig. 1 Basic flow scheme for CO2 absorption using chemical solvent [2]
The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 247

basic flow scheme and apparatus commonly used industrially for CO2 absorption
with chemical solvent [2].
Industrially, CO2 capture using a chemical solvent (amine) is commonly con-
ducted by passing the warm exhaust gas stream into the bottom of absorption
column. From base of the column, gas rises through it meeting a stream of
absorbing solution passed counter-current to the gas to absorb CO2. The CO2
absorbs and reacts with components in the solution, and the gas stream gradually
loses its CO2 while moving up the column.
At the top, gas with low CO2 content is released into the atmosphere. CO2
content of the solution increases as solution moves down the column. The liquid
stream is typically at 90–95% of equilibrium with incoming exhaust gas at bottom
of the column. At the bottom, rich concentrated CO2 solution is taken out and is
channeled to a second column termed stripper (desorber). In stripper, temperature
and/or pressure are set so that the chemical equilibrium in the liquid are reversed
and the CO2 is released into gas phase. Pressure release is very common in natural
gas applications whereas changing the temperature is the most common approach
for exhaust gas treatment. Change in temperature is usually achieved by adding heat
as steam in reboiler below the stripper column. A gas phase consisting only of CO2
and steam is taken out at top of the column. The steam is separated from CO2 in the
overhead condenser and CO2 can be compressed to facilitate its transportation and
storage. Regenerated chemical solvent from the stripper has low CO2 concentration
and is again recycled for CO2 absorption. The amine solvent recycled keeps cir-
culating between absorber and stripper columns, transporting CO2 between the
columns. In an industrial process, the absorber will often be operated at tempera-
tures around 40–55 °C while the stripper will be operating at around 120 °C.
Economically, operating conventional amine technology for CO2 capture is still at a
high cost notwithstanding the improvements made on the process since inception
till date. This is due to the associated practical problems like high energy demand
for its solvent regeneration, absorbent losses, and high corrosion rate limiting its use
at high amine concentration to increase reactivity. These setbacks result in high
operating costs. Hence, to reduce the operating cost, there is need to obtain and use
better solvents in the CO2 separation process. Many experimental works have been
published on various aspects of CO2 capture processes and therefore a brief
overview is given here on the nature of published data. A number of studies deal
with gas-liquid equilibrium experiments. The CO2 partial pressure and CO2 con-
centration is determined where the system is taken to be in a pre-fixed pressure and
temperature values. Obtained experimental results are then presented as plots of
CO2 partial pressure (kPa) versus CO2 uptake (loading) so as to determine CO2
solubility in a given amine. Another important form of experiment is the study of
kinetics. The experimental set-up for kinetics studies varies significantly but the
general approach is to measure the rate of CO2 uptake in a liquid at a given set of
conditions. The conditions set are usually temperature, pressure and liquid
composition.
248 O.Y. Orhan et al.

The CO2 uptake does not necessarily reflect a single rate of reaction. Some
analysis work is usually required to extract reaction kinetics data from the exper-
imental results.
The importance of chemical kinetics study with respect to reducing energy
demand by carbon capture process cannot be overemphasized. Reaction kinetics
gives us the insight of reaction rates of various solvents proposed for application in
carbon capture. It provides knowledge of affects of various variables in the system
thereby enabling one to determine the optimum condition for a targeted output with
minimum waste. Kinetic studies helps in many ways to determine energy efficiency
of carbon capture. Specifically, it provides details of reaction rates of various sol-
vents and mediums which enables one to make a choice of most efficient system to
be applied in the carbon capture process. The kinetic studies also relate effect of
variables, intermediates reaction with overall reaction mechanism that can generate
mathematical models that will describe chemical reactions. This makes it easier for
one to make choice of the most efficiency solvent option that can be applied for
CO2 capture with minimum energy input in CO2 absorption, stripping and regen-
eration of the solvent in use there by improving the economic efficiency of the
process.
In kinetics study, calorimetric experiments can also be used to obtain informa-
tion on the enthalpy of CO2 absorption. The same kind of measurements can also be
used to determine heat capacities. Many pure amines and amine-water systems have
available data of their measured physical properties. The extent of such data is
however more limited than for common organic molecules. The primary amine,
MEA is commonly used industrially for CO2 separation. MEA solution reacts fairly
fast and absorbs CO2 and H2S simultaneously. It reacts also with COS, CS2 and
mercaptans. The high vapour pressure property of monoethanolamine makes it a
good option for CO2 removal especially in flue gas stream with very low H2S
concentrations and without COS or CS2 [3]. The kinetic results for reaction of MEA
with CO2 in aqueous solution using a direct stopped flow technique compare
favourably with those obtained using indirect methods. Ali et al. [4], experimentally
studied kinetics of MEA with CO2 using stopped flow method, and the results they
obtained were in good agreement with similar experimental results published by
other Authors [5, 6].
Secondary amine like DEA helps overcome the limitation of MEA, and is used
in the presence of COS and CS2. DEA is less basic and its reaction rate does not
equal the performance of MEA, but it is easier to regenerate. Thus, it is still an
industrially accepted amine [7].
It should be noted that tertiary amines can result in several processing advan-
tages over the use of MEA and DEA. First of all, it can selectively remove H2S
from gas streams, because the reaction rate with CO2 is finite and slow [8]. Tertiary
amines do not form carbamate but contribute to the formation of bicarbonate
making it possible for its equilibrium to be easily reversed in the stripper. Because
amine-CO2 stoichiometry with respect to bicarbonate formation is 1 to 1, tertiary
amines do also have the potential to absorb higher amounts of CO2. However, they
tend to have low reaction rates. For natural gas treatment the tertiary amine
The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 249

N-methyldiethanolamine (MDEA) is widely used [9]. In industrial flue gases,


concentration of CO2 in the gas phase is lower thus MDEA is thought to have too
low reactivity to be efficiently applied in such a case. Tertiary amines are often
combined with promoters in order to take advantage of the shuttle-effect [10, 11].

4.2 Sterically Hindered Amines

Research interest in sterically hindered amines (SHAs) stimulates on its ability to


form unstable carbamates. These cabamate ions formed as SHAs react with CO2 are
unstable and so lead to formation of biocabonates and free amines.
2-amino-2-methylpropanol (AMP) is an example of SHAs in frequent use. It offers
a good alternative to conventional amines for CO2 capture as it has ability of higher
CO2 absorption capacity, favourable selectivity of CO2 in presence of gas mixtures
and degradation resistance advantages [12].
Reaction rate of aqueous AHPD—CO2 system has been reported though there is
discrepancy between the rates from other authors [13] but all the published data
gave low reaction rates indicating that sterically hindered AHPD will not be suitable
alone. It was therefore, noted that AHPD is a potential SHA as CO2 absorbent if the
reaction rate is enhanced by an additional amine such as piperazine. AHPD offers a
higher absorption capacity, as well as lower regeneration heat over conventional
amines [13].

4.3 Blends of Amines

In recent years, the usage of suitable amine mixtures over selection of single amines
has become a popular approach in gas treating processes. This is mainly due to the
high capacity coupled with better regeneration economy of blended amines over
single ones especially when blend involves tertiary amine or sterically hindered
amine with high reaction rates primary or secondary amines [14]. Alkanolamines
are primary, secondary or tertiary amines containing one or more hydroxyl mole-
cule(s). The amino group provides the basicity to absorb acid gases by a bronsted
type acid-base reaction, while the OH functional group brings about decrease in
vapour pressure of the alkanolamine leading to its higher solubility in water.
Stoichiometrically, the ratio of 2:1 absorption capacity is achieved for a reaction
between moles of primary or secondary amines and that of CO2 respectively, while
a loading of 1 or higher can be obtained in the case of tertiary amines. Blends of
primary or secondary alkanolamines with tertiary alkanolamines have become
common to utilize the high absorption rates of the former along with the high
loading capacity of the latter.
There is a vast range of possible amines that can be blended to achieve desired
properties but suitability of blending amines and its relative composition has to be
250 O.Y. Orhan et al.

determined experimentally to obtain the optimum blending ratio of two or more


solvents. It should also be noted that though blended amines give observed reaction
rate constant, ko which might be higher in value compared to addition of rate
constants, ko of the respective pure amines in the blend at its equivalent concen-
trations under the same conditions but generally, this is not true for all blends.
Notwithstanding the earlier challenges or problems associated with amine tech-
nology, it is worthy to note that the work on ultrasonic desorption of CO2 indicates
that the use of monoethanolamine system for CO2 capture can still stand a test of
time. They noted that the use of ultrasonic treatment of aqueous CO2-containing
amine solutions, in particular ethanol amines and blends of amines, generally leads
to an accelerated degassing of CO2 [15]. It revealed further that thermal degradation
is minor when the reboiler temperature is held below 110 °C. Based on this claim,
if amine solvents can be regenerated within temperature range of 60–80 °C from
the present 100–120 °C in practice, it implies significant energy savings and
reduction in thermal degradation effect. This will be an outstanding advantage for
amine solvents.

4.4 CO2 Binding Organic Liquids (CO2BOLs)

Recently, a completely different approach was proposed by Jessop et al. in order to


handle disadvantages of aqueous systems by switching to an organic base and
high-boiling liquid compounds [16].
CO2-binding organic liquids (CO2BOLs), which are also known as switchable
solvents, comprised of strong amidine/guanidine base and alcohols that capture CO2
to form amidinium or guanidinium alkyl carbonate salts. CO2BOLs can reversibly
switch from a non-polar form to a polar form when exposed to carbon dioxide and
causes a dramatic change in polarity (Fig. 2) [17]. Then, the high-polarity solvent
can revert back to its non-ionic form by the removal of CO2 from solution [18]. The
removal of CO2 is achieved by heating the solution below its boiling point or by
sweeping with an inert gas such as nitrogen. The main advantages of using
all-organic solvents are their lower cost of energy due to the elimination of the
vaporization of water, lower stripper reboiler temperature, and their tunable
physicochemical properties. Therefore, they have considerable potential to be an
efficient CO2 capture solvent due to high CO2 binding capacities, low heat capaci-
ties, less solvent loss during CO2 stripping and lower energy requirement for
regeneration than the traditional aqueous amine systems [19]. Consequently, there

Fig. 2 First generation of CO2BOL reacting with CO2. R = (CH2)nCH3 [17]


The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 251

are several studies examining the absorption performance and reversibility of


CO2BOL systems [16, 17, 19]. However, intrinsic reaction kinetics of CO2BOL
composed of a mixture of 1,8-diazabicyclo [5,4,0] undec-7-ene (DBU) amidine base
in 1-hexanol or 1-propanol were investigated by Ozturk et al. [20].
Further still, Ozturk et al. [20], studied the mechanism and kinetics of CO2
absorption in CO2-binding organic liquids (CO2BOLs) system. They noted that
1,8-Diazabicyclo[5.4.0]undec-7-ene (DBU):1-Alkanol mixtures are promising CO2
absorption systems with beneficial properties for industrial processes, such as lower
specific heats and almost complete regeneration with simple heating well below the
boiling point of the mixture [21]. They also stated that second order rate constants
of DBU system and its observed reaction order is very much competitive with
commonly used industrial CO2 absorption aqueous solvents such as AMP and
comparable to MEA and DGA. Therefore, with further advancement, CO2BOLs are
believed to be of future potential CO2 capture technology [22].

4.5 Ionic Liquids

Another alternative is the use of ionic liquids instead of the aqueous solution
[23, 24]. Ionic liquids are molten salts composed of a large organic cation such as
1-alkyl-3-alkyl imidazolium and a small inorganic anion such as tetrafluoroborate.
The melting point of most of the ionic liquids is below the room temperature.
Therefore, these liquids are generally known as room temperature ionic liquids
(RTILs). Ionic liquids seem to have a certain potential in separation processes due
to their low vapour pressure and selective solubility for CO2. Conceptually, CO2
can be selectively absorbed from the gas mixture at high pressure and low tem-
perature, and then can desorb at low pressure and high temperature. In such an
innovative process, as there is no need to CO2 dehydration, there is no loss of the
solution due to negligible vapour pressure. It is not energy-intensive for it did not
require evaporation. However, generally only physical absorption will exist with
existing ionic liquids, mass transfer (absorption) rate will be slow and solution
capacity will be limited by the physical solubility. The absorption capacity can be
enhanced by functional amine groups. However, this increases the viscosity and can
cause difficulties in operation [25]. Although, variety of cations and anions are
infinite that can create ionic liquids, commonly used anions and cations are shown
in Table 1.
Some of the possible cations; imidazolium, pyridinium, pyrrolidinium, phos-
phonium, ammonium or sulfonium. Possible anions; hexafluorophosphate [PF6−],
tetrafluoroborate [BF4−], bis (trifluoromethylsulfonyl) imide [(CF3SO2)2 N−], tri-
flate [CF3SO3−], acetate [CH3CO2−], trifluoroacetate [CF3CO2−], dicyanamide
[(CN)2 N−], nitrate [NO3−], chloride [CI−], bromide [BR−] or iodide [I−]. The
chemical structure of [EMIm]+ [Tf2N] is shown in Fig. 3.
252

Table 1 Some anions and cations constituting the ionic liquids


Cation (organic) Anion (organic) Anion (inorganic)
Imidazolium Alkylsulfate Bis(trifluoromethylsulfonyl)
imide

Pyridinium Tosylate Hexafluoro-phosphate

Pyrrolidinium Methanesulfonate Tetrafluoro-borate

Phosphonium Halide

Ammonium

Sulfonium
O.Y. Orhan et al.
The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 253

Fig. 3 Chemical structure of [EMIm]+ [Tf2N]−

There is ongoing interest in ionic liquids because of their advantages such as


high thermal stability, negligible vapour pressure, adjustable physicochemical
characteristics and high CO2 loading capacity [24, 26]. The major disadvantage of
ionic liquids is the high viscosity as previously mentioned. The viscosity of the
ionic liquids is determined by Van der Waals forces and hydrogen bonds. However,
the viscosity can be adjusted within an acceptable range, such as 50–10.000 cP, by
selecting a suitable combination of cations and anions [25].
For imidazolium-based cations, viscosity of ionic liquids depends on the length
of the alkyl chain as well as upon the nature of the anion. An increase in viscosity is
clearly stated in Table 2 with increasing alkyl chain lengths, because of the
increased probability of van der Waals interactions between cations.
Experimental and simulation studies have shown that CO2 is more soluble in
alkyl imidazolium-based ionic liquids. The reason for this high solubility is that
anion dominates the interactions of CO2 in ionic liquids and cation plays secondary
role [26]. The length of alkyl-side chain of imidazolium based cations also affects
the CO2 solubility (Table 1). Fluoro-substituted side chains increase CO2 uptake
greatly rather than fluorine-nonsubstituted side chains but substantially this causes

Table 2 Viscosity of various ionic liquids


Ionic liquid Viscosity (cP) Temperature (°C) Source
[C2mim][BF4] 43 25 [27]
[C2mim][PF6] 23.4 70 [28]
[C2mim][Tf2N] 32.6 25 [27]
[C2mim][OTf] 50 20 [28]
[C4mim][PF6] 450 25 [29]
[C4mim]Cl 1534 50 [28]
[C4mim][BF4] 219 25 [29]
[C4mim][I] 1110 25 [29]
[C4mim][Tf2N] 69 25 [29]
[C6mim][PF6] 585 25 [29]
[C6mim]Cl 716 25 [29]
[C6mim][BF4] 314 20 [28]
[C8mim]Cl 337 25 [29]
[C8mim][PF6] 682 25 [29]
254 O.Y. Orhan et al.

Table 3 Henry constant at different temperatures for carbon dioxide in ionic liquids [33]
Ionic liquid HCO2 (bar)
10 °C 20 °C 25 °C 30 °C 40 °C 50 °C
[bmim][Tf2N] 28 ± 2 30.7 ± 0.3 34.3 ± 0.8 42 ± 2 45 ± 3 51 ± 2
[bmim][BF4] 41.9 ± 0.2 52 ± 2 56 ± 2 63 ± 2 73 ± 1 84 ± 4

increase in the viscosity [30–32]. Anion appears to be of more powerful effect on


the gas solubility than cations. Imidazolium-based ionic liquids containing [Tf2N]
anions show higher CO2 solubility (Table 3).
Presence of long alkyl chains on the cation of the ionic liquid can create steric
hindrance between the CO2-cation interactions. When the length of the alkyl side
chain on the cation is increased, cation-anion interactions decrease and an increase
in CO2 solubility generally results because of the increased free volume available
for CO2 (Tables 3 and 4) [33, 34].

Table 4 Henry constant for carbon dioxide in different ionic liquids


Ionic liquid Abbreviation HCO2 Source
(bar)
1-ethyl-3-methylimidazolium bis [emim][Tf2N] 35.6 [26]
(trifluoromethanesulfonyl)imide
1-butyl-3-methylimidazolium bis [bmim][Tf2N] 33.0 [26]
(trifluoromethanesulfonyl)imide
1-hexyl-3-methylimidazolium bis [hmim][Tf2N] 31.6 [26]
(trifluoromethanesulfonyl)imide
1-butyl-3-methylimidazolium hexafluorophosphate [bmim][PF6] 53.4 [35]
1-ethyl-3-methylimidazolium tetrafluoroborate [emim][BF4] 80.0 [36]
1-butyl-3-methylimidazolium tetrafluoroborate [bmim][BF4] 59.0 [35]
1-methyl-3-(3,3,4,4,5,5,6,6,6-nonafluorohekzil) [C6H4F9mim] 28.4 [26]
imidazolium bis(trifluoromethanesulfonyl)imide [Tf2N]
1-methyl-3- [C8H4F13mim] 27.3 [26]
(3,3,4,4,5,5,6,6,7,7,8,8,8-tridecafluorooctyl) [Tf2N]
imidazolium bis(trifluoromethanesulfonyl)imide
1-hexyl-3-methylimidazolium bis [hmpy][Tf2N] 32.8 [26]
(trifluoromethanesulfonyl)imide
1-hexyl-3-methylimidazolium tris(pentafluoroetil) [hmim][eFAP] 25.2 [26]
trifluorofosfat
1-hexyl-3-methylimidazolium tris(heptafluoropropil) [hmim] 21.6 [26]
trifluorophosphate [pFAP]
1-butyl-2,3-methylimidazolium tetrafluoroborate [bmmim] 61.0 [26]
[BF4]
1-ethyl-2,3-methylimidazolium bis [emmim] 39.6 [26]
(trifluoromethanesulfonyl)imide [Tf2N]
1-ethyl-3-methylimidazolium [emim][OTf] 73 [37]
trifluoromethanephosphate
The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 255

Viscosity of the commonly used ionic liquids is very high at room temperature.
At the same temperature (33 °C), the viscosity of [bmim] [BF4] (79.5 cP) is 40 times
higher than that of the 30% mass MEA solution [38]. To cope with this limitation of
viscosity of ionic liquids, they can be mixed with water or ordinary organic solvents
[39]. However, the addition of such liquids causes the reduction of gas capture
ability. Gas solubility reduces with increasing amount of organic liquid [eg. poly-
ethylene glycol 200 (PEG)]. This phenomenon can be explained by the low solu-
bility of CO2 in PEG 200. It was determined that by adding an appropriate amount of
PEG 200, both desorption rate and absorption can be increased. This case can be
explained by the PEG 200 solvent properties or decrease in viscosity [40].
Another more viable option is the combination of non-volatile and stable ionic
liquids with alkanolamine systems. Thus, negligible vapour pressure, high thermal
stability and low heat capacity of ionic liquids can be combined with fast absorption
kinetics of alkanolamines in the amine functional ionic liquids [41]. CO2 absorption
capacity of the ionic liquids functionalized with suitable amino groups (“Task
specific”) are three times higher than room temperature ionic liquids (RTILs)
[39, 42]. Chemical and physical absorption increase with using task specific ionic
liquid while working at high pressure. The task specific ionic liquids when exposed
to CO2 for 3 h at room temperature and pressure, the mass gain was 7.4% which
corresponds to 0.5 molar uptake of CO2. They remain stable even after five gas
absorption/desorption cycles without any detectable loss in efficiency [25].

4.6 Hybrid Solvents

Separation of carbon dioxide from gas mixture by absorption is most widely used
method. Conventional aqueous alkanolamine solutions used in “absorber-stripper”
system have some problems such as low capacity (leading to high circulation rate)
and high cost of energy in “boiler”. Capacity problems can be overcome by using
hindered amines which have unstable carbamate ions but reversible reaction
requires temperatures such 120–130 °C which leads to evaporation losses and high
energy demand. The previous studies were to improve absorption rate and regen-
eration capacity of widely used alkanolamine solution. For this purpose, the
advantages of the various amine reaction mechanisms were combined with acti-
vators, such as piperazine, imidazolium were added to the amine solution. Low
production costs and high reaction rates of primary and secondary amines; high
absorption capacity and less energy demand during regeneration due to their low
heat of reaction of tertiary the amine are advantageous parameters to mix them.
Because of these reasons, use of ionic liquids which remain liquid at room
temperature and have high boiling point (at least 200 °C) has been raised.
Commercial ionic liquids pass through several synthesis and purification steps so
they are not as cheap as amines. Another limitation is their high viscosity. Solvents
with high viscosity often result in high energy consumption during absorption.
Various disadvantages of this kind of solvents used in carbon dioxide capture lead
256 O.Y. Orhan et al.

to high interest for new hybrid solvents that can combine the advantages of different
amine systems. Therefore, a new solvent formulation which is superior in terms of
energy requirement has been developed. The addition of ionic liquid to CO2BOLs
is expected to bring high thermal stability and reduce the effect of volatile alcohol
(eg. Hexanol) in the CO2BOL. The preparation of hybrid solutions are more
practical and quick than adding functional amine groups to ionic liquids and they do
not require high synthesis costs. Furthermore, the existing viscosity problems of
task specific ionic liquids (TSIL) will not be encountered [43].
In this chapter, CO2-binding organic liquid—ionic liquid hybrid solvents have
been proposed as good alternatives to existing alkanolamine solutions [44]. Hybrid
solvents composed of such advantageous of CO2BOLs (high CO2 loading capac-
ities, low specific heat capacities and a less energy intensive solvent regeneration)
and ionic liquids (have negligible vapour pressures, high thermal stabilities,
favorable CO2 solubilities and can remain liquid over a large temperature range
with tunable physicochemical properties) provide to increase the absorption
capacity and reaction rate. Although, there have been several studies which showed
the synergetic effect on CO2 capture of ionic liquids–amine solutions there is
basically no information about the reaction kinetics of ionic liquid—CO2BOL
systems [45–49].

5 Reaction Mechanisms

Generally, the reaction of carbon dioxide with amines can be described by two
established mechanisms. These mechanisms are known as zwitterion mechanism
and termolecular mechanism. The zwitterion mechanism has originally proposed by
Caplow [50] and then reintroduced by Danckwerts [51] while termolecular
mechanism has proposed by Crooks and Donnellan [52] and supported by Alper
[53]. The zwitterion mechanism consists of a two-step process. In the first step, a
zwitterion intermediate is formed rather than one-step carbamate formation. In the
next step, deprotonation of the zwitterion takes place to produce carbamate ion and
a protonated base.
Zwitterion formation for a primary amine is as follows:

k2
CO2 þ RNH2 ! RN þ H2 COO ð1Þ
k1

Deprotonation by any base (B) present in the solution

RN þ H2 COO þ B ! kB RNHCOO þ BH þ ð2Þ

In this reaction, an amine or an alcohol can act as the base depending on whether
amine or alcohol is carrying out the deporotonation. The resulting net reaction is
given in Eq. (3).
The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 257

CO2 þ 2 RNH2  RNHCOO    RNH3þ ð3Þ

In a termolecular reaction mechanism, a molecule of amine (R2NH) reacts with a


molecule of carbon dioxide and a molecule of a base (B) to form a loosely-bound
encounter complex in a single step as represented by Eq. (4).

CO2 þ RNH2    B  RNHCOO    BH þ ð4Þ

In this mechanism the bonding between amine and CO2, and the proton transfer
take place simultaneously (Fig. 4).
While a portion of the weakly bound intermediates is transforming to the
reactants (carbon dioxide and amine), even a small portion is reacted with one more
amine molecule and ionic products form. The observed rate equation with respect to
CO2 absorption is given in Eq. (5).

robs ¼ ko ½CO2  ð5Þ

According to termolecular reaction mechanism, Eq. (6) is valid for the


pseudo-first-order rate constants of hybrid systems consisting of ionic liquid
(IL) and amidine (or guanidine)/linear alcohol ((B)/n-hexanol).

ko ¼ kOH ½OH ½IL þ k0OH ½OH½B þ kIL ½IL½IL þ kB ½B½B þ kILB ½IL½B ð6Þ

Considering that the alcohol is in excess and nearly at constant concentration,


k = kOH [OH] and k ¼ k0OH [OH] are assumed so Eq. (7) is obtained.

ko ¼ ðk þ kIL ½ILÞ½IL þ ðk þ kB ½BÞ½B þ kILB ½IL½B ð7Þ

In the experiments carried out with stopped-flow technique, [IL] is kept constant
and [B] is changed. When [IL] is kept constant at [IL]0, the following equations are
obtained:

ko ¼ k þ kIL ½IL0 ½IL0 þ ðk þ kB ½BÞ½B þ kILB ½IL0 ½B ð8Þ

ko ¼ k4 þ k3 ½B þ kA ½B½B ð9Þ

Fig. 4 Single-step reaction mechanism [52]


258 O.Y. Orhan et al.

where, k4 = (k + kIL [IL]0) [IL]0 and k3 = k* + kIL-B[IL]0, are fixed under exper-
imental conditions.
The kinetic equations valid for the amine mixtures given by Eqs. (6)–(9) are
original and have not been previously proposed in the literature. In this study, the
rate constants for hybrid mixtures were obtained using Eq. (9).
If the degree of the reaction according to the amine is 1.00 under constant ionic
liquid concentration, based on termolecular reaction mechanism, pseudo first order
observed rate constant is given by Eq. (10).

ko ¼ k½B ð10Þ

According to Eq. (10), if ko (s−1) is plotted versus [B], the rate constant
k (m3/kmol.s) is obtained from the slope by linear regression.
If the reaction order is fractional between 1.00 and 2.00, according to ter-
molecular reaction mechanism, affecting base can be [B] or 1-hexanol and observed
pseudo first order rate constant can be defined as:

ko ¼ k½B þ kB ½B2 ð11Þ

According to Eq. (11), if ko (s−1) is plotted versus [B], the rate constants
k (m3/kmol.s) and kB (m6/kmol2.s) are obtained from the slope by polynomial
regression.
If the degree of reaction by the amine is accepted 2.00, according to termoleculer
and zwitterion mechanisms the base which affected the reaction is considered
probably of that amine. The effect of hexanol as a base to the reaction is negligible
beside amines. The obtained pseudo first order reaction rate constant equation
simplifies to:

ko ¼ kB ½B2 ð12Þ

According to Eq. (12), if ko (s−1) is plotted versus ([B]2), the rate constant kB
(m /kmol2.s) is obtained from the slope by linear regression.
6

The zwitterion mechanism becomes equivalent to the termolecular mechanism


and gives rise to similar expressions for reaction kinetics when the lifetime of the
zwitterions intermediate approaches zero.

References

1. Astarita G (1983) Gas treating with chemical solvents. Wiley-Interscience Publication,


New York
2. Kohl A, Nielsen R (1997) Gas purification, 5th edn. Gulf Publishing Co., Texas
3. Li J. (2006) Kinetic study of CO2 and alkanolamines both in aqueous and non-aqueous
solutions using the stopped flow technique. Master’s Thesis, University of Regina, Canada
The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 259

4. Ali SH, Merchant SQ, Fahim MA (2002) Reaction kinetics of some secondary alkanolamines
with carbon dioxide in aqueous solutions by stopped flow technique. Sep Purif Technol
27:121–136
5. Alper E (1990) Kinetics of reactions of carbon dioxide with diglycolamine and morpholine.
Chem Eng J 44:107–111
6. Mimura T, Suda T, Iwaki I, Honda A, Kumazawa H (1998) Kinetics of reaction between
carbon dioxide and sterically hindered amines for carbon dioxide recovery from power plant
flue gases. Chem Eng Commun 170:245–260
7. Ali SH (2004) Kinetic study of the reaction of diethanolamine with carbon dioxide in aqueous
and mixed solvent systems: application to acid gas cleaning. Sep Purif Technol 38:281–296
8. Versteeg GF, van Swaaij WPM (1988) On the kinetics between CO2 and alkanolamines both
in aqueous and non-aqueous solutions-I. Primary and secondary amines. Chem Eng Sci
43:573–585
9. de Koeijer G, Solbraa E. (2004) High pressure gas sweetening with amines for reducing CO2
emissions. 7th international conference on greenhouse gas control technologies. Vancouver,
Canada
10. Bishnoi S, Rochelle GT (2002) Thermodynamics of piperazine/methyldiethanolamine/
water/CO2. Ind Eng Chem Res 41:604–612
11. Zhang X, Wang J, Zhang CF, Yang YH, Xu JJ (2003) Absorption rate into a MDEA aqueous
solution blended with piperazine under a high CO2 partial pressure. Ind Eng Chem Res
42:118–122
12. Sartori G, Savage DW (1983) Sterically hindered amines for CO2 removal from gases. Ind
Eng Chem Fundam 22:239–249
13. Ume CS, Alper E (2012) Reaction kinetics of carbon dioxide and 2-amino-2-hydroxy-
methyl-1,3-propanediol in aqueous solution obtained from stopped flow method. Turk J
Chem 36:427–435
14. Ali SH (2005) Kinetic of the reaction of carbon dioxide with blends of amines in aqueous
media using the stopped-flow technique. Int J Chem Kinet 37(7):391–405
15. Gantert S, Möller D (2012) Ultrasonic desorption of CO2: a new technology to save energy
and prevent solvent degradation. Chem Eng Technol 35(3):576–578
16. Jessop PG, Heldebrant DJ, Li X, Eckert CA, Liotta CL (2005) Green chemistry: reversible
nonpolar-to-polar solvent. Nature 436:1102
17. Jessop PG, Mercer SM, Heldebrant DJ (2012) CO2-triggered switchable solvents, surfactants,
and other materials. Energy Environ Sci 5(6):7240–7253
18. Yu T, Gaviola GC, Weiss RG (2008) Carbon dioxide and molecular nitrogen as switches
between ionic and uncharged room-temperature liquids comprised of amidines and chiral
amino alcohols. Chem Mater 20:5337–5344
19. Mathias PM, Afshar K, Zheng F, Bearden MD, Freeman CJ, Andrea T, Koech PK,
Kutnyakov I, Zwoster A, Smith AR, Jessop PG, Nik OG, Heldebrant DJ (2013) Improving
the regeneration of CO2-binding organic liquids with a polarity change. Energy Environ Sci 6
(7):2233–2242
20. Ozturk MC, Ume CS, Alper E (2012) Reaction mechanism and kinetics of 1,8-Diazabicyclo
[5.4.0]undec-7-ene and carbon dioxide in alkanol solutions. Chem Eng Technol 35(12):
2093–2098
21. Heldebrant DJ, Koech PK, Rainbolt JE, Zheng FR (2011) CO2-binding organic liquids: An
integrated acid gas capture system. Energy Procedia 4:216–223
22. Yüksel Orhan Ö, Ozturk MC, Seker A, Alper E (2015) Kinetics and performance studies of a
switchable solvent TMG (1,1,3,3-tetramethylguanidine)/1-propanol/carbon dioxide system.
Turk J Chem 39:13–24
23. Anthony JL, Maginn EJ, Brennecke JF (2002) Solubilities and thermodynamic properties of
gases in the ionic liquid 1-n-butyl-3-methylimidazolium hexafluorophosphate. J Phys Chem B
106:7315–7320
24. Muldoon MJ, Aki SN, Anderson VK, Dixon JL, Brennecke JK (2007) Improving carbon
dioxide solubility in ionic liquids. J Phys Chem B 111:9001–9009
260 O.Y. Orhan et al.

25. Hasib-ur-Rahman M, Siaj M, Larachi F (2010) Ionic liquids for CO2 capture-development
and progress. Chem Eng Process 49(4):313–322
26. Cadena C, Anthony JL, Shah JK, Morrow TI (2004) Why is CO2 so soluble in imidazolium-
based ionic liquids. J Am Chem Soc 126:5300–5308
27. McEwen AB, Ngo HL, LeCompte K, Goldman JL (1999) Electrochemical properties of
imidazolium salt electrolytes for electrochemical capacitor applications. J Electrochem Soc
146:1687–1695
28. Seddon KR, Stark A, Torres MJ (2000) Influence of chloride, water, and organic solvents on
the physical properties of ionic solids. Pure Appl Chem 72:2275–2287
29. Huddleston JG, Visser AE, Reichert WM, Willauer HD, Broker GA, Rogers RD (2001)
Characterization and comparison of hydrophilic an hydrophobic room temperature ionic
liquids incorporating the imidazolium cation. Green Chem 3:156–164
30. Shin EK, Lee BC (2008) High-pressure phase behavior of carbon dioxide with ionic liquids,
1-alkyl-3-methylimidazolium trifluoromethanesulfonate. J Chem Eng Data 53:2728–2734
31. Palgunadi J, Kang JE, Nguyen DQ, Kim JH, Min BK, Lee SD, Kim H, Kim HS (2009)
Solubility of CO2 in dialkylimidazolium dialkylphosphate ionic liquids. Thermochim Acta
494:94–98
32. Baltus RE, Culbertson BH, Dai S, Luo H, DePaoli DW (2004) Low-Pressure solubility of
carbon dioxide in room-temperature ionic liquids measured with a quartz crystal microbal-
ance. J Phys Chem B 108:721–727
33. Hou Y, Baltus RE (2007) Experimental measurement of the solubility and diffusivity of CO2
in room-temperature ionic liquids using a transient thin-liquid-film method. Ind Eng Chem
Res 46:8166–8175
34. Shiflett MB, Yokozeki A (2005) Solubilities and diffusivities of carbon dioxide in ionic
liquids: [Bmim][PF6] and [Bmim][BF4]. Ind Eng Chem Res 44:4453–4464
35. Anthony JL, Anderson JL, Maginn EJ, Brennecke JF (2005) Anion effects on gas solubility in
ionic liquids. J Phys Chem B 109:6366–6374
36. Camper D, Becker C, Koval C, Noble R (2005) Low pressure hydrocarbon solubility in room
temperature Ils containing imidazolium rings interpreted using regular solution theory. Ind
Eng Chem Res 44:1928–1933
37. Scovazzo P, Camper D, Kieft J, Poshusta J, Koval C, Noble R (2004) Regular solution theory
and CO2 gas solubility in room temperature ionic liquids. Ind Eng Chem Res 43:6855–6860
38. Ion IH, Muhammad R, Faiçal L (2014) CO2 absorption in diethanolamine/ionic liquid
emulsions—chemical kinetics and mass transfer study. Chem Eng J 240:16–23
39. Carvalho PJ, Alvarez VH, Schröder B, Gil AM, Marrucho IM, Aznar M, Santos MNBF,
Countinho JAP (2009) Specific solvation interactions of CO2 on acetate and trifluoroacetate
imidazolium based ionic liquids at high pressures. J Phys Chem B 113:6803–6812
40. Li X, Hou M, Zhang Z, Han B, Yang G, Wang X, Zou L (2008) Absorption of CO2 By ionic
liquid/polyethylene glycol mixture and the thermodynamic parameters. Green Chem 10:
879–884
41. Camper D, Bara JE, Gin DL, Noble RD (2008) Room-temperature ionic liquid-amine
solutions: tunable solvents for efficient and reversible capture of CO2. Ind Eng Chem Res
47:8496–8498
42. Carvalho PJ, Alvarez VH, Marrucho I, Aznar M, Coutinho JAP (2009) High pressure phase
behavior of carbon dioxide in 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide
and 1-butyl-3-methylimidazolium dicyanamide ionic liquids. J Supercrit Fluids 50:105–111
43. Ahmady A, Hasmin A, Aroua M (2010) Experimental investigation on the solubility and
initial rate of absorption of CO2 in aqueous mixtures of methyldiethanolamine with the ionic
liquid 1-butyl-3-methylimidazolium tetrafluoroborate. J Chem Eng Data 55(12):5733–5738
44. Arslan B, Şeker A, Alper E (2012) İyonik Sıvı—Karbon Dioksit Tutan Organik Sıvı Hibrit
Sistemleri ile Karbon Dioksit Yakalanması. Ulusal Kimya Mühendisliği Kongresi. 3-5 Eylül,
Türkiye
The Absorption Kinetics of CO2 into Ionic Liquid—CO2 … 261

45. Ahmady A, Hashim MA, Aroua MK (2011) Absorption of carbon dioxide in the aqueous
mixtures of methyldiethanolamine with three types of imidazolium-based ionic liquids. Fluid
Phase Equilibr 309(1):76–82
46. Ahmady A, Hashim MA, Aroua MK (2012) Kinetics of carbon dioxide absorption into
aqueous MDEA+[bmim][BF4] solutions from 303 to 333K. Chem Eng J 200:317–328
47. Lu BH, Jin JJ, Zhang L, Li W (2012) Absorption of carbon dioxide into aqueous blend of
monoethanolamine and 1-butyl-3-methylimidazolium tetrafluoroborate. Int J Greenhouse Gas
Control 11:152–157
48. Lv B, Sun C, Liu N, Li W, Li S (2015) Mass transfer and kinetics of CO2 absorption into
aqueous monoethano)lamine/1-hydroxyethy-3-methyl imidazolium glycinate solution. Chem
Eng J 280:695–702
49. Iliuta I, Hasib-ur-Rahman M, Larachi F (2014) CO2 absorption in diethanolamine/ionic liquid
emulsions—chemical kinetics and mass transfer study. Chem Eng J 240:16–23
50. Caplow M (1968) Kinetics of carbamate formation and breakdown. J Am Chem Soc
90:6795–6803
51. Danckwerts PV, Kennedy AM (1954) Kinetics of liquid-film processes in gas absorption, Part
II: measurements of transient absorption rates. Trans Inst Chem Eng 32:53–59
52. Crooks JE, Donnellan JP (1989) Kinetics and mechanism of the reaction between carbon
dioxide and amines in aqueous solution. J Chem Soc Perkin Trans 1989:331–333
53. Alper E (1990) Reaction-mechanism and kinetics of aqueous-solutions of 2-amino-2-methyl-
1-propanol and carbon-dioxide. Ind Eng Chem Res 29:1725–1728
Solubility of Carbon Dioxide in Aqueous
Solutions of Linear Polyamines

Jian Chen, Ruilei Zhang, Zhongjie Du and Jianguo Mi

Abstract For the analysis of energy consumption for carbon dioxide capture pro-
cesses from flue gases, CO2 solubility in aqueous amine solutions of various amines
at different temperatures and pressures are crucial. In this work, solubility of CO2 in
aqueous solutions of five linear polyamines were determined at 313.15 and 393.15 K
and CO2 partial pressure of about 1–500 kPa using the constant-volume method
combined with gas chromatography analysis. The amines are diethylenetriamine,
dipropylenetriamine, trimethylenediamine, tetraethylenepentamine, triethylene-tet-
ramine. The relationship between molecular structure of these polyamines and
capture performance is discussed. The results show that the capture performance is
affected by the species and number of amino groups, the carbon number between the
amino groups, and the chain length. The corresponding values of CO2 absorption
reaction heat were estimated employing the Gibbs-Helmholtz equation and were also
discussed with the molecular conformations. Compared with original solvents,
polyamines are more energy efficient solvents.

1 Introduction

Global warming and climate change associated with increased emissions of CO2
from anthropogenic sources have become one of the most critical worldwide issues
of the current age. A multitude of technological advances have been developed to
reduce CO2 emissions from combustion exhaust gases [1–3]. Nowadays
amine-based aqueous solutions are the most commonly used absorbents for the
absorption process [4]. In this case, finding new solvents that offer highly cyclic
absorption capacity for CO2 and less energy consumption for solvent regeneration

J. Chen (&)  R. Zhang


State Key Laboratory of Chemical Engineering, Tsinghua University, Beijing 100084, China
e-mail: cj-dce@tsinghua.edu.cn
R. Zhang  Z. Du  J. Mi
State Key Laboratory of Organic-Inorganic Composites, Beijing University of Chemical
Technology, Beijing 100029, China

© Springer International Publishing AG 2017 263


W.M. Budzianowski (ed.), Energy Efficient Solvents for CO2 Capture
by Gas–Liquid Absorption, Green Energy and Technology,
DOI 10.1007/978-3-319-47262-1_12
264 J. Chen et al.

is important. For the selection of better energy efficient solvents, CO2 solubility data
measurements in new solvents are crucial.
The first used amine solvent is methanolamine (MEA). MEA is a primary amine
and a strong base. For CO2 absorption, it has advantages as high absorption ability,
high reaction rate and high absorption ratio. But it also has disadvantages as high
regeneration heat, high degradation rate and serious facility corrosion. The study on
CO2 solubility in aqueous MEA solutions show that CO2 loading (mole CO2/mole
MEA) is about 0.5–0.6 at the conditions for absorption, while at the conditions of
regeneration the CO2 loading is about 0.15–0.25 [5–7]. So the circle loading is about
0.35. Cyclic absorption capacity is the solubility difference between absorption and
desorption. Another kind of solvents are secondary amines, as diethanolamine
(DEA) [8], diglycolamine (DGA) [9] and diisoproanolamine (DIPA) [10].
Compared with MEA, these secondary amines have advantages as high boiling
temperature and less degradation, with the similar absorption ability with MEA.
Their disadvantage is higher molecular weights, which leads more heat capacity.
And tertiary amines have advantages as high CO2 solubility at high CO2 partial
pressure and less degradation. The typical solvent is N-methyl-diethanolamine
(MDEA), which is widely used in acid gas removal from high pressure gases. The
circle loading at high CO2 partial pressure can be higher as 0.65 [11, 12], which is
higher than those of primary and secondary amines. Nowadays another kind of
amine solvents are steric-hindred amines, as 2-amino-2-methyl-1-propanol (AMP),
with its CO2 absorption mechanism is mainly bicarbonate over carbarmate [13]. So
its CO2 loading is close to 0.8, and circle loading is as more as 0.6. Chakraborty et al.
[14], Sartori and Savage [15], and Singh et al. [16] studied the steric-hindered
effects, and pointed out that these amines have a-substituted groups, and because of
hydrolysis there are about 1 mol of free amines liberated [17], leads higher con-
centration of free amines to react with more CO2. Based on AMP, for increase of
solvent boiling points, and decrease of solvent evaporation lose,
amino-methyl-propanediol and amino-ethyl-propanediol are used to absorb CO2, but
the second hydroxyl group leads lower CO2 solubility [18, 19]. Cycle amines are
always the research direction, and the typical one is piperazine (PZ). The CO2
solubility in aqueous PZ solution has been reported [20], and CO2 loading at usual
CO2 partial pressure is as more as 0.8, the circle loading is about 0.5. At high CO2
partial pressure, CO2 loading is more than 1.0 [21]. In multi-amino solvents,
amino-ethyl- ethanolamine (AEEA) is the amine studied frequently [22]. Compared
with MEA, it has advantages of high boiling temperature, and low solvent evapo-
ration loss, with CO2 solubility about 20 % higher than in MEA.
In order to find absorption solvent with low energy consumption, relationship of
molecular structure with CO2 absorption ability has been studied, e.g., chain length
of alkanolamines, chain lengths of alkylamines, chain length of diamines, and also
effects of side chain number and amino numbers [16]. It is also found that increase
of chain length between hydroxyl group and amino group can increase the CO2
solubility in solvents [23]. The key point was pointed out that the electronegativity
Solubility of Carbon Dioxide in Aqueous Solutions … 265

of the nitrogen atom in alkanolamines is the most important effects for molecular
structure on CO2 absorption ability [24].
For different requirements of CO2 absorption amounts, kinetic rates, mixtures of
amines have been used and CO2 solubility measured in these mixed amines,
including MEA-MDEA [25, 26], MEA-AMP [27], DEA-AMP [28, 29], AMP-PZ
[30], MDEA-PZ [31–33] and AMP-MDEA [34]. These mixed amine solutions are
mainly mixtures of primary or secondary amines with tertiary or steric-hindered
amines, for overall consideration of solubility and kinetic rate.
Compared with monoamine solvents, polyamines [35] solvents are expected to
have lower volatility along with higher CO2 loading capacity and mass transfer rate.
For example, diethylenetriamine (DETA) entails a higher mass transfer rate [36],
higher cyclic capacity [37–39], and significant lower heat of absorption [40] than
monoethanolamine (MEA). tetraethylenepentamine (TEPA) shows outstanding
potential for CO2 absorption, and it can maintain a high absorption rate as well as
cyclic absorption capacity [41]. 1.0 mol TEPA removes three times more CO2 per
cycle than 1.0 mol MEA. These investigations reveal that polyamines make a good
prospect for chemical absorption. Singh et al. showed that an increase in chain
length between the amine and different functional groups, result in a decrease of
absorption rate whereas, the absorption capacity was increased in most absorbents
[42, 43]. Machida et al. observed that alkyl chain length between two amines has an
important role in CO2 solubility [44].
Although some fragment solubility data for CO2 in polyamines have been
reported in previous literature, the solubility data for CO2 in polyamines in different
structures are insufficient. In this work, the vapor–liquid equilibrium data of CO2 in
ten polyamine solutions were measured at 313.15 K (absorption process) and
393.15 K (desorption process). Based on the solubility data, the corresponding
absorption heats (DHabs ) were calculated to analyze the energy consumptions for
absorbent regeneration. These data could be used for detailed analysis on the
performance of these solvents and on possible energy consumption for CO2
capture.

2 Measurement

2.1 Materials

CO2 and N2 with a volume fraction of 0.99999 were supplied by Millennium City
Gas. The solvents used are listed in Table 1, and their molecular structures are
shown in Table 2. The solution was prepared using ultrapure water, which was
taken from the Center 120 FV-S Ultrapure Water Machine. The resistivity of
ultrapure water is 18.2 MXcm at 298.15 K. All components were used without
further purification.
266 J. Chen et al.

Table 1 Chemicals used in this work


Component Abbreviation Molecular CAS Purity Source
formula number (%)
Diethylenetriamine DETA C4H13N3 111-40-0 99 J&KScientific
Dipropylenetriamine DPTA C6H17N3 56-18-8 >98 TCI Shanghai
Trimethylenediamine TMDA C3H10N2 109-76-2 98 Alfa Aesar
Tetraethylenepentamine TEPA C8H23N5 112-57-2 >90 Sinopharm
chemical
Triethylenetetramine TETA C6H18N4 112-24-3 >70 Sinopharm
chemical

2.2 Apparatus and Procedure

The constant-volume method, combined with gas chromatography analysis was


used to measure the differing levels of solubility. A detailed description of the
vapor–liquid equilibrium apparatus can be obtained from previous literature [45,
46]. It is comprised of two stainless steel tanks for buffer and reaction. Both tanks
are equipped with temperature transducers (PT-100, Kunlunhaian Co.) and pressure
transducers (JYB-KO-HAG, Kunlunhaian Co.). The accuracy of the temperature
and pressure transducers is ±0.1 K and ±0.5 %, respectively. In accordance with
the Peng-Robinson (PR) equation, the precise amount of CO2 in the gas phase was
determined using its volume, pressure, and temperature [47]. The principle of this
method is to accomplish a known volume of gas with the known-volume polyamine
solutions. The amount of CO2 gas introduced into the reaction tank was determined
by the change in the buffer tank, before and after injection. After equilibrium was
achieved at a constant temperature, the amount of CO2 gas absorbed in the solutions
was equal to total amount of CO2 subtracted by the amount of CO2 in the vapor
phase. The CO2 solubility, described as CO2 loading, in the liquid phase was
defined as the mole amount of CO2 in the liquid phase divided by the mole number
or mass of amine. The experimental error of CO2 loading is estimated to be ±8 %.
At low partial pressure (<10 kPa), the pressure transducer error may have a
greater impact on results. Therefore, the CO2 partial pressure was obtained using gas
chromatography (Agilent 7890). The gas phase is primarily composed of CO2, with
minor components of N2, H2O, and amine. The partial pressure of amine is small
enough to be considered negligible. The partial pressure of water was calculated
according to Raoult’s law. Prior to the introduction of CO2 into the reaction tank, the
partial pressure of N2 was calculated by subtracting H2O partial pressure from total
pressure. Correspondent to the increase in CO2 amount, the solution’s volume
change remained negligible and the N2 partial pressure was assumed constant.
The apparatus was verified by determining the solution of CO2 in 30 mass%
MEA solution at T = 313.15 K and 393.15 K. The results, listed in Fig. 1, are
compared with reported data [48–50]. Our results are agree with reported data.
Table 2 Molecular structure of chemicals
Abbreviation DETA DPTA TETA TEPA TMDA
Molecular structure
Solubility of Carbon Dioxide in Aqueous Solutions …
267
268 J. Chen et al.

(a) (b)
10000
1000

100 1000
PCO2/kPa

PCO2/kPa
10 100

1 10

0.1 1

0.01 0.1
0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8
, mole CO2/mole amine , mole CO2/mole amine

Fig. 1 Solubility of CO2 in a 30 mass % MEA solution. a T = 313.15 K. Filled circle this work;
open square Lee [5]; open triangle Shen [26]. b T = 393.15 K. Filled circle this work; open
square Lee [5]; open triangle Ma’mun [7]; open circle Jou [6]

3 Results and Discussions

CO2 solubility levels in ten polyamine aqueous solutions at 313.15 and 393.15 K
were measured using the constant-volume method combined with gas chromatog-
raphy analysis, and the results are listed in Tables 3 and 4. The corresponding
solubility levels are expressed in two forms, both involving the mole number of
CO2; in one form, it is dissolved in per mole of amine, and in the second, per
kilogram of amine. The relationship between solubility and molecular structure can
be seen in Figs. 2, 3 and 4.

3.1 Relationship Between the Solubility and Molecular


Structure

1. Effect of carbon numbers between the amino groups


The structure of DPTA differs from DETA by the addition of a CH2 group in each
of the alkyl chain, which increases the carbon number between amino groups.
Compared with DETA, DPTA has higher solubility at 313.15 K, as shown in
Fig. 2. The superior performance of DPTA is caused by the high basicity of
nitrogen owing to the reduction of the interaction between the amino groups. The
solubility in DETA at 393.15 K is higher than that in DPTA below 100 kPa, but it
is lower above 100 kPa. Because DETA and DPTA are long-chain amines, the
intermolecular interactions can affect the reaction significantly. At high tempera-
tures, the effect can be reduced, and DETA is more vulnerable by the temperature
than DPTA. Once the pressure reaches above 100 kPa, amines can be fully utilized
owing to the increased CO2 partial pressure; thus the regular pattern is consistent
with that at 313.15 K. The solubility data for DETA in the literature [38] are also
given in Fig. 2. These data are consistent with our measurement.
Table 3 CO2 solubility aCO2 in 2.5 mol/L amine aqueous solution under different pressures, at 313.15 K. aCO2 (mol CO2/mol amine), pCO2 (kPa)a
aCO2 pCO2 aCO2 pCO2 aCO2 pCO2
DETA 1.98 ± 0.16 292. ± 6. 1.08 ± 0.09 1.35 ± 0.11
0.352 ± 0.028 0.0139 ± 0.0003 2.02 ± 0.16 363. ± 7. 1.32 ± 0.11 1.40 ± 0.11
0.477 ± 0.038 0.0142 ± 0.0003 TEPA 1.48 ± 0.12 1.45 ± 0.12
0.863 ± 0.069 0.0628 ± 0.0013 0.601 ± 0.048 0.0230 ± 0.0005 1.63 ± 0.13 18.0 ± 0.4
0.985 ± 0.079 0.131 ± 0.003 1.079 ± 0.086 0.0681 ± 0.0014 1.71 ± 0.14 35.2 ± 0.7
1.11 ± 0.09 0.352 ± 0.007 1.34 ± 0.11 0.263 ± 0.005 1.78 ± 0.14 74.9 ± 1.5
1.24 ± 0.10 1.299 ± 0.026 1.57 ± 0.13 0.667 ± 0.013 1.81 ± 0.15 123. ± 2.
1.36 ± 0.11 6.09 ± 0.12 1.81 ± 0.15 3.99 ± 0.080 1.84 ± 0.15 173. ± 3.
1.49 ± 0.12 29.3 ± 0.6 1.97 ± 0.16 16.4 ± 0.3 1.89 ± 0.15 264. ± 5.
1.59 ± 0.13 73.7 ± 1.5 2.03 ± 0.16 39.8 ± 0.8 1.94 ± 0.15 354. ± 7.
1.64 ± 0.13 384. ± 8. 2.08 ± 0.17 77.8 ± 1.6 1.97 ± 0.16 451. ± 9.
DPTA 2.14 ± 0.17 127. ± 3. 1.99 ± 0.16 518. ± 10.
1.04 ± 0.08 0.0735 ± 0.0015 2.20 ± 0.18 175. ± 4. 2.01 ± 0.16 564. ± 11.
Solubility of Carbon Dioxide in Aqueous Solutions …

1.19 ± 0.10 0.210 ± 0.004 2.25 ± 0.18 243. ± 5. TMDA


1.33 ± 0.11 0.355 ± 0.007 2.30 ± 0.18 342. ± 7. 0.726 ± 0.060 0.0519 ± 0.0011
1.45 ± 0.12 1.75 ± 0.04 2.33 ± 0.19 428. ± 9. 1.01 ± 0.08 10.60 ± 0.21
1.58 ± 0.13 12.3 ± 0.2 2.36 ± 0.19 483. ± 10. 1.16 ± 0.09 57.8 ± 1.2
1.70 ± 0.14 41.0 ± 0.8 2.38 ± 0.19 526. ± 11. 1.25 ± 0.10 110. ± 2.
1.78 ± 0.14 76.2 ± 1.5 TETA 1.40 ± 0.11 1.35 ± 0.11
1.86 ± 0.15 147. ± 3. 0.618 ± 0.050 1.16 ± 0.09 1.40 ± 0.11 360. ± 7.
1.91 ± 0.15 207. ± 4. 0.826 ± 0.066 1.25 ± 0.10 1.45 ± 0.12 477. ± 10.
a
Standard uncertainties u(c) = 0.01 mol/L, u(T) = 0.2 K
269
Table 4 CO2 solubility aCO2 in 2.5 mol/L amine aqueous solution under different pressures, at 393.15 K. aCO2 (mol CO2/mol amine), pCO2 (kPa)a
270

aCO2 pCO2 aCO2 pCO2 aCO2 pCO2


DETA 1.1387 ± 0.0911 145.20 ± 2.90 0.4297 ± 0.0344 12.00 ± 0.24
0.1074 ± 0.0086 1.70 ± 0.03 1.2449 ± 0.0996 207.20 ± 4.14 0.5396 ± 0.0432 18.40 ± 0.37
0.1527 ± 0.0122 2.90 ± 0.06 1.3207 ± 0.1057 287.60 ± 5.75 0.6584 ± 0.0527 27.80 ± 0.56
0.2091 ± 0.0167 3.70 ± 0.07 1.3764 ± 0.1101 402.40 ± 8.05 0.7744 ± 0.0620 35.20 ± 0.70
0.2732 ± 0.0219 6.90 ± 0.14 1.4115 ± 0.1129 466.20 ± 9.32 0.8943 ± 0.0715 56.00 ± 1.12
0.3391 ± 0.0271 7.30 ± 0.15 1.4391 ± 0.1151 571.40 ± 11.43 1.0135 ± 0.0811 87.00 ± 1.74
0.4350 ± 0.0348 9.50 ± 0.19 TEPA 1.0970 ± 0.0878 123.00 ± 2.46
0.4914 ± 0.0393 10.20 ± 0.20 0.2137 ± 0.0171 2.50 ± 0.05 1.1773 ± 0.0942 171.80 ± 3.44
0.5864 ± 0.0469 15.80 ± 0.32 0.3435 ± 0.0275 5.60 ± 0.11 1.2614 ± 0.1009 243.20 ± 4.86
0.6672 ± 0.0534 21.30 ± 0.43 0.4779 ± 0.0382 11.30 ± 0.23 1.3266 ± 0.1061 328.50 ± 6.57
0.7443 ± 0.0595 31.70 ± 0.63 0.6132 ± 0.0491 17.20 ± 0.34 1.3749 ± 0.1100 410.00 ± 8.20
0.8245 ± 0.0660 43.90 ± 0.88 0.7289 ± 0.0583 23.20 ± 0.46 1.4211 ± 0.1137 495.20 ± 9.90
0.9160 ± 0.0733 67.50 ± 1.35 0.8665 ± 0.0693 35.50 ± 0.71 1.4492 ± 0.1159 565.00 ± 11.30
1.0154 ± 0.0812 112.60 ± 2.25 0.9836 ± 0.0787 50.20 ± 1.00 TMDA
1.0857 ± 0.0869 164.30 ± 3.29 1.1624 ± 0.0930 72.60 ± 1.45 0.1177 ± 0.0094 4.60 ± 0.09
1.1492 ± 0.0919 241.50 ± 4.83 1.2802 ± 0.1024 104.00 ± 2.08 0.2257 ± 0.0181 5.90 ± 0.12
1.1938 ± 0.0955 316.50 ± 6.33 1.4071 ± 0.1126 159.00 ± 3.18 0.3312 ± 0.0265 11.10 ± 0.22
1.2293 ± 0.0983 391.30 ± 7.83 1.5217 ± 0.1217 220.60 ± 4.41 0.4311 ± 0.0345 16.70 ± 0.33
1.2607 ± 0.1009 480.90 ± 9.62 1.6214 ± 0.1297 309.90 ± 6.20 0.5307 ± 0.0425 26.00 ± 0.52
DPTA 1.6971 ± 0.1358 400.40 ± 8.01 0.6418 ± 0.0513 43.10 ± 0.86
0.1312 ± 0.0105 6.60 ± 0.1320 1.7644 ± 0.1412 463.60 ± 9.27 0.7446 ± 0.0596 71.60 ± 1.43
0.2522 ± 0.0202 11.20 ± 0.2240 1.8224 ± 0.1458 530.40 ± 10.61 0.8560 ± 0.0685 137.00 ± 2.74
0.3757 ± 0.0301 24.80 ± 0.4960 1.8727 ± 0.1498 562.3 ± 11.25 0.9314 ± 0.0745 233.30 ± 4.67
(continued)
J. Chen et al.
Table 4 (continued)
aCO2 pCO2 aCO2 pCO2 aCO2 pCO2
0.5069 ± 0.0406 35.20 ± 0.7040 1.9089 ± 0.1527 620.0 ± 12.40 1.0070 ± 0.0806 398.00 ± 7.96
0.6385 ± 0.0511 51.40 ± 1.0280 TETA 1.0517 ± 0.0841 544.70 ± 10.89
0.7680 ± 0.0614 65.90 ± 1.3180 0.0868 ± 0.0069 1.60 ± 0.03 1.0719 ± 0.0858 627.10 ± 12.54
0.8987 ± 0.0719 78.40 ± 1.5680 0.1997 ± 0.0160 3.70 ± 0.07
1.0235 ± 0.0819 99.40 ± 1.99 0.3085 ± 0.0247 7.30 ± 0.15
a
CO2 solubility aCO2
Solubility of Carbon Dioxide in Aqueous Solutions …
271
272 J. Chen et al.

(a) 1000 (b) 1000


100 100

10 10

PCO2/kPa
PCO2/kPa

1 1

0.1 0.1

0.01 0.01
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0 2 4 6 8 10 12 14 16
, mole CO2/mole amine , mole CO2/kg amine

Fig. 2 Solubility of CO2 in 2.5 molL−1 DETA, and DPTA solutions. Filled square DETA
313.15 K; open square DETA 393.15 K; filled circle DPTA 313.15 K; open circle DPTA 393.15 K;
filled triangle DETA 313.15 K, Hartono [38]; open triangle DETA 393.15 K, Hartono [38]

(a) 1000 (b) 1000


100 100
PCO /kPa

10 10
PCO /kPa

2
2

1 1

0.1 0.1

0.01 0.01
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 0 2 4 6 8 10 12 14 16
, mole CO2/mole amine , mole CO2/kg amine

Fig. 3 Solubility of CO2 in 2.5 molL−1 DETA, TETA, and TEPA solutions. Filled triangle
DETA 313.15 K; open triangle DETA 393.15 K; filled circle TETA 313.15 K; open circle TETA
393.15 K; filled square TEPA 313.15 K; open square TEPA 393.15 K

(a) 1000 (b) 1000


100 100

10 10
PCO2/kPa

PCO2 /kPa

1 1

0.1 0.1

0.01 0.01
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0 2 4 6 8 10 12 14 16 18 20
, mole CO2/mole amine , mole CO2/kg amine

Fig. 4 Solubility of CO2 in 2.5 molL−1 TMDA, and DPTA solutions. Filled square TMDA
313.15 K; open square TMDA 393.15 K; filled circle DPTA 313.15 K; open circle DPTA
393.15 K
Solubility of Carbon Dioxide in Aqueous Solutions … 273

2. Effect of the number of amino groups


While DETA, TETA, and TEPA are similar in structure, they all have the enamine
unit but differ in the number of amino groups. A numerical increase of amino
groups in the structure leads to the increased availability of reaction sites. As shown
in Fig. 3a, the solubility increases with increasing amino groups. In particular, the
enhancement amplitude can be enlarged as the pressure increases. For further
comparison, the solubility curves per kilogram of amines are presented in Fig. 3b.
The solubility in DETA is the highest at the absorption and desorption process. As
the number of amino groups increases, the steric hindrance increases, leading to
reduced utilization of nitrogen. In addition, due to the high reactivity associated
with high temperatures, the tendencies are more obvious than at low temperatures.
In conclusion, DETA has the highest cyclic absorption capacity in the unit of mole
CO2/kg amine.
Again, the results for TMDA and DPTA are compared in the Fig. 4. They both
have allylamines, while TMDA has two amino groups and DPTA has three amino
groups. As shown in Fig. 4a, solubility of CO2 is clearly increased from TMDA to
DPTA at the temperature 313.15 K. While at the temperature 393.15 K and low
pressure around 10 kPa, CO2 solubility is similar for two amines. Also for further
comparison, the solubility curves per kilogram of amines are presented in Fig. 4b.
The solubility in TMDA is higher than DPTA at both the absorption and desorption
temperatures. But considering the cyclic absorption ability, both amines are similar
with each other.

3.2 CO2 Absorption Heat

As one of the important properties, the absorption heat of CO2 in the aqueous solution
of amine can be estimated using the following Gibbs-Helmholtz equation [51]:
 
DHabs dlnfCO2
¼ ð1Þ
R d1=T xCO
2

where xCO2 is the mole fraction or equilibrium loading of CO2, fCO2 is the fugacity
of CO2.
In order to facilitate the calculation, the fugacity of CO2 is approximately equal
to the partial pressure of CO2. According to the equation, the absorption heat of
CO2 in different amine aqueous solutions can be calculated from the corresponding
solubility data. In this article, the solubility data were measured at temperatures as
313.15 and 393.1 K. This means that the absorption heat of CO2 in these amines are
the mean value between these two temperatures. Results are presented in Fig. 5. It
is shown that DETA, DPTA, TEPA, and TETA have approximately the same
absorption heat of CO2. The value is approximately −85 kJ/mol. By comparing the
structurally similar molecules DETA, TEPA, and TETA, one can find that adding
274 J. Chen et al.

Fig. 5 Heats of absorption of 100


CO2 in aqueous solutions of
amines: filled square DETA; 90
filled circle DPTA; open

-Heat of absorbtion(KJ/mol)
circle TMDA; white 80
down-pointing triangle
TEPA; star TETA 70

60

50

40

30
0.4 0.6 0.8 1.0 1.2 1.4
, mole CO 2 /mole amine

the number of amino groups has little influence on the absorption heat of CO2. The
absorption heat of CO2 in DETA is smaller than it is in DPTA, evincing a trend: an
increase in alkyl groups between the two amino groups induces the increase of the
absorption heat of CO2. In particular, TMDA have lowest absorption heat in the
measurement range, but the highest point is also around 90 kJ/mol, which means
for the loading less than 0.8 the absorption heat will be also larger than or around
90 kJ/mol. Because TMDA has only two amino group, the maximum solubility is
around 0.8. For the solubility range from 0.8 to 1.0, more and more CO2 will be
absorbed as physical solvation, so the absorption heat become small.

3.3 CO2 Cyclic Capacity and Energy Consumption

As suggested by Li et al. [46], CO2 cyclic capacity in a capture process can be


estimated with the solubility at 313 K and CO2 partial pressure of 15 kPa as the rich
loading and the solubility at 393 K and CO2 partial pressure of 15 kPa as the lean
loading. With the increased cyclic capacity, energy consumption for CO2 capture
can be decreased with less heat needed for heating the liquid solution. Table 5 lists
the cyclic capacity of CO2 in different amine solutions. The results show that CO2
cyclic capacity in aqueous DETA, DPTA, TETA and TEPA solutions are larger
than that in the aqueous 4.91 M MEA solution by about 50–100 %, among them
DPTA has the highest cyclic capacity. In a usual capture process with the aqueous
30 % MEA solution, about 15 % of energy is used for liquid heating. Considering
the absorption reaction heats are similar for these amine solutions, 50–100 %
increase in the CO2 cyclic capacity could lead to save about one-third to half in the
energy for heating the solvent, i.e., about 5.0–7.5 % saving in the energy con-
sumption in a CO2 capture process.
Solubility of Carbon Dioxide in Aqueous Solutions … 275

Table 5 Estimated cyclic capacity of CO2 in aqueous solutions


Amine type MEA DETA DPTA TETA TEPA TMDA
Amine molecular weight 61 103 131 146 189 74
Amine concentration, M 4.91 2.5 2.5 2.5 2.5 2.5
Amine concentration, wt% 30 25.8 32.8 36.5 47.3 18.5
Cyclic capacity of CO2, g/kg 65 99 140 132 154 71.5
solution
Cyclic capacity estimated in 65 115 128 108 97 116
30 %, g/kg solution

4 Conclusions

The solubility of CO2 in aqueous linear polyamines have been measured. The
amines used are diethylenetriamine (DETA), dipropylenetriamine (DPTA),
trimethylenediamine (TMDA), tetraethylenepentamine (TEPA), triethylenete-
tramine (TETA). For absorption of CO2 at the low partial pressure in flue gas, the
cyclic capacity of CO2 in these solvents are about 50–100 % more than that in the
aqueous MEA solution, and DPTA is the best of five solvents. The results show that
these polyamines are new energy efficient solvents for CO2 capture.

Acknowledgments This work was supported by the National Natural Science Foundation of
China (Nos. 51134017, 21276010 and 51373019), National Science and Technology Support
Program of China (No. 2015BAC04B01 and 2015BAC04B02) and State Key Laboratory of
Chemical Engineering of China (SKL-ChE-12Z01).

References

1. Budzianowski WM (2015) Single solvents, solvent blends, and advanced solvent systems in
CO2 capture by absorption: a review. Int J Global Warming 7(2):184–225
2. Abu-Zahra MRM, Abbas Z, Singh P, Feron P. (2013) Carbon dioxide post-combustion
capture: solvent technologies overview, status and future directions. In: Méndez-Vilas E
(ed) Materials and processes for energy: communicating current research and technological
developments, vol 1. Formatex Research Center, pp 923–934
3. Rochelle GT (2009) Amine Scrubbing for CO2 capture. Science 325:1652–1654
4. Rao AB, Rubin ES (2002) A technical, economic, and environmental assessment of
amine-based CO2 capture technology for power plant greenhouse gas control. Environ Sci
Technol 36:4467–4475
5. Lee JI, Frederick DO, Mather AE (1976) Equilibrium between carbon dioxide and aqueous
monoethanolamine solutions. J Appl Chem Biotech 26(10):541–546
6. Jou FY, Mather AE, Otto FD (1995) The solubility of CO2 in a 30 mass percent
monoethanolamine solution. Can J Chem Eng 73(1):140–147
7. Ma’mum S, Nilsen R, Svendsen HF (2005) Solubility of carbon dioxide in 30 mass %
monoethanolamine and 50 mass % methyldiethanolamine solutions. J Chem Eng Data 50
(2):630–634
276 J. Chen et al.

8. Lawson JD, Garst AW (1976) Gas sweetening data: equilibrium solubility of hydrogen sulfide
and carbon dioxide in aqueous monoethanolamine and aqueous diethanolamine solutions.
J Chem Eng Data 21(1):20–30
9. Martin JL, Otto FD, Mather AE (1978) Solubility of hydrogen sulfide and carbon dioxide in a
diglycolamine solution. J Chem Eng Data 23(2):163–164
10. Isaacs EE, Otto FD, Mather AE (1977) Solubility of mixtures of carbon dioxide and hydrogen
sulphide in an aqueous DIPA solution. Can J Chem Eng 55(2):210–212
11. Jou FY, Carroll JJ, Mather AE, Otto FD (1993) Solubility of carbon dioxide and hydrogen
sulfide in a 35 wt% aqueous solution of methyl- diethanolamine. Can J Chem Eng 71(2):264–
268
12. Kuranov G, Rumpf B, Smirnova NA, Maurer G (1996) Solubility of single gas carbon
dioxide and hydrogen sulfide in aqueous solutions of N-methyldiethanolamine in the
temperature range 313–413 K at pressures up to 5 MPa. Ind Eng Chem Res 35(6):1959–1966
13. Rochelle GT, Goff GS, Cullinane JT. (2002) Research needs for CO2 capture from flue gas by
aqueous absorption/stripping[C]. Laurance reid gas conditioning conference. Oklahoma City,
OK, USA
14. Chakraborty AK, Astarita G, Bischoff KB (1986) CO2 absorption in aqueous solutions of
hindered amines. Chem Eng Sci 41(4):997–1003
15. Sartori G, Savage DW (1983) Sterically hindered amines for CO2 removal from gases. Ind
Eng Chem Fundam 22(2):239–249
16. Singh P, Niederer JPM, Versteeg GF (2007) Structure and activity relationships for amine
based CO2 absorbents. Int J Greenhouse Gas Control 1(1):5–10
17. Tontlwachwuthlkul P, Melsen A, Llm CJ (1991) Solubility of CO2 in
2-Amino-2-methyl-1-propanol Solutions. J Chem Eng Data 36(1):130–133
18. Baek JI, Yoon JH (1998) Solubility of carbon dioxide in aqueous solutions of
2-amino-2-methyl-1,3-propanediol. J Chem Eng Data 43(4):635–637
19. Park JY, Yoon SJ, Lee H, Yoon JH, Shim JG, Lee JK, Min BY, Eum HM, Kang MC (2002)
Solubility of carbon dioxide in aqueous solutions of 2-amino-2-ethyl-1,3-propanediol. Fluid
Phase Equilib 202(2):359–366
20. Bishnoi S, Rochelle GT (2000) Absorption of carbon dioxide into aqueous piperazine:
reaction kinetics, mass transfer and solubility. Chem Eng Sci 55(22):5531–5543
21. Kadiwala S, Rayer AV, Henni A (2010) High pressure solubility of carbon dioxide (CO2) in
aqueous piperazine solutions. Fluid Phase Equilib 292(1–2):20–28
22. Ma’mun S, Jakobsen JP, Svendsen HF, Juliussen O (2006) Experimental and modeling study
of the solubility of carbon dioxide in aqueous 30 mass% 2-((2-aminoethyl)amino)ethanol
solution. Ind Eng Chem Res 45(8):2505–2512
23. Singh P, Versteeg GF (2008) Structure and activity relationships for CO2 regeneration from
aqueous amine-based absorbents. Process Saf Environ Prot 86(5):347–359
24. Liu YX, Dong LH, MI JG, Chen J (2012) Study on molecular structure of alkanolamines and
their CO2 capture ability. Sci Sinica Chim 42(3):291–296
25. Li MH, Shen KP (1992) Densities and solubilities of solutions of carbon dioxide in water +
monoethanolamine + N-methyldiethanolamine. J Chem Eng Data 37(3):288–290
26. Shen KP, Li MH (1992) Solubility of carbon dioxide in aqueous mixtures of
monoethanolamine with methyldiethanolamine. J Chem Eng Data 37(1):96–100
27. Li MH, Chang BC (1995) Solubility of mixtures of carbon dioxide and hydrogen sulfide in
water + monoethanolamine + 2-amino-2-methyl-1-propanol. J Chem Eng Data 40(1):328–
331
28. Seo DJ, Hong WH (1996) Solubilities of carbon dioxide in aqueous mixtures of
diethanolamine and 2-amino-2-methyl-1-propanol. J Chem Eng Data 41(2):258–260
29. Jane IS, Li MH (1997) Solubilities of mixtures of carbon dioxide and hydrogen sulfide in
water + diethanolamine + 2-amino-2-methyl-1-propanol. J Chem Eng Data 42(1):98–105
30. Dash SK, Samanta AN, Bandyopadhyay SS (2012) Experimental and theoretical investigation
of solubility of carbon dioxide in concentrated aqueous solution of
2-amino-2-methyl-1-propanol and piperazine. J Chem Thermodyn 51(1):120–125
Solubility of Carbon Dioxide in Aqueous Solutions … 277

31. Derks PWJ, Hogendoorn JA, Versteeg GF (2010) Experimental and theoretical study of the
solubility of carbon dioxide in aqueous blends of piperazine and N-methyldiethanolamine.
J Chem Thermodyn 42(1):151–163
32. Speyer D, Ermatchkov V, Maurer G (2010) Solubility of carbon dioxide in aqueous solutions
of N-methyldiethanolamine and piperazine in the low gas loading region. J Chem Eng Data
55(1):283–290
33. Ermatchkov V, Maurer G (2011) Solubility of carbon dioxide in aqueous solutions of
N-methyl-diethanol-amine and piperazine: prediction and correlation. Fluid Phase Equilib 302
(1–2):338–346
34. Silkenbäumer D, Rumpf B, Lichtenthaler RN (1998) Solubility of carbon dioxide in aqueous
solutions of 2-amino-2-methyl-1-propanol and N-methyldiethanolamine and their mixtures in
the temperature range from 313 to 353 K and pressures up to 2.7 MPa. Ind Eng Chem Res 37
(8):3133–3141
35. Ma’mun S, Svendsen HF, Hoff KA, Juliussen O (2007) Selection of new absorbents for
carbon dioxide capture. Energy Convers Manag 48:251–258
36. Hartono A, da Silva EF, Svendsen HF (2009) Kinetics of carbon dioxide absorption in
aqueous solution of diethylenetriamine (DETA). Chem Eng Sci 64:3205–3213
37. Singh P, Niederer JPM, Versteeg GF (2009) Structure and activity relationships for
amine-based CO2 absorbents-II. Chem Eng Res Des 87:135–144
38. Hartono A, Hoff KA, Mejdell T, Svendsen HF (2011) Solubility of carbon dioxide in aqueous
2.5 M of diethylenetriamine (DETA) solution. Energy Procedia 4:179–186
39. Chang YC, Leron RB, Li MH (2013) Equilibrium solubility of carbon dioxide in aqueous
solutions of (diethylenetriamine + piperazine). J Chem Thermodyn 64:106–113
40. Kim I (2009) Heat of reaction and VLE of post combustion CO2 absorbent. Norwegian
University Trondheim, Norway
41. Aronu UE, Svendsen HF, Hoff KA, Juliussen O (2009) Solvent selection for carbon dioxide
absorption. Energy Procedia 1:1051–1057
42. Singh P, Niederer JPM, Versteeg GF (2007) Structure and activity relationships for
amine-based CO2 absorbents-I. Int J Greenhouse Gas Control 1:5–10
43. Singh P, Versteeg GF (2008) Structure and activity relationships for CO2 regeneration from
aqueous amine-based absorbents. Process Saf Environ Prot 86:347–359
44. Machida H, Yamada H, Fujioka Y, Yamamoto S (2015) CO2 Solubility measurements and
modeling for tertiary diamines. J Chem Eng Data 60:814–820
45. Dong L, Chen J, Gao G (2010) Solubility of carbon dioxide in aqueous solutions of
3-amino-1-propanol. J Chem Eng Data 55:1030–1034
46. Li H, Moullec YL, Lu JH, Chen J, Marcos JCV, Chen GF (2014) Solubility and energy
analysis for CO2 absorption in piperazine derivatives and their mixtures. Int J Greenhouse Gas
Control 31:25–32
47. Peng DY, Robinson DB (1976) 0 A new two-constant equation of state. Ind Eng Chem
Fundam 15:59–64
48. Lee JI, Otto FD, Mather AE (1976) The measurement and prediction of the solubility of
mixtures of carbon dioxide and hydrogen sulphide in a 2.5 N monoethanolamine solution.
Can J Chem Eng 54:214–219
49. Ma’mun S, Nilsen R, Svendsen HF (2005) Solubility of carbon dioxide in 30 mass %
monoethanolamine and 50 mass % methyldiethanolamine solutions. J Chem Eng Data
50:630–634
50. Jou FY, Mather AE, Otto FD (1995) The solubility of CO2 in a 30 mass percent
monoethanolamine solution. Can J Chem Eng 73:140–147
51. Sherwood AE, Prausnitz JM (1962) The heat of solution of gases at high pressure. AIChE J
8:519–521

Potrebbero piacerti anche