Sunteți pe pagina 1din 62

Literature Review

Development and integration of novel


adsorbents for post combustion carbon capture
Christopher Starkie
August 2012
Literature review August 2012

Executive summary

The purpose of this literature review is to examine the current developments in


solid adsorbents for the application of post combustion carbon dioxide (CO 2) separation.
Common to much of the existing literature, this review will analyse the CO 2 adsorption
capacity of the materials but further to this we will assess their potential for use at scale
in real world applications. Presently much of the research focuses on ideal conditions
with token efforts to consider feasibility, regeneration and flue gas composition. The aim
is to survey the literature for materials with high potential for commercial full scale
applications but with current surmountable barriers. This incorporates the objectives of
the project to develop functioning adsorbents for the separation of CO 2. After reviewing
all prevalent classes of materials metal organic frameworks and solid supported amines
were reviewed in greater detail and future directions explored. From this several
potential research directions were highlighted. Coordinatively unsaturated metal organic
frameworks exhibit high heats of adsorption for CO2 allowing them to be used as
potential chemisorbents. Furthermore the surface functionalisation of metal organic
frameworks is an emerging area and little work in this area has focused on the
developments of chemisorbents materials. Novel supports for immobilised amines such
as those that encapsulate the impregnated amine were found to improve the stability of
class one amine adsorbents. The development of class three adsorbents combining the
benefits of high capacities of class one materials with the stability of class two materials
show excellent potential for CO2 selective materials. With further development of metal
organic frameworks and solid amine adsorbents and their subsequent process integration
will solid adsorption systems could lead to greatly improved post combustion carbon
capture processes.

2
Literature review August 2012

Glossary

AP 3-aminopropyltrimethoxysilane
APTMS 3-Aminopropyltrimethoxysilane
BDC-NH2 2-amino-1,4-benzenedicarboxylate
btc 1,3,5-benzenetricarboxylate
bttb 4,4',4'',4'''-benzene-1,2,4,5-tetraaryltetrabenzoate
BTTri 1,3,5-tri-(1-H-1,2,3-triazol-4-yl)benzene
cnc 4-Carboxycinnamic
CNT Carbon nanotube
CO2 Carbon dioxide
CTAB hexadecyltrimethylammonium bromide
cyclam 1,4,8,11- tetraazacyclotetradecane
DEA diethanol amine
DMAP 3-dimethylaminopropyltrimethoxysilane
DABCO 1,4-diazabicyclo(2,2,2)octane
dobdc 2,5-dihydroxyterephthalic acid
dpt 3,6-Di-4-pyridyl-1,2,4,5-tetrazine
en enamine
HAS Hyperbranched amino silica
ICA Imidazole carbonyl aldehyde
IR Infrared spectroscopy
KIT-x Korean Institute of Science and Technology material
MAP 3-(methylamino)propyltrimethoxysilane
MC400 Mesoporous capsule material
MCM-x Mobil composition of matter -x
MEA Monoethnanol amine
MIL-X Material of Institute Lavier-x
mtb methanetetrabenzoate
NMR Nuclear magnetic resonance spectroscopy
PAP 3-(pheylamino)propyltrimethoxysilane
pbmp N,N'-piperazinebismethylenephosphonate
p-cdc deprotonated form of 1,12-dihydroxycarbonyl-1,12 dicarba-closo-dodecaborane
PEHAS pore-expanded hyperbranched aminosilica
PEI Polyethyleneimine
PE-MCM pore-expanded Mobil composition of matter
PMMA Poly(methyl methacrylate)
PSA Pressure swing adsorption
py-CF3 4-trifluoromethylpyridine
SBA-x Santa Barbara Amorphous type material
TEOS Tetraethyl orthosilicate
TEPA Tetraethylenepentamine
TSA Temperature swing adsorption
VSA Vacuum swing adsorption
ZIF Zeolitic imidazolate framework

3
Literature review August 2012

Contents

Executive summary 2

Glossary 3

1.0 Introduction 5

1.1 Current amine capture systems 6

1.2 Challenges for new materials 8

1.3 Solid adsorbents for carbon capture 11

1.4 Down selection of materials 15

1.5 References 16

2.0 Metal organic frameworks 18

2.1 Relevance to carbon capture and storage 19

2.2 Metal organic frameworks with vacant coordination sites for selective 20
CO2 capture

2.3 Interpenetrated metal organic frameworks for CO2 separation 23

2.4 Flexible frameworks for selective CO2 separation 24

2.5 Functionalised metal organic frameworks for selective CO2 capture 25

2.6 Potential of metal organic frameworks for use in post combustion 30


capture

2.7 References 31

3.0 Solid supported amines 34

3.1 Class 1: Physically adsorbed amines 34

3.2 Class 2: Covalently bound amines 45

3.3 Class 3: Covalently bound amine polymers 53

3.4 Potential for solid supported amine adsorbents for use in carbon 55
capture systems

3.5 References 56

4.0 Conclusion 60

5.0 Outlook 62

4
Literature review August 2012

1.0 Introduction

Increasing levels of anthropogenic CO2 emissions brought about through the


combustion of fossil fuels is believed to contribute towards global warming and
subsequent climate change.[1] A potential strategy for the mitigation of further CO 2
emissions is to use carbon capture and storage technology on large point sources of CO2
such as thermal power stations. This will allow the continued use of fossil fuels with
further release of CO2 emissions. A range of approaches have been suggested to capture
CO2 as  the  current  approach  dating  from  the  1930’s  is  a sub-optimal solution leading to
unacceptable increases in electricity prices. Solid CO2 selective adsorbents are being
investigated as a replacement technology and have the potential to vastly decrease the
costs associated with CO2 capture. This project aims to develop novel solid adsorbents
that function under realistic operating conditions. These materials have the potential to
reduce the costs associated with carbon capture and storage and lead to the formation of
both practical and economic systems.

5
Literature review August 2012

1.1 Current carbon capture systems

Current carbon capture and storage (CCS) systems use liquid amine solvents to
reversibility capture CO2 by formation of a carbamate species. Briefly a 15-30% amine
solution in sprayed counter to the flue gases motion in the adsorber and the CO2 reacts
in the amine solution. The amine solution is then heated releasing the CO 2 and
regenerating the amine solvent. The details of this process are outlined in box 1.1.

Primary, secondary and tertiary amines can abdsorb CO2 via a small range of
mechanisms leading to overall reversible chemisoprtion. Primary and secondary amines
react via the nitrogen based lone pair with CO2 to form zwitterionic intermediates. An
arbitrary base with can either be a secondary amine, hydroxyl species or even as water
molecule proceeds to deprotonate the zwitterionic species to form a carbamate species.
Under dry conditions those without the presence of water and hydroxyl anions the molar
ratio between CO2 and amines is limited to 0.5. This is termed the amine efficiency and
can be increase towards 1 under humid conditions where water can act as the base.
Tertiary amines react with CO2 via a different mechanism; being more basic they readily
deprotonate water to form quaternary amine and an associated hydroxyl ion. This
hydroxyl anion then reacts with CO2 to form a bicarbonate which is strongly associated
with the quaternary amines. It is possible for primary and secondary amines to undergo
a similar reaction pathway, however their lower pKa leads to the carbamate pathway to
be more competitive.

Box 1.1: An overview of current amine carbon capture systems

6
Literature review August 2012

This technique has several limitations which add to the overall cost of the system
leading to an acceptable increase is cost of electricity. The principal disadvantage is the
large energy requirement of regenerating the amine solvent. This is due to the large
amounts of water present which has a significant heat capacity as a result of the
hydrogen bonding network. The high pH of the amine used and the carbonic acid formed
on dissolution can lead to high levels of corrosion. To limit said corrosion amine levels
are kept below 30% which greatly reduces the capture efficiency of the system. As the
amine is sprayed down a contact column of packed trays the contact efficiency of the
system could also be improved. Additives are also required to prevent foaming to the
amine solvent systems and to promote the reaction. The stability of amine solvent
systems is also questionable as the amines used react irreversibility with SO2 and are
degraded by NOX and oxygen. Amine slip, a by-product of volatile amine use could
become an environmental problem if used on large scale such as a power station. It was
noted that for each tonne of CO2 captured 0.0032 tonnes of MEA were released to the
atmosphere.[2] Taking into account a typical 2.4 GW pulverised coal plant produces 30-
50 tonnes CO2/min the extent of amine slip becomes apparent.
The pitfalls of this technology currently make it an unattractive carbon abatement
technology when compared on a cost basis to other technologies (Figure 1.1). Carbon
capture technology is regarded as being in its infancy with large potential gains in
efficiency possible. Research suggests that current amine systems consume ~24% of a
power stations output, 16% separation and 8% compression. The theoretical minimum is
8% of a power stations output distributed as 3.2% separation and 4.6% compression
leading increasing cost competitiveness with other low carbon technologies.[3] Fossil
fuels power plants that use CCS technology also have benefits over other technologies
shown is Figure 1.1. Thermal powers stations provide power on demand and are not
influenced by environmental factors and are able to meet fluctuating demand.
Furthermore they do not produce persistent harmful waste and well understood pollution
control technologies are available.

Figure 1.1: The levelised cost of electricity for differing methods of power generation.[4]

7
Literature review August 2012

1.2 Challenges for new materials

A pertinent point that should not be overlooked is the magnitude of


anthropogenic CO2 emissions produced, estimated to be 31.6 billion metric tonnes per
annum.[5] Thus any material used to capture CO2 irreversibly will quickly exhaust global
supplies. Further to this any material produced from CO2 would rapidly saturate the
market need for that product making it less competitive. Hence it is of prime importance
that the materials used to sequester CO2 be completely regenerable. The sorbent
materials must satisfy strict criteria in order to be successful operationally as well as
economically. The criterion were selected to ensure that solid adsorbents are vastly
economically superior to that of liquid amine systems. The factors given below drawn
from the current literature and reflect the properties and functionality required by solid
adsorbent systems [6-9]

1.2.1 Process integration

The process as a whole should have an energy reduction of no less than 30-50 %
of the energy requirements for the wet MEA process. This is partially down to the
accepted fact that solid adsorbents with have increased capital costs which need to be
justified by greater operational savings. The adsorbents must be able to function within
solid particle adsorption beds with a minimum pressure drop of the flue gas.

1.2.2 Adsorption capacity

The equilibrium adsorption capacity is often quoted in the literature as a measure


of a materials performance. However, the adsorption capacity after a short time or
working capacity is of more use in practical situations as equilibrium capacity is unlikely
to be reached. The working capacity is taken as the difference in uptake of adsorption
isotherms between the adsorption and regeneration temperature. It has been estimated
that for solid adsorbents to be cost competitive with liquid amine systems a capture
capacity of at least 3.0 mmolCO2 g-1 is required.[10] The adsorption capacity is a crucial
variable as it determines the amount of adsorbent required and the size of the adsorber
vessel.

1.2.3 Selectivity for CO2

To make the processes of compression, storage and transport economical a high


selectivity for CO2 over other molecules is required in face of the similar properties of the
gaseous mixture. Discrimination by size or kinetic diameter (which takes into account
molecular geometry) cannot be applied to CO2 separation as CO2 is in the mid range of
molecular sizes meaning that all size separation is complicated as oppose to merely
excluding large molecules against a background of smaller molecules. Cryogenic
distillation has also been suggested as an alternative but requires the cooling vast
amounts of flue gas requiring similarly vast amounts of energy, making this method
unfeasible. Differences do exist in the electronic properties of the flue gas molecules,
namely nitrogen, oxygen, water, carbon monoxide and carbon dioxide. These differences
include polarisability, molecular orbitals and quadrupole moment, which underpin their
molecular interactions between difference species. As such, this allows us a level of
molecular control via exploiting differences between chemical reactivity of the flue gas
species.

8
Literature review August 2012

1.2.4 Kinetics of both adsorption and desorption

To ensure economic feasibility the kinetics of CO2 capture must be very rapid
under the typical operating conditions of the unit. Rapid kinetics allow for shorter
adsorption columns as less adsorbent is required. Ideal kinetics yield a sharp sigmodial
breakthrough curve allowing for immediate uptake and rapid arrival at a significant
capacity. The observed kinetics of the adsorption process is a combination of mass
transfer to the active site and the rate of reaction between CO 2 and the active species.
The porous support and the embedded functionality must strike a balance between little
diffusional resistance but sufficient activation by confined pores.

1.2.5 Thermal, Chemical and Cyclic stability

Capture of CO2 from post combustion gases involves the removal of low
concentrations of CO2 against a large background of predominantly nitrogen. Included is
these background gases are many reactive species such as water, halogens, heavy
metals, sulphur and nitrogen oxides. These species pose significant challenges as the
rate of degradation whilst a fact of life has to be within acceptable limits. Sorbent
stability is of crucial importance as the demands of practical systems disallows any
significant degradation over the sorbent lifetime. The hydrolytic stability of many second
generation materials such as MOFs and COF has been shown to be extremely poor.[11-
13] Water has also been shown to degrade the pore structure of supported amines.
Water has also been shown to competitively onto zeolites and carbon materials where it
gradually oxidises the surface leading to a reduction is capacity. [14, 15] First generation
amine solvent systems show oxidative degradation to the acetate, with pilot scale testing
showing links to the amount of chloride present in the flue gas stream.[16] Cyclic
stability is also a large area of concern as the sorbent will have to be cycled numerous
with minimal loss of capacity. The sorbent stability plays a crucial role in solid adsorbents
as the economics is effected by the adsorbent cost/lifetime ratio.

1.2.6 Mechanical strength of particles

Physical strength of the solid sorbent particles is an important but often


overlooked consideration of solid sorbent materials. On the basis of the design of the
pilot scale reactors a consistent and durable particles size is required for optimum
operation and regeneration processes. The microporous architecture and sorbent
morphology must be stable throughout multiple cycles to maintain rapid kinetics and the
required capacity for a given timescale. Conditions such as the flow rate of flue gas,
vibration, operating temperatures or mechanical movements should not have a
detrimental effect on the particles. Disintegration of the sorbent particles alters the
operating capability of the systems and increases the adsorbent make-up rate.

9
Literature review August 2012

1.2.7 Regeneration

The regeneration energy is highly important as this is directly proportional to the


energy penalty of CCS systems. The regeneration energy depends on the heat of
adsorption of the material with aqueous amine systems having very large regeneration
energies. The heat of adsorption for physisorbents is typically -25 to -50 kJmol-1 and -25
to-50 kJmol-1 for chemisorbents. A target regeneration energy deemed suitable for post
combustion capture systems is below 3600 kJ/kg (1548 BTU/lb). The regeneration
energy is compromised of the thermodynamic interaction energy of the material and its
overall heat capacity. As mentioned previously the large heat capacity of water is a
hindrance when dealing with liquid amine solvent systems. During the development of
novel adsorbents consideration should be given to the heat capacity of the support as
this will directly affect the regeneration energy. The sorbent must be regenerable under
sensible conditions whilst maintaining its adsorbent properties. The sorbent should
ideally be regenerated via a temperature swing process or via steam regeneration
without the need for an inert gas purge.

1.2.8 Availability and Cost of adsorbent

In order for solid adsorbents to be utilised in such large scale applications such as
CCS, production is required to be both economic and practical at scale. Economic
analysis by Tarka reported that the cost of a sorbent should be around $10/kg with $5
being very favourable and $15 /kg uneconomical

These targets have been identified by the scientific community and represent the
challenges facing carbon capture systems to date.[1] At present no material satisfies all f
this criterion. Numerous materials have exhibited capacities several times greater than
the target capacity although these are often under best conditions with the material
having drawbacks elsewhere. It is unlikely that a single adsorbent will satisfy all
parameters however compromises can be made in one area if the material excels at one
or several other areas.

10
Literature review August 2012

1.3 Solid adsorbents for carbon capture

Liquid amine systems are called 1st generation systems and given the large
potential for improvement a number of 2 nd and 3rd generation systems are being
developed (see figure 1.2). [17] These new materials over significant cost reduction
benefits by increasing capacity, reducing regeneration energy and offering improved
stability. Solid adsorbents are a promising alternative to liquid amine capture systems. A
principal driver in the use of solid adsorbents is the potential reduction in regeneration
energy. This is reduced by a number of factors such as the supports possessing a much
lower heat capacity than that of water. For example the heat capacity for an aqueous
solution of 30% MEA is 4.0 kJ/kg-1K-1 whilst the mesoporous silica SBA-15 has a heat
capacity of 0.8 kJ/kg-1K-1. [18] Energy is also saved via omitting the need for phase
changes and subsequent condensation. By incorporating chemical functionality onto
supports losses by evaporation are minimised. Corrosion of liquid amine solutions is
highly problematic but as the majority of the amines are bound within the pores of the
support corrosion issues are reduced. Solid adsorbents are used in fluidised bed reactors
which would be incompatible with aqueous amine solutions. Fluidised bed reactors offer
an increased contact time, better heat transfer and a reduction in plant size owing to an
additional capacity. It is estimated that solid adsorbents can reduce the energy loading
by 30-50% leading to a vast reduction in parasitic energy.

Figure 1.2: Potential materials that are being investigated for selective carbon capture.

To date a wide selection of solid adsorbents have been studied for capture of
CO2 .These all have differing properties and hence different approaches in terms of
designing carbon capture systems. Each classification of material is discussed below and
its suitability towards post combustion carbon capture systems described. Each material
will be assed against the set criteria established above.

11
Literature review August 2012

1.3.1 Physical adsorbents

Carbon dioxide may be captured using physisorbent solid materials in which the
CO2 is adsorbed through weak Van der Walls attractions between the gas species and the
surface of the solid. Owing to the weak nature of physisorption regeneration these
materials are well suited to low temperature separations.

1.3.2 Carbon based adsorbents

Activated carbons are well understood adsorbent materials used in numerous


industrial processes. They can be produced from a large range of carboneous materials
making them cheap to produce and widely available. To date porous activated carbons,
graphenes and carbon nanotubes (CNTs) have all been studied for utilisation in carbon
capture. [19] The range of materials and preparations leads to a variety of structures
and pores sizes allowing application specific carbon adsorbents to be used. Activated
carbons have fast kinetics and low regeneration requirements however they have poor
selectivity. Their capacity is reduced at flue gas temperatures due to the weak nature of
the physisorption. At low pressures activated carbons exhibit lower adsorption capacity
than zeolites owing to their unfavourable adsorption isotherms. [20] Recent
developments regarding CNTs have shown potential for use in high pressure CO2
separations, however economic production of CNT is far from a possibility.[21]

1.3.3 Zeolites

Zeolites are porous aluminosilicates formed from an extended array of MO4


tetrahedra (M= Si or Al). The inclusion of Al or other metal species imparts a negative
charge on the framework balanced by positive cations within the pore spaces. Variation
of the Si/Al ratio and ion exchange can alter the pore sizes or adsorption characteristics
of the sorbent.[22] Zeolites are widely used in the field of gas separation and purification
and as such have been studied as CO2 adsorbents. Zeolites mainly adsorb CO2 via
physisorption with a small proportion chemisorbed as either a carbonate or a carboxylate.
Zeolites have fast kinetics reaching equilibrium capacity within a matter of minutes and
work well under low temperatures and high pressures. Desorption occurs readily under
elevated temperatures and capacities are low with low gas partial pressures. Zeolites
readily adsorb water leading to high regeneration temperatures under humid conditions
as well as a reduction in capacity.[23]

12
Literature review August 2012

1.3.4 Metal organic frameworks

Metal organic frameworks (MOFs) are an emerging class of materials consisting of


single atom metal centres surrounded by a coordination sphere of multidentate linkers
which in turn bind to further metal centres.[24] Potentially there are numerous
combinations of metal centre and ligand combinations leading to tuneable material
properties. The first reported cases of porous metal organic frameworks collapsed in the
absence of a guest species and were unstable in the presence of moisture.[25] Recent
work has led to the development of a range of metal organic frameworks stable without
a host and furthermore in the presence of later amounts of water. These materials
primarily exhibit physisorption with CO2 bound by weak interactions between the ligand
and metal centres with the adsorbed gas species. Owing to this, these materials are
good low temperature high pressure adsorbents currently exhibiting the largest
adsorption capacity recorded (MOF-210 CO2 uptake capacity ~54.mmolCO2 CO2 g-
1
sorbent, 298 K, 50 atm).[26] There have been reports of CO2 being strongly bound to
vacant metal coordination sites or functionalised frameworks in a fashion similar to
chemisorptions allowing CO2 separation under lower partial pressures and at elevated
temperatures. MOFs exhibit limited stability to moisture and typically have poor
capacities under proposed operating conditions. MOFs are currently being extensively
researched for carbon capture systems with potential for the development of successful
adsorbents.

1.3.5 Chemisorbents

Most physisorbents highlighted above suffer from low CO 2 uptake capacities at


low CO2 partial pressures and often a lack of selectivity at high pressures Great
selectivity for CO2 can be achieved through the use of chemisorbent materials. These
selectively react with CO2 to via the formation of a formal bond to yield a reversible
species. Chemisorbents have a higher heat of adsorption leading to an increase in
regeneration requirements.

1.3.6 Regenerable Alkali metal carbonate based sorbents

Numerous metal oxides can adsorb acidic gases such as CO2 via interaction
between basic sites, especially those with a high charge density such as CaO, MgO and
K2O. Many metal oxides react to form the respective carbonate as highlighted in the
following exothermic reaction (where M can be any group 2 metal):

In the case of calcium oxide the 1:1 reaction stoichiometry implies that the
maximum capacity of CaO is 17.8mmolg-1 however capacities of this magnitude are not
reached due to solid defects and slow kinetics after initial adsorption of CO2.[27] The
kinetics for this process occur very slowly at room temperature and function best at high
temperatures (723-923 K) with regeneration at even higher temperatures. Such
temperatures omit this materials form post combustion carbon capture and also deform
the porous structure limiting reducing the capacity with each cycle. Magnesium oxides
have also been studied as solid adsorbents as they behave in a similar fashion to CaO
albeit lower regeneration requirements.[28] In spite of this, these materials exhibit low
capacities and sluggish kinetics at room temperature and still operate under high
pressures.

13
Literature review August 2012

Lithium zirconates (Li2ZrO3) and lithium orthosilicates (Li4SiO4) have been


explored as potential adsorbents for post combustion capture systems as they offer good
capacities with excellent thermal stability.[29] These materials function at lower
temperatures relative to other metal oxides but still too high temperatures for ideal post
combustion capture systems. The principal limitations of these materials are their
sluggish kinetics. Although exhibiting capacities of up to 4.5mmolg-1 the time taken to
heat this can be up to the order of days.[30]
Hydrotatalcites alternatively know as double layer hydroxides consist of
positively charged brucite-like [M2+1-xM3+(OH)2]x+ structures with divalent cations at the
centre of octahedral sites partially substituted for trivalent metal cations. The excess
charge is balanced by intercalating species such as carbonate anions and solvent leading
to a charge balanced framework. [31] These materials exhibit fast kinetics but have
capacities below 1mmolg-1 which is positively affected by the presence of water.[32]
These materials exhibit better stability but still function best at elevated temperatures
not suitable for post combustion capture.

1.3.7 Solid supported amines

Amines can be bound to solid surfaces to form solid analogues of liquid amine
species used in current post combustion capture systems.[33] Materials of this type have
been heavily researched for applications in post combustion capture systems. A wide
variety of amines exist and these can be attached to numerous high surface area
supports increasing the availability and capture capacity of the amines. Amines can
either be physisorbed or covalently bound to the support. These materials have been
shown to have good capture capacities under conditions similar to those proposed for
post combustion carbon capture systems. The materials have shown to be regenerable
but issues with stability in the presence of moisture and flue gas contaminants needs to
be improved.

14
Literature review August 2012

1.4 Down selection of materials

With the given constraints it is not possible to investigate all of these materials
simultaneously with a sufficient degree of rigour. Hence the materials attributes of these
materials were compared to the criteria listed previously with the comparison highlighted
below in table 1.1. After studying the potential classes, metal organic frameworks and
solid supported amines were down selected to become the focus of this study and are
discussed in great depth in sections 2 and 3 respectively. Metal organic frameworks were
nominated on the basis of recent literature highlighting successful example of
chemisorption under low partial pressures. Inclusive to this post synthetic modification
towards post combustion carbon capture is a relatively unexplored field and could yield
excellent materials. Solid supported amines have been show to be excellent materials for
post combustion capture however increases in stability and capacity under real flue gas
conditions require development. Physisorbents such as activated carbons and zeolites
are not suitable for post combustion capture applications as their capacity decreases to
impractical levels with typical post combustion temperatures and gas partial pressures.
Furthermore these materials are often adversely effected by the presence of moisture.
The process requirements for high temperature adsorbents based on metal oxides would
render the process uneconomical for post combustion capture systems.

Material CO2 Temp. Pressure Regen. Stability Kinetics Selectivity Cost


Capacity regime regime

Activated Low Low High Facile Medium Fast Poor Low


Carbon

Carbon Low Low High Facile Unknown Fast Ok High


nanotubes

Zeolites High Low Low Facile Poor Fast OK Medium

Metal High Low (some High OK Poor Fast OK-good High


organic medium) (some
frameworks Low)

Metal Low High All Difficult Poor Slow Excellent Low


oxides

Solid Good Medium Low OK Variable Fast Excellent Medium


supported
amines

Table 1.1: An overview comparing different classes of solid adsorbent towards there
suitability for post combustion capture. The adsorbents are required to selectively
capture CO2 at 40-80 °C and be regenerated between 80-120 °C.

The following sections will provide a detailed overview of the research


opportunities and challenges facing MOFs and solid support amines for use in post
combustion carbon capture. The findings will further be summarised in the conclusion
and further work highlighted.

15
Literature review August 2012

1.5 References

1. Stern, N., The Stern Review: The economics of climate change, H.M. Treasury,
2007, H M Government: London, UK.
2. Improvement in Power Generation with Post-Combustion Capture of CO2, Report
PH4/33. 2004, IEA GHG.
3. Rochelle, G.T., Amine Scrubbing for CO(2) Capture. Science, 2009. 325(5948): p.
1652-1654.
4. Institute, G.C., The costs of CCS and other low-carbon technologies. 2011.
5. Davis, S.J., G.P. Peters, and K. Caldeira, The supply chain of CO2 emissions. Proc.
Natl. Acad. Sci. U. S. A. 108(45): p. 18554-18559, S18554/1-S18554/9.
6. H. W. Pennline, J.S.H., M. L. Gray, R. V. Siriwardane, D. J. Fauth and G. A.
Richards, NETL In-house Postcombustion Sorbent-Based Carbon Dioxide Capture
Research. Annual IEP Contractors Meeting, 2009. 2009.
7. Yong, Z., V. Mata, and A.Ì.r.E. Rodrigues, Adsorption of carbon dioxide at high
temperature a review. Separation and Purification Technology, 2002. 26(3): p.
195-205.
8. Addleman, R.S., et al., Amine Functionalized Nanoporous Materials for Carbon
Dioxide Capture, in Environmental Applications Of Nanomaterials. p. 285-312.
9. Drage, T.C., et al., Materials challenges for the development of solid sorbents for
post-combustion carbon capture. Journal of Materials Chemistry. 22(7): p. 2815-
2823.
10. G.F. Schutte, E.G.L., J. B. Corss and E. Esen, in Fifth Annual Conference on
Carbon Capture and Sequestration. 2006: Alexandria, Virginia.
11. Millward, A.R. and O.M. Yaghi, Metal Organic Frameworks with Exceptionally High
Capacity for Storage of Carbon Dioxide at Room Temperature. Journal of the
American Chemical Society, 2005. 127(51): p. 17998-17999.
12. Babarao, R. and J. Jiang, Molecular Screening of Metal Organic Frameworks for
CO2 Storage. Langmuir, 2008. 24(12): p. 6270-6278.
13. Hunt, J.R., et al., Reticular Synthesis of Covalent Organic Borosilicate Frameworks.
Journal of the American Chemical Society, 2008. 130(36): p. 11872-11873.
14. Adams, L.B., et al., An examination of how exposure to humid air can result in
changes in the adsorption properties of activated carbons. Carbon, 1988. 26(4):
p. 451-459.
15. Verma, S.K. and P.L. Walker Jr, Carbon molecular sieves with stable hydrophobic
surfaces. Carbon, 1992. 30(6): p. 837-844.
16. Moser, P., et al., Performance of MEA in a long-term test at the post-combustion
capture pilot plant in Niederaussem. International Journal of Greenhouse Gas
Control. 5(4): p. 620-627.
17. Figueroa, J.D., et al., Advances in CO2 capture technology: The U.S. Department
of Energy's Carbon Sequestration Program. International Journal of Greenhouse
Gas Control, 2008. 2(1): p. 9-20.
18. Khatri, R.A., et al., Carbon dioxide capture by diamine-grafted SBA-15: A
combined Fourier transform infrared and mass spectrometry study. Industrial &
Engineering Chemistry Research, 2005. 44(10): p. 3702-3708.
19. Ruthven, D.M., Principles of Adsorption and Adsorption processes. 1984, New
York: Wiley-Interscience.
20. Chue, K.T., et al., Comparison of Activated Carbon and Zeolite 13X for CO2
Recovery from Flue Gas by Pressure Swing Adsorption. Industrial & Engineering
Chemistry Research, 1995. 34(2): p. 591-598.
21. Khoerunnisa, F., et al., Enhanced CO2 Adsorptivity of Partially Charged Single
Walled Carbon Nanotubes by Methylene Blue Encapsulation. The Journal of
Physical Chemistry C. 116(20): p. 11216-11222.
22. Zelenak, V., et al., Amine-modified ordered mesoporous silica: Effect of pore size
on carbon dioxide capture. Chemical Engineering Journal, 2008. 144(2): p. 336-
342.

16
Literature review August 2012

23. Brandani, F. and D.M. Ruthven, The Effect of Water on the Adsorption of CO2 and
C3H8 on Type X Zeolites. Industrial & Engineering Chemistry Research, 2004.
43(26): p. 8339-8344.
24. Li, J.-R., R.J. Kuppler, and H.-C. Zhou, ChemInform Abstract: Selective Gas
Adsorption and Separation in Metal—Organic Frameworks. ChemInform, 2009.
40(29): p. no-no.
25. Kitagawa, S., R. Kitaura, and S.-i. Noro, Functional Porous Coordination Polymers.
Angewandte Chemie International Edition, 2004. 43(18): p. 2334-2375.
26. Furukawa, H., et al., Ultrahigh Porosity in Metal-Organic Frameworks. Science.
329(5990): p. 424-428.
27. Wang, Y., S. Lin, and Y. Suzuki, Study of Limestone Calcination with CO2 Capture:
Decomposition Behavior in a CO2 Atmosphere. Energy & Fuels, 2007. 21(6): p.
3317-3321.
28. Feng, B., H. An, and E. Tan, Screening of CO2 Adsorbing Materials for Zero
Emission  Power  Generation  Systems†Energy & Fuels, 2007. 21(2): p. 426-434.
29. Kato M., Y.S.a.N.K., Carbon dioxide adsorption be lithium orthosilicate in a wide
range of temperature and carbon dioxide concentrations. Journal of Materials
Science Letters, 2002. 21(6): p. 485-487.
30. Ida, J.-i., R. Xiong, and Y.S. Lin, Synthesis and CO2 sorption properties of pure
and modified lithium zirconate. Separation and Purification Technology, 2004.
36(1): p. 41-51.
31. Yong, Z. and A.Ì.r.E. Rodrigues, Hydrotalcite-like compounds as adsorbents for
carbon dioxide. Energy Conversion and Management, 2002. 43(14): p. 1865-
1876.
32. Oliveira, E.L.G., C.A. Grande, and A.r.E. Rodrigues, CO2 sorption on hydrotalcite
and alkali-modified (K and Cs) hydrotalcites at high temperatures. Separation and
Purification Technology, 2008. 62(1): p. 137-147.
33. Samanta, A., et al., Post-Combustion CO2 Capture Using Solid Sorbents: A
Review. Industrial & Engineering Chemistry Research. 51(4): p. 1438-1463.

17
Literature review August 2012

2.0 Metal organic frameworks

The past 20 years has witnessed the emergence and advancement of a new class
of crystalline porous materials namely metal organic frameworks (MOFs). Owing to their
desirable properties and unique chemistry, these materials have become a significant
component of materials science and inorganic chemistry research. MOFs are
compromised of metal centres surrounded by a coordination sphere of multidentate
ligands which are bound to further metal atoms to form a highly ordered molecular array.
The geometry of which is defined by the coordination geometry of the metal and the
directionality of the ligands. This forms a highly stable ordered structure, held together
by a large number of coordinate bonds.
The synthesis of MOFs usually involves template synthesis using a guest molecule
to act as a molecular scaffold, bringing together the different components into close
proximity and in the required orientation. MOF synthesis follows a modular approach
allowing a wide range of metal coordination geometries and ligand choices to be used,
providing almost limitless combinations and excellent material tuneability. This control
allows facile variation of pore size and surface functionality, a difficult feat for activated
carbons or zeolite materials. Furthermore post synthetic modification has been shown to
be a viable method of altering the MOF properties, exploiting ligand based reactive
centres and free metal coordination sites. MOFs exhibit many interesting properties; they
have high porosity enabling fast mass transfer kinetics, high surface area, high internal
volume and excellent adsorption behaviour. Being highly crystalline materials MOFs can
be easily characterised, allowing the determination of important structure activity
relationships.
Rigid MOFs posses stable porous networks that do not change with external
stimuli. They have a fixed structure  that  isn’t  significantly  disrupted  by  the  introduction  
of guests. Flexible MOFs have a dynamic structure that responds to external stimuli such
as the presence of guest molecules as well as changes in pressure and temperature. This
flexibility arises due to the cooperative binding present in the structure; a large number
of weak H-bonding, p-p stacking and Van der Waals interactions. These interactions are
weak dynamic interactions, constantly breaking and reforming. The number of
interactions ensures complex stability as only a small proportion of the total population
are unbound at a particular point in time. The introduction of a significant concentration
of guest species disrupts these interactions, distorting the structure of the framework to
incorporate the guest species. This is normally a thermodynamically favourable process
as there is a net energy gain from reorganising to accommodate for the interaction of
the guest species. This flexibility is different to that observed by classes of zeolites in the
fact that it occurs under much lower temperatures and milder conditions and the MOF
structure shows a greater range of movement. Early examples of MOF structures
collapsed upon removal of host species however Yaghi et al. discovered MOF structures
that retain their porous structure in the absence of a guest species. [1]. This discovery
allowed the potential use of MOF is gas separation technologies.

18
Literature review August 2012

2.1 Relevance to carbon capture and storage

Gas separations by MOFs can be classified by either adsorption or kinetic based


separations. Adsorption based processes involve the reversible adsorption of a gases
within the pores of a substrate at a higher density than that found in the equilibrium gas
phase. These adsorbed gases are then removed by either a pressure or temperature
swing process. The principal controlling factor of such systems is the adsorption
selectivity for the target gas which is closely related to the difference s in the heat of
adsorption of the target gas and that of its nearest neighbour. Kinetic separations are
based around the differences of adsorption affinities and the diffusion rates of the
species within the porous material. Kinetic separations using MOFs have recently been
adapted to utilising the benefits of metal organic frameworks within membrane systems.
Few examples of kinetic separations using thin film MOFs exist but most involve the
separation at high pressures. [2, 3] Presently there has only been proof of concept MOF
membranes which exhibit rather poor levels of selectivity. It is suggested that as the
pore volume in MOFS, which control molecular diffusion, are significantly larger than
typical gas molecules the effect of the porous material on diffusion is limited. [4]
Much of the earlier work in this field focused on equilibrium adsorption isotherms
with later studies considering practical applications of the MOFs, by testing under more
stringent conditions. MOFs have been use as gas separation materials utilising their
tuneable pore size and to effect shape and size exclusion. However due to the very small
differences in the kinetic diameters of flue gas species this has had limited success in the
application of CCS. Material of Institute Lavoisier-96 (MIL-96) and guest free [Zn2(cnc)2
(dpt)] exhibited size selection between CO2 and CH4.[5, 6] There have been no reports
to date where CO2 has been separated from other flue gas species on the basis or size
exclusion. This is undoubtedly due to the similarity in the kinetic diameters of the flue
gas components and that of the pore system.
The governing factor of molecular selectivity in metal organic frameworks is the
adsorbate – adsorbent interactions with differences in size only having a limited
influence. This exploits the differences in polarity, quadruple moment and bonding
interactions between the adsorbate and the porous MOF structure. In MOFs both the
effects of the ligands and that of the metal centres can be exploited to induce high levels
of selectivity. An example combining all 3 levels of interaction is interaction is that of
Er2(pda)3 .[7] This MOF contains coordinatively unsaturated ErIII metal centres with 1D
circular channels with a diameter or 3.4 Å. This material exhibited selective adsorption of
CO2 over N2 thought to be due to combined effects of partial size exclusion and
electronic affects arising from the π donor ligands and charged metal centre. This
electric field is thought to induce a dipole of CO 2 furthering the adsorption energy of that
species. In addition the unsaturated ErIII metal centres could loosely bind via a donor-
acceptor mechanism aiding the separation process.
To date four distinct types of MOFs have been utilised for CO 2 separation; Those
containing vacant metal coordination sites, interpenetrated structures, dynamic
structures that change with the introduction of a host gas species and functionalised
MOFs.

19
Literature review August 2012

2.2 Metal organic frameworks with vacant coordination sites for selective CO 2
capture

Vacant metal sites within MOF structures feature protruding highly directional
orbitals which can interact strongly with molecular orbitals of other molecules such as
CO2. The strength of these interactions depends on the orbital overlap (energies)
between the two and owing to the distinctive electronic structure of molecules allows a
basis for gas separation. These co-ordinately unsaturated MOFs bind to CO2 in preference
to other gaseous species such as CH 4 and N2 owing to the higher polarisability and
quadruple moment of CO2 . To date the majority of these materials have been tested for
gas separations at high pressures not suitable for post combustion systems.[8] The
adsorption behaviour exhibited for these systems is physisorption owing to the
increasing capacity which increases with pressure but decreases with temperature. This
was also confirmed by the low to moderate enthalpy of adsorption a selection of which
are given in table 2.1. At high pressures many of these materials have exhibited
excellent CO2 capacities but many of these decrease considerably with decreasing
pressure.

Material CO2 uptake (Conditions -ΔHads ( kJ mol-1) Ref.


in parenthesis)

Cu3(btc)2 10.9 mmolg-1 (298 K, 6 30 [9]


bar)

Ni2(pbmp) 2.5 mmolg-1 (304 K, 1 35 [10]


bar)

Mg2(dobdc) 23.6 wt % (298 K, 0.1 40 [11]


atm.)

Zn3(OH)(p-cdc)2.5 0.566 mmolg-1 (298K, - [12]


0.5 bar)

Zn2(bttb) 0.4 mmolg-1 (298 K, 0.15 - [13]


bar)
-1
HCu[(Cu4Cl)3(BTTri)8(en)5] 0.366 mmolg (298 K, 90 [14]
0.06 atm.)

-1
Activated carbon- 0.5 mmolg ( 298 K, 1 14.6 [15]
MAXSORB atm.)

Table 2.1: A comparison of MOF and the known physisorbent activated carbon. This table
should only be used as a guide as differences in equipment configuration will effect the
observed capacity.

A key limitation of these studies are that they often studied high pressure
situations involving separation of CO2 from CH4. The binding mode of CO2 onto
uncoordinated metal centres has been reported for a number of complexes. [8, 16] It
was reported that the CO2 bound to the metal centre via direct end on coordination
leading to moderate binding interactions greater than those of weak physisorption (see
figure 2.1).[17] The nature of the surrounding ligands in the microporous structure
influence the amount of electron density available at the metal centre and hence control
the binding energy.

20
Literature review August 2012

Figure 2.1 : Infrared studies showing the formation of the Cr···O=C=O adduct on the
surface of [Cr3F(H2O)2O(btc)]3 (MIL-100). Similar bonding methods are thought to exist
for other metal centres.

The effect of changing the metal centre has been studied utilising the
isostructural framework [M2(dobc)(H2O)2 ](M = Ni, Co, Zn, Mg and Mn).[11] Indeed
changing the metal influences the CO2 uptake however in the case of Mg high capacities
of CO2 were reported at elevated temperatures with an initial heat of adsorption of
47kj/mol or 40kj/mol.[18] This value decreased slightly after the sites with the largest
affinity are filled first. It is thought that large increase in the sequestration capacity is a
net product of the increased character of the Mg-O bond. It is not believed to chemisorb
owing to the rigid nature of the framework preventing insertion into the Mg-O ligand
bonds, but rather the increased ionic character of this bond forms a stronger interaction
with the polarisable CO2. It should also be noted that Mg is significantly lighter than
other metals used in frameworks and this helps increase the observed %wt uptake.
Such materials with free metal coordinate sites exhibit excellent CO2 separation
capacities under dry conditions and when separating against N2. More recent work has
focused on testing these materials under more realistic conditions. [19] In the case of
[M2(dobc)(H2O)2] large reductions in CO2 adsorption capacity were reported after a
single regeneration in a humid atmosphere (see Figure 2.2). This publication highlights
the importance of screening several criteria in the application of MOFs. Chaffee et al.
studied the effect of water on the widely studied [Cu3(BTC)2(H2O)], firstly reported by
Chui and remains a widely studied MOF for gas adsorption. [20, 21] Their report
concluded that there is a strong guest host interaction with H 2O with an uptake much
higher than that of coordination alone. It was also reported a level of hysteresis during
desorption with slight structural changes as observed by X-ray diffraction and a large
reduction in CO2 adsorption capacity of 83%. Initial studies were completed at a level of
97% relative humidity (RH), when this was reduced to 30% RH the decline was noted to
be more gradual, stabilising around 75% of the original capacity. This is in agreement
with a previous study by Li who highlighted a degree of stability to water adsorption of
[Cu3 (BTC)2(H2O)] with low levels of humidity.[22]

21
Literature review August 2012

Figure 2.2: Comparison of CO2 capacities of pristine [Mg2(DOBDC)] and regenerated


[Mg2(DOBDC)]samples after exposure to simulated flue gas consisting of CO 2, N2 and
H20. The system was then purged with Argon and the samples tested with dry simulated
flue gas to determine their post-regeneration performance.

In contrast to these findings computational and experimental studies by Snurr et


al. reported that the presence of water influences the CO2 uptake by Cu3 (BTC)2(H2O).
[23]. Computational studies suggested that the presence of water should increase the
adsorption capacity. In practice up to 4% wt hydration increased the uptake
considerably with anything greater than 8% leading to a rapid reduction in capacity. The
simulations  conducted  didn’t  agree  well  with  higher  moisture  loadings  as  the  structure  
was treated as rigid allowing no change in structure for increased water content. It is
thought that the presence of water mediated the interaction between the Cu II centre and
the adsorbed CO2. Little is known whether this is an isolated phenomena or such effects
are present is other vacant coordination MOF structures. Few studies have focussed on
stability or regenerability studies of MOF with free coordination sites and even less under
testing for real flue gas conditions. It appears that the free coordination sites are highly
reactive and unless selectivity is built into the MOF they will react with water or other
flue gas species. Furthermore much of free coordinate MOF structures have focused on
ambient conditions or under higher pressures owing to the physisorption interaction.

22
Literature review August 2012

2.3 Interpenetrated metal organic frameworks for CO2 separation:

Interpenetrated structures are the result of two or more metal organic


frameworks which intercalate during synthesis to form structures made up of two or
more separate MOFs (see Figure 2.3).[24, 25] Interpenetrated MOFs are often
thermodynamically favourable due to the increased number of favourable interactions
but are kinetically slow to form. Hence long reaction times, high reagent concentrations
and high temperatures favour the formation of intercalated frameworks. Generally the
gas adsorption properties of interpenetrated species is inferior to those of individual
frameworks. These structures are highly stable and have very small pore sizes on the
order of 1-10 Å. Small pore sizes allow use in gas separations of very small gaseous
species such as CO2 , N2 and CH4 as larger molecules are unable to enter the pores[26,
27]. These materials have mostly been applied in the separation of alkanes as oppose to
smaller gases as there has been limitations in reducing the pore size[28, 29].
Separations of binary mixtures of CO2 and N2 often have low levels of selectivity
owing to the small difference in kinetic diameters of the species.[27] The selectivity is
brought about by the smaller kinetic diameter of CO 2 in comparison to N2 allowing
preferred entrance on the basis of size exclusion. The electronic effects of the ligands are
thought to interact favourably with the polarisable CO2 molecules leading to a degree of
hysteresis at low CO2 pressures.[30] As of the case of [Ni(cyclam)2(mtb)]
interpenetrated frameworks have low surface areas but exhibit selectivity for CO2 on the
basis of size exclusion .[31] Intercalated MOFs are still at a very early stage for gas
separation technologies as few successfully example exist for the separation of small gas
molecules. A key limitation to this area is the lack of information regarding stability and
regeneration

Figure 2.3: Crystal structures of the lone framework a and the interpenetrated
framework b.

23
Literature review August 2012

2.4 Flexible frameworks for selective CO2 separation:

Metal organic frameworks compromise a number of coordinate bonds which are


much weaker than rigid inorganic structures such as zeolites. The linkers between metal
centres are less constrained leading to an overall structural flexibility. This flexibility
provides MOF with fundamentally different mechanical and physical properties. Dynamic
MOFs undergo reversible structural transformations due to physical or chemical stimuli.
Upon removal of the guest species of 1 st generation MOFs the structure would collapse
but reform with reintroduction of the guest species. Further developments led to the
design of rigid frameworks that retained their structure upon guest removal and are
utilised as adsorbents. Dynamic MOF undergo a reversible structure transformation upon
addition and removal of a guest species. [32]
In relevance to gas adsorption these materials exhibit highly interesting
behaviour; at distinct pressures of guest the material expands as the gas molecule
enters the pores.  The  extent  of  this  ‘opening’  is  dependent  upon  the  material  as  in  some  
cases little adsorption occurs until the pores become accessible (see figure 2.4).[33]
Flexible MOFs characteristically exhibit hysteresis upon adsorption/desorption isotherms
the extent of which is dependent upon the nature of the material and cycling
temperature.[34] The gate mechanism of opening has only been studied under single
isotherms and it is expected that the materials may be poor for separating mixtures of
gases. Long suggested that if a single component triggers opening of the framework, the
pores will be unable to discriminate against species entering the structure. [35] At
present flexible MOFs have little potential for application in post combustion capture as
their selectivity is based on size and ligand interactions discussed previously.

Figure 2.4: The multiple component CO2 adsorption isotherms for SNU-M10. The material
initially adsorbs very little CO2 but once a gate opening pressure is reached CO2 uptake
increases rapidly.

24
Literature review August 2012

2.5 Functionalised metal organic frameworks for selective CO2 capture

To improve the CO2 capture capacity of MOFs at low pressures post synthesis
modification has been employed. This entails coordinating ligands with embedded
functionalisation imparting a high affinity towards CO2 onto vacant metal centres. This
post synthesis modification is similar to amine grafting but allows a much finer level of
molecular control. Owing to the numerous MOF with free coordination sites plentiful new
materials can be formed. As described previously the Cu MOF with vacant metal
coordination sites can be used as CO2 capture materials but suffer from low capacities
flue gas concentrations or CO2 . Long et al. functionalised the water stable sodalite type
framework [Cu-BTTri] which features vacant Cu coordination sites with ethylenediamine
(en). [14]. Alkylamine functionalisation led to a reduction is surface area but an
increased CO2 uptake capacity at low partial pressures of CO2. Functionalisation also
increased the CO2 / N2 selectivity and the material displayed an isoteric heat of
adsorption of 90 kJ/mol-1 indicative of chemisorptions. Cyclic testing was conducted in a
15% CO2 / N2 stream with the adsorbent showing hysteresis and gradual loss of capacity
after 10 cycles with a N2 purge. The choice of MOF influences the final pore size after
modification which in turn influences the observed kinetics.
As with liquid amine systems the choice of amine effects the heat of adsorption.
Recent work by Long led to the development of [PE-M2(dobpdc)] an expanded variant of
[M2(dobdc)] highlighted earlier, M=Mg.[36] The related framework M2(dobdc) has
smaller pores and it was believed that this hampered diffusion of the diamines as
materials of this type were unable to be synthesised. This material exhibits good low
temperature CO2 adsorption at temperatures up to 75°C (see figure2.6). The cyclic
working capacity from dry flue gas (15% CO2 85% N2 ) was reported as 2.52 mmolg-1
with the sample cycled between 40-120°C. No studies were undertaken on moisture
stability but given the rapid decomposition in a humid atmosphere of Mg(dobdc) one
predicts this compound to reach a similar fate. The reason for this lack of stability is the
completive coordination of water onto the reactive metal centres. Applying the Pearson
acid base concept oxygen atoms in water will bind more favourably to the hard Mg
centres as oppose to the amines; the shear concentration differences between the two
also favours water coordination

Figure 2.5: The functionalisation of Mg2(dobpdc) by coordination of enamine ligands.

25
Literature review August 2012

Figure 2.6: Rapid adsorption kinetics of en-Mg2(dobpc) showing high CO2 capacities at
low partial pressures.

As well as functionalising with amines other molecules with an affinity for CO2
have also been used. Snurr et al. enhanced the selectivity of [Zn2(bttb)(py-CF3)2] by
displacing coordinated solvent molecules with the highly polar py-CF3 moiety. [13] This
functionalisation imparted vastly improved selectivity against N2 but poor uptakes at low
partial pressures of CO2 relative to other adsorbents.
Further efforts to impart greater selectivity and better adsorption properties of
gases have lead to the development of MOFs with added functionality. Synthetic
modification of the structure by covalent modification of reactive ligand sites have been
used to further functionalise the surface. Introduction of functional groups to MOF can be
challenging as the groups can interfere with MOF synthesis if present during formation or
methods of post synthetic treatment may destroy the structure. [37] In spite of this
many examples of MOF post synthetic modification have been reported for varying
applications.[38-40]. In relevance to CO2 adsorption there are only a limited number of
examples of successful post synthetic modification enhancing gas separation. Cohen et al.
reported that the gate opening effect in flexible MOF could be controlled by post
synthetic covalent functionalisation of [Zn2((2-amino-1,4-
benezenedicarboxylate)(DABCO)) with anhydrides and methylaziridines.[41] Modification
was found to lead to an expected reduction in surface area and the onset of gate
opening at a lower pressure. CO2 isotherms were only conducted at 196K with no data on
other temperatures reported. At such low temperature the primary and secondary
amines present would not have enough energy to react with CO2 hence chemisorption
would have not been observed.

Alternative work by Jeong et al. highlighted a different approach; membrane


separation of propane and CO2 using functionalised (Zn4O(BDC-NH2)3) membranes.[42].
They observed that amine functionalised MOF imparted a great selectivity for CO 2 owing
to the favourable interaction. They functionalised the MOF with an aniline unit meaning
that the N lone pair is delocalised within the ring structure prohibiting chemisorption of
the CO2 . Computational studies by Torrisi et al. illustrated that functional group
modification of MIL-53 could enhance low pressure CO2 adsorption.[43] They suggested
that OH functionalised MIL-53 should have the best low temperature adsorption capacity.
Further work by Biswas synthesized a range of MIL-X complexes and studied their

26
Literature review August 2012

adsorption behaviour .[44] In contrast to experimental prediction they found that MIL-
53-NH2 to capture the highest amount of CO2 with MIL-53-OH having the lowest capacity.
MILs are known to exhibit breathing effects dependant on the incorporated metal ion and
it was suggested that the computational study assumed an open structure whereas at
the low pressures used the structure was closed. The adsorption isotherms are given
below in Figure 2.7 and shown moderate levels of single component CO2 adsorption. It
was proposed that the pore opening is influenced by the intraframework hydrogen
bonding.

Figure 2.7: CO2 adsorption isotherms of MIL-53-X, 1-Cl (black, circles), 2-Br (blue,
squares), 3-CH3 (green, triangles), 4-NO2 (red, stars), 5-(OH)2 (magenta, pentagons),
and Al-MIL-53-NH2 (violet, hexagons) measured at 25 °C.

Recently work by Xiang et al. conducted covalent modification of very high


surface area UMCM-1-NH with maleic anhydride for use in post combustion carbon
capture (see figure 2.8).[45] The resulting functionalization lead to the formation of a
conjugated amide alkene carboxylic acid bound to the framework. This was selected as it
was believed to provide a stronger interaction with CO2 based on computational studies,
however experimental adsorption isotherms revealed that amine functionalised materials
captured a greater capacity of CO2 at high pressures but the carboxylic acid
functionalised material had better capacity at low pressures suitable for post combustion
capture. As to be expected materials synthesised still exhibited weak physisorption and
only and have little suitability towards post combustion carbon capture.

27
Literature review August 2012

Figure 2.8: The post synthetic modification strategy used by Xiang et al. The reactions
used are directly analogous to those conducted in solution with standard organic
chemistry

The vast majority of MOF structures discussed previously exhibit hydroscopic


structures and are often highly unstable requiring storage under inert atmospheres.
Zeolite imidazolate frameworks (ZIFs) are a recent addition to the MOF family and are
both thermally and chemically stable able to retain their structure and properties after
refluxing aqueous acid.[46] ZIFs are isostructural with zeolite frameworks by
substituting tetrahedral SiIV and AlIII found in zeolites with similarly tetrahedral ZnII or
CoII. Bridging oxygen anions are replaced with imidazolate ligands (see Figure 2.9).[47]
In contrast to zeolites the structural units are determined by the linker interactions as
appose to structural directing agents used in zeolite preparation. These ZIF have
tuneable pore sizes, large surface area, large pore volumes and can be easily be
functionalised using ligand modification.

Figure: 2.9: A range of imidazole based ligands that have been used to synthesis ZIFs
for CO2 capture. The wide range of functional groups tolerated allows many options for
post-synthetic modification.

28
Literature review August 2012

ZIFs have been show to interact favourably with CO2 in binary gas mixture
breakthrough experiments and have improved performance compared to zeolite 13X and
BPL carbon under similar conditions.[48] At present most of the functionisation has
focused on tuning the electronic nature of the structure to encourage stronger
physisorption interactions. The excellent stability of ZIFs under a range of conditions has
led to covalent post synthetic modification in an effort to induce chemisorbents
behaviour. Morris et successfully functionalised [Zn(ICA)2] (ZIF-90) by reaction with the
pendant aldehyde group shown below in figure 2.10[49]. The framework was remained
intact during the functionalisation with MEA however the pores became blocked with the
majority of functionalisation occurring on the most available sites. If materials with
larger pore sized could be used then MOF chemisorbents could be produced. Owing to
the unique nature of the support such materials would exhibit highly interesting
properties.

Figure 2.10: Post synthetic functionalisation routes of ZIF-90. Left, the reduction with
sodium borohydride and right, reaction with monoethanolamine.

29
Literature review August 2012

2.6 Potential of metal organic frameworks for use in post combustion capture:

The use of MOF solid adsorbents in post combustion capture relies on the
fulfilment of the criteria highlighted in section 1. The materials require an adsorption
capacity of at least 3mmol/g at a CO2 partial pressure of 0.15 bar with a high CO2 /N2
selectivity. Figure 2.11 highlights that numerous well known examples of MOFs such as
Cu-BTC, ZIF 8, ZIF-78, ZIF-82, Zn-MOF, Co carboranes and a biological based MOF fail
to meet this capacity.[50] The isotherms show in Figure 2.11 show that the CPO-27
family of MOF surpass the criteria for capacity due to the strong interaction between CO 2
and the Mg or Ni vacant coordination sites. [11, 51] Initially this isotherm is steep as the
free metal sites are preferentially bound, whilst enabling a high capacity at low pressure
it makes regeneration problematic. The stability of such frameworks is also a concern as
low partial pressures of H2O quickly led to saturation of the open metal sites and
reducing the adsorption capacity of the sample. The linear isotherms ZIF 8 and Cu BTC
highlight the prevalent weak physisorption which is highly dependent on partial pressure,
whilst being easy to regenerate these materials lack the required capacity.[52] Of
interest bioMOF11 a biologically derived framework consisting of Cobalt and adenine
ligands has a greater capacity than the ZIF materials with a promising curved
isotherm.[53]
The level of regenerative stability and working capacity is a major barrier for MOF
applications as these materials have yet to be tested in cyclic studies with few reports of
multicomponent adsorption. A significant and often overlooked barrier to the mainstream
application of MOF is the use of expensive and rare late transition metals. There is not
sufficient global supplies to produce the materials required and their use cannot be
justified in the presence of well performing cheaper materials.

Figure 2.11: CO2 isotherms for selected MOFs at 298 K at low partial pressures of CO 2 .
[50]

30
Literature review August 2012

2.7 References

1. Li, H., et al., Design and synthesis of an exceptionally stable and highly porous
metal-organic framework. Nature, 1999. 402(6759): p. 276-279.
2. Liu, Y., et al., Synthesis of continuous MOF-5 membranes on porous alumina
substrates. Microporous and Mesoporous Materials, 2009. 118(13): p. 296-301.
3. Bux, H., et al., Zeolitic Imidazolate Framework Membrane with Molecular Sieving
Properties by Microwave-Assisted Solvothermal Synthesis. Journal of the
American Chemical Society, 2009. 131(44): p. 16000-16001.
4. Keskin, S. and D.S. Sholl, Assessment of a Metal Organic Framework Membrane
for Gas Separations Using Atomically Detailed Calculations: CO2, CH4, N2, H2
Mixtures in MOF-5. Industrial & Engineering Chemistry Research, 2008. 48(2): p.
914-922.
5. Loiseau, T., et al., MIL-96, a Porous Aluminum Trimesate 3D Structure
Constructed from a Hexagonal Network of 18-Membered Rings and 3-Oxo-
Centered Trinuclear Units. Journal of the American Chemical Society, 2006.
128(31): p. 10223-10230.
6. Xue, M., et al., Robust Metal Organic Framework Enforced by Triple-Framework
Interpenetration Exhibiting High H2 Storage Density. Inorganic Chemistry, 2008.
47(15): p. 6825-6828.
7. Pan, L., et al., Porous Lanthanide-Organic Frameworks: Synthesis,
Characterization, and Unprecedented Gas Adsorption Properties. Journal of the
American Chemical Society, 2003. 125(10): p. 3062-3067.
8. Llewellyn, P.L., et al., High Uptakes of CO2 and CH4 in Mesoporous Metal Organic
Frameworks MIL-100 and MIL-101. Langmuir, 2008. 24(14): p. 7245-7250.
9. Min Wang, Q., et al., Metallo-organic molecular sieve for gas separation and
purification. Microporous and Mesoporous Materials, 2002. 55(2): p. 217-230.
10. Miller, S.R., et al., Structural Transformations and Adsorption of Fuel-Related
Gases of a Structurally Responsive Nickel Phosphonate Metal Organic Framework,
Ni-STA-12. Journal of the American Chemical Society, 2008. 130(47): p. 15967-
15981.
11. Dietzel, P.D.C., V. Besikiotis, and R. Blom, Application of metal-organic
frameworks with coordinatively unsaturated metal sites in storage and separation
of methane and carbon dioxide. Journal of Materials Chemistry, 2009. 19(39): p.
7362-7370.
12. Bae, Y.-S., et al., Separation of CO2 from CH4 Using Mixed-Ligand Metal Organic
Frameworks. Langmuir, 2008. 24(16): p. 8592-8598.
13. Bae, Y.-S., et al., Enhancement of CO2/N2 selectivity in a metal-organic
framework by cavity modification. Journal of Materials Chemistry, 2009. 19(15):
p. 2131-2134.
14. Demessence, A., et al., Strong CO2 Binding in a Water-Stable, Triazolate-Bridged
Metal Organic Framework Functionalized with Ethylenediamine. Journal of the
American Chemical Society, 2009. 131(25): p. 8784-8786.
15. Dreisbach F., S.R., Keller J. U., High pressure adsorption data of methane,
nitrogen, carbon Dioxide and their binary and ternary mixtures on activated
carbon. Adsorption, 1999. 5(3): p. 215-227.
16. Bordiga, S., et al., Adsorption properties of HKUST-1 toward hydrogen and other
small molecules monitored by IR. Physical Chemistry Chemical Physics, 2007.
9(21): p. 2676-2685.
17. Llewellyn, P.L., et al., High uptakes of CO2 and CH4 in mesoporous metal-organic
frameworks MIL-100 and MIL-101. Langmuir, 2008. 24(14): p. 7245-7250.
18. Caskey, S.R., A.G. Wong-Foy, and A.J. Matzger, Dramatic Tuning of Carbon
Dioxide Uptake via Metal Substitution in a Coordination Polymer with Cylindrical
Pores. Journal of the American Chemical Society, 2008. 130(33): p. 10870-
10871.

31
Literature review August 2012

19. Kizzie, A.C., A.G. Wong-Foy, and A.J. Matzger, Effect of Humidity on the
Performance of Microporous Coordination Polymers as Adsorbents for CO2
Capture. Langmuir. 27(10): p. 6368-6373.
20. Liang, Z., M. Marshall, and A.L. Chaffee, CO2 Adsorption-Based Separation by
Metal Organic Framework (Cu-BTC) versus Zeolite (13X). Energy & Fuels, 2009.
23(5): p. 2785-2789.
21. Chui, S.S.-Y., et al., A Chemically Functionalizable Nanoporous Material
[Cu3(TMA)2(H2O)3]n. Science, 1999. 283(5405): p. 1148-1150.
22. Li, Y. and R.T. Yang, Hydrogen storage in metal-organic and covalent-organic
frameworks by spillover. AIChE Journal, 2008. 54(1): p. 269-279.
23. Yazayda A., et al., Enhanced CO2 Adsorption in Metal-Organic Frameworks via
Occupation of Open-Metal Sites by Coordinated Water Molecules. Chemistry of
Materials, 2009. 21(8): p. 1425-1430.
24. Chen, B., et al., Rationally  Designed  Micropores  within  a  Metal−Organic  
Framework for Selective Sorption of Gas Molecules. Inorganic Chemistry, 2007.
46(4): p. 1233-1236.
25. Chen, B., et al., A Triply Interpenetrated Microporous Metal Organic Framework
for Selective Sorption of Gas Molecules. Inorganic Chemistry, 2007. 46(21): p.
8490-8492.
26. Dybtsev, D.N., et al., Microporous Manganese Formate: A Simple Metal Organic
Porous Material with High Framework Stability and Highly Selective Gas Sorption
Properties. Journal of the American Chemical Society, 2003. 126(1): p. 32-33.
27. Yao, Q., et al., Interpenetrated metal-organic frameworks and their uptake of
CO2 at relatively low pressures. Journal of Materials Chemistry. 22(20): p.
10345-10351.
28. Chen, B., et al., A Microporous Metal–Organic Framework for Gas-
Chromatographic Separation of Alkanes. Angewandte Chemie International
Edition, 2006. 45(9): p. 1390-1393.
29. Bárcia, P.S., et al., Kinetic Separation of Hexane Isomers by Fixed-Bed
Adsorption with a Microporous Metal Organic Framework. The Journal of Physical
Chemistry B, 2007. 111(22): p. 6101-6103.
30. Yang, S., et al., A partially interpenetrated metal organic framework for selective
hysteretic sorption of carbon dioxide. Nat Mater. advance online publication.
31. Cheon, Y.E. and M.P. Suh, Multifunctional Fourfold Interpenetrating Diamondoid
Network: Gas Separation and Fabrication of Palladium Nanoparticles. Chemistry –
A European Journal, 2008. 14(13): p. 3961-3967.
32. Horike, S., S. Shimomura, and S. Kitagawa, Soft porous crystals. Nat Chem,
2009. 1(9): p. 695-704.
33. Suh, M.P., Y.E. Cheon, and E.Y. Lee, Syntheses and functions of porous
metallosupramolecular networks. Coordination Chemistry Reviews, 2008. 252(9):
p. 1007-1026.
34. Choi, H.-S. and M.P. Suh, Highly Selective CO2 Capture in Flexible 3D
Coordination Polymer Networks. Angewandte Chemie International Edition, 2009.
48(37): p. 6865-6869.
35. D'Alessandro, D.M., B. Smit, and J.R. Long, Carbon Dioxide Capture: Prospects
for New Materials. Angewandte Chemie International Edition. 49(35): p. 6058-
6082.
36. McDonald, T.M., et al., Capture of Carbon Dioxide from Air and Flue Gas in the
Alkylamine-Appended Metal Organic Framework mmen-Mg2(dobpdc). Journal of
the American Chemical Society. 134(16): p. 7056-7065.
37. Wang, Z. and S.M. Cohen, Postsynthetic modification of metal-organic
frameworks. Chemical Society Reviews, 2009. 38(5): p. 1315-1329.
38. Burrows, A.D., et al., Post-Synthetic Modification of Tagged Metal-Organic
Frameworks. Angewandte Chemie-International Edition, 2008. 47(44): p. 8482-
8486.

32
Literature review August 2012

39. Savonnet, M., et al., Generic Postfunctionalization Route from Amino-Derived


Metal-Organic Frameworks. Journal of the American Chemical Society. 132(13):
p. 4518-4521.
40. Cohen, S.M., Postsynthetic Methods for the Functionalization of Metal-Organic
Frameworks. Chemical Reviews. 112(2): p. 970-1000.
41. Wang, Z. and S.M. Cohen, Modulating Metal Organic Frameworks To Breathe: A
Postsynthetic Covalent Modification Approach. Journal of the American Chemical
Society, 2009. 131(46): p. 16675-16677.
42. Yoo, Y., V. Varela-Guerrero, and H.-K. Jeong, Isoreticular Metal Organic
Frameworks and Their Membranes with Enhanced Crack Resistance and Moisture
Stability by Surfactant-Assisted Drying. Langmuir. 27(6): p. 2652-2657.
43. Torrisi, A., R.G. Bell, and C. Mellot-Draznieks, Functionalized MOFs for Enhanced
CO2 Capture. Crystal Growth & Design. 10(7): p. 2839-2841.
44. Biswas, S., T. Ahnfeldt, and N. Stock, New Functionalized Flexible Al-MIL-53-X (X
= -Cl, -Br, -CH3, -NO2, -(OH)2) Solids: Syntheses, Characterization, Sorption,
and Breathing Behavior. Inorganic Chemistry. 50(19): p. 9518-9526.
45. Xiang, Z., S. Leng, and D. Cao, Functional Group Modification of Metal Organic
Frameworks for CO2 Capture. The Journal of Physical Chemistry C. 116(19): p.
10573-10579.
46. Hayashi, H., et al., Zeolite A imidazolate frameworks. Nat Mater, 2007. 6(7): p.
501-506.
47. Phan, A., et al., Synthesis, Structure, and Carbon Dioxide Capture Properties of
Zeolitic Imidazolate Frameworks. Acc. Chem. Res. 43(1): p. 58-67.
48. Wang, B., et al., Colossal cages in zeolitic imidazolate frameworks as selective
carbon dioxide reservoirs. Nature, 2008. 453(7192): p. 207-211.
49. Morris, W., et al., Crystals as Molecules: Postsynthesis Covalent Functionalization
of Zeolitic Imidazolate Frameworks. J. Am. Chem. Soc., 2008. 130(38): p.
12626-12627.
50. Pirngruber, G.D. and P.L. Llewellyn, Opportunities for MOFs in CO2 Capture from
Flue Gases, Natural Gas, and Syngas by Adsorption, in Metal-Organic Frameworks,
Wiley-VCH Verlag GmbH & Co. KGaA. p. 99-119.
51. Yazayd, A., et al., Screening of Metal Organic Frameworks for Carbon Dioxide
Capture from Flue Gas Using a Combined Experimental and Modeling Approach.
Journal of the American Chemical Society, 2009. 131(51): p. 18198-18199.
52. Pérez-Pellitero, J., et al., Adsorption of CO2, CH4, and N2 on Zeolitic Imidazolate
Frameworks: Experiments and Simulations. Chemistry – A European Journal.
16(5): p. 1560-1571.
53. An, J., S.J. Geib, and N.L. Rosi, High and Selective CO2 Uptake in a Cobalt
Adeninate Metal Organic Framework Exhibiting Pyrimidine- and Amino-Decorated
Pores. Journal of the American Chemical Society, 2009. 132(1): p. 38-39.

33
Literature review August 2012

3.0 Solid supported amines:

Supported amine sorbents have been separated into 3 classes of adsorbents. [1]
Class 1 amines are immobilised on a range of solid supports impregnation without the
formation of a covalent bond. Class 2 amine sorbents are cases where the amine is
covalently tethered to the solid support. Class 3 amine adsorbent are solid porous
supports in which amine containing polymers are polymerised in situ using a covalently
bound monomer.

3.1 Class 1: Physically adsorbed amines

Physically adsorbed amines on a range of solid supports can be synthesised by


utilising the straightforward wet impregnation technique. A general synthetic strategy is
so suspend a solid support in a solution containing the desired amine and an inert
volatile solvent such as ethanol. Subsequent removal of the solvent leads to
immobilisation of the amine within the porous structure of the support or coating of the
outer surface. In relation to carbon capture immobilisation within the porous structure is
preferred as deposition on the exterior of the particles can lead the formation of to
agglomerates. In principal the amine density is controlled by the synthetic stoichiometry,
however this is often not the case as control is limited. Deviation between the expected
and observed amine density can be caused by the number of accessible amines.
Inaccessible amines are those unable to capture CO2 due to either steric constrains or
chemical deactivation.

Figure 3.1: Impregnation of PEI on a porous molecular support to yield a molecular


basket sorbent

Physically adsorbed amines immobilised on a high surface area support were first
reported by Song et al. and termed molecular basket adsorbents(see figure 3.1).[2] In
such molecular baskets the sterically branched polymer polyethylenimine (PEI) is
impregnated within the pores of the porous support Mobil composition of matter no. 41
(MCM-41). PEI was selected as branched amines have lower regeneration requirements
owing to the steric hindrance between the bulky polymer chains and adsorbed gases.
Furthermore the immobilised PEI had a very low vapour pressure and decomposed
above 150 °C, a temperature above most post combustion capture systems.
Functionalisation was straightforward via wet impregnation of PEI in methanol and led to
no degradation of the MCM-41 support. PEI-MCM-41 led to a marked increase in CO2
uptake compared to MCM-41 or PEI alone. This was attributed to the synergistic effects
of utilising a high surface area support and a material that strongly interacts with CO2.

34
Literature review August 2012

A high surface area support allows a greater dispersion of PEI exposing more CO2
adsorbent sites. Interestingly the largest theoretical amount of PEI loading was 50%
however MCM-41-PEI with a PEI loading of 75% yielded the highest CO 2 uptake capacity.
This suggests that the PEI forms mutlilayers within the pores as agglomeration was not
observed at this level of loading. A further key point to note is that at best, 44% of the
available amine reacted with CO2 suggesting a large extent of inaccessible or deactivated
amines. The authors reported that the capacity increases with temperature to a set
extent, seemingly contrasting the notion that an exothermic reaction between amines
and CO2 would be retarded by an increase in temperature. They proposed that a diffusion
limiting process is underway and at low temperatures access to the sites is restricted.
Moving to a higher temperature regime leads to a breakdown of the nanoparticle
structures within the pores leading to an increase in the number of sites available for
CO2 adsorption (see figure 3.2). Under dry CO2 the material had an equilibrium capture
capacity of 2.27mmolg-1 with a working capacity of 1.59mmolg-1 under pure CO2 with a
nitrogen purge. Initial studies found that the adsorbents were stable to seven cycles of
regeneration with no loses in CO2 uptake capacity reported.

Figure 3.2: The differences between the (A) low temperature and (B) high temperature
regime of PEI inside molecular basket adsorbents.

Further to this Xu et al. explored the effects of material preparation in a bid to


bolster the capture capacity.[3] It was found that the impregnation solvent had little
effect on the dispersion of the PEI although water proved troublesome to remove,
leading to support breakdown. Addition of polyethylene glycol (PEG) led to an increase in
capture capacity as well as an increase in adsorption and desorption kinetics. This was
also  the  case  with  Satyapal  modified  NASA’s  HSC+  solid  amine  adsorbent.  [4] They
found that PEG attracts more water enabling a faster reaction and a larger capacity. This
was also reported by Chaffee et al. who showed that the ratio of CO2 per available N
atom doubled with in the presence of a nearby hydroxyl group. [5] They hypothesised
that in the presence of a hydroxyl group reactions with CO2 proceed via a different
mechanism, one not dissimilar to the mechanism in the presence of water. To date much
work has been undertaken towards the development of impregnated amines with the
principal findings discussed within this work.

35
Literature review August 2012

3.1.1 Effect of water:

It is widely accepted that in the presence of water the amine efficiency should be
increased from 0.5 to 1.0 owing to the formation of carbonates. Many adsorbents
competitively adsorb water and CO2 with the water being difficult to remove. Hicks et al.
highlighted the potential problem of water causing increased levels of leeching.[6] There
are many conflicting reports on the role of water; Song et al. studied the effect of
different concentrations on water on the CO2 adsorption of MCM-41-PEI. [7] They found
that the presence of water led to an increase in breakthrough time suggesting more CO 2
is adsorbed . With a concentration of water lower than that of CO2, the capacity
increases with the increasing moisture content. When the moisture concentration
surpasses that of the CO2 the adsorption capacity does not increase further. A maximum
is found at a water content almost equal to the CO 2 concentration. The CO2 is thought to
dissolve into water as part of the adsorption mechanism, hence the greater the amount
of water present the higher the amount of CO2 that can be solvated . In contrast Zhu et
al. reported that the presence of moisture in the flue gas led to a reduction in CO 2
capture capacity for a binary mixture of TEPA and DEA on a SBA-15 support. [8] This
material was modified to contain a free hydroxyl group within the pores to act as a
proton source during carbamate formation. It appears that the presence of this hydroxyl
group has a high affinity for water leading to a level of competitive adsorption between
water andCO2 .

3.1.2 Types of amine

Many amines are suitable for wet impregnation onto the surface and within the
pores of solid supports. Materials with high weight percentage of nitrogen are preferred
as these will act as sites for CO2 adsorption. This is not the only consideration to bear in
mind; the kinetics, impregnation behaviour, stability and heats of adsorption are all
crucial factors. The most commonly used material being low molecular weight PEI either
in linear or branched form. PEI contains a high amine weight consisting of a mixture of
1°, 2°, 3° amines often assumed to be in a near 1:1:1 ratio. [9] However studies by
Jones and Kissel reported that PEI contains a higher proportion of primary amines at the
expense of tertiary amines, with the exact amount dependent on synthesis conditions. [6,
10]. Work by Drage et al., studied the effect of molecular weight upon PEI regeneration
on a silica support. [11] It was reported that higher molecular weight PEI retained a
higher level of capacity after regeneration, with 423MM PEI losing mass in regeneration
under nitrogen. However there is an upper limit towards molecular weight substrates as
very high molecular weight polymers may not enter pores and causes blockages leading
to a reduction in capacity.

Figure 3.3 : Amines used in the synthesis of solid support amine adsorbents.

36
Literature review August 2012

PEI is by no means the only amine utilised to impregnated amine sorbents, a


wide range of polymeric and monomer amines have been reported in physisorbed amine
sorbents. Others that have been studied included diethanolamine(DEA) , and
tetraethylenepentamine(TEPA) (show above if figure 3.3). [12, 13] These materials have
a higher vapour pressure than PEI but still exhibit good stability. This could be attributed
to higher proportion of strongly interacting nitrogen atoms present in the molecules.

3.1.3 Kinetics

Kinetics are a crucial factor for all adsorbents as they influence the amount of
adsorbent required by a process. Impregnated adsorbents typically display a rapid
adsorption phase levelling off with slow approach to equilibrium as a result of mass
transfer limitations. The amount of amine loading influences the kinetics of the material
as in low loading concentrations CO2 adsorption is limited by the availability of the amine.
At high loadings the maximum adsorption rate is increased owing to increasing
availability of amines but, at such loadings the diffusional resistance becomes a limiting
factor. This was highlighted in the work by Sayari who noted a maximum in the rate of
adsorption just above the theoretical pore saturation point. [13]. The pore size is
another crucial factor as KIT-6 with a much larger pore size than MCM-41 when
impregnated with PEI reached 70% of its capacity approximately six times faster than
PEI-MCM-41.[14] Generally the larger the pore size the better the mass transport and
hence faster kinetics. As highlighted in Figure 3.4, the physical form of the adsorbent,
being linked to mass transport influences the breakthrough time of the molecular basket
adsorbents developed by XU et al.[15]

Figure 3.4: Xu et al. noted a reduction in breakthrough time when changing the
morphology of the sorbent from a pellet to a powder.

37
Literature review August 2012

3.1.4 Capacity

Given the long times required for amine impregnated adsorbents to reach
equilibrium the equilibrium capacity is often a redundant measurement. With this in
mind working capacities are better measure of the achievable capacity of an adsorbent.
This is defined as the difference in capacity between the adsorbed and desorbed capacity
for a limited short cycle time. The working capacity of impregnated amine adsorbents is
found to vary with the operating temperature, gas composition and physical form of the
adsorbent. Schuette et al. evaluated that for solid adsorbents to be cost competitive with
MEA a sorbent capacity of less than 3 mmolCO2 g-1 will favour MEA whilst solid
adsorbents favoured with capacities above 4 mmolg-1.[16]
Class 1 adsorbents exhibit many of the highest capacities to date, which can be
attributed to their respective high levels of amine loading. This high amine loading is
typically accompanied with low amine efficiency, hence for optimum capacity a balance
between amine loading and utilisation is required. Table 3.1 below highlights significant
capacities from the literature. It should be noted that the measurement technique and
gas conditions all influence the observed capacity and this table should be used as a
guide. Measurement of capacity under humid conditions is particularly problematic using
gravimetric methods as the uptake of CO2 and H2O is difficult to separate.

38
Literature review August 2012

Amine Support Capacity Conditions Amine Method Ref


(mmolCO2 g-1) efficiency

PEI Proprietary 2.4 Dry, T= 0.26 TGA [4]


inorganic 348K, PCO2
support = 1bar

DEA PE-MCM- 2.93 Dry, T= 0.4 TGA [13]


41 298K, PCO2
= 0.05bar

PEI MCM-41 3.02 Dry, T= 0.17 TGA [2]


348K, PCO2
= 1bar

PEI MCM-41 2.02 Dry, T= 0.17 GC [7]


348K, PCO2
= 0.15bar

PEI KIT-6 3.07 Dry, T= 0.26 TGA [17]


348K, PCO2
= 1bar

PEI MCM-41 3.02 Dry, T= 0.17 TGA [3]


348K, PCO2
= 1bar

PEI MCM-41 2.55 Humid, T= 0.22 GC [15]


348K, PCO2
= 1bar

TEPA SBA-15 3.93 Dry, T= 0.30 TPD/GC [12]


348K, PCO2
= 1bar

DEA/TEPA SBA-15 4.00 Dry, T= 0.37 TPD/GC [8]


348K, PCO2
= 1bar

DEA/TEPA SBA-15 3.18 Humid, T= 0.3 TPD/GC [8]


348K, PCO2
= 1bar

TEPA MCM-41 5.02 Dry, T= 0.32 TPD/GC [18]


308K, PCO2
= 1bar

PEI SBA-15 3.2 Dry, T= - GC [19]


348K, PCO2
= 0.15bar

TEPA MC400 7.93 Humid, T= - GC-MS [20]


348K, PCO2
= 0.1bar

Table 3.1: CO2 uptake capacities and measurement conditions for class 1 adsorbents.

39
Literature review August 2012

3.1.5 Support effects

A wide range of supports can be utilised for the immobilisation of amines with
highly porous materials preferred owing to their high surface areas and large surface
areas. Physisorbed molecules are weakly bound to the surface of support and as a result
can be easily removed. However, materials contained within porous networks exhibit a
much larger degree of support due to the many simultaneous interacts between the
molecule and the pore surface. Most of the supports studied to date are highly porous
silicas such as SBA-15, MCM-48 with the surface silanols having a strong interaction with
impregnated amines. A study by Son et al. synthesised a range of mesoporous silica
supports and impregnated them with 50wt % PEI (see Table 3.2).[17] A general trend
emerged that increasing capacity followed the average pore diameter of the material.
Increasing pore diameter also correlated to an increase in adsorption kinetics.
Supported PEI on KIT-6 lead to a reduction in decomposition temperature for sub
60% loadings of PEI, a similar phenomenon was observed by Xu et al.[3] This reduction
is decomposition temperature was attributed to a small PEI particle size having a higher
volatility. [21] With higher loadings ~70 %wt. PEI was found in increasing amounts on
the surface or the support where it behaved akin to bulk PEI.

Material BET Surface area (m 2/g) Pore volume (cm3/g) Pore diameter (BJH)
(nm)

MCM-41 1042 0.85 2.8

MCM-41-PEI 50 4 0.01 -

MCM-48 1162 1.17 3.1

MCM-48-PEI 50 26 0.10 -

SBA-15 753 0.94 5.5

SBA-15 PEI 50 13 0.04 -

SBA-16 736 0.75 4.1

SBA-16 PEI 50 23 0.0 -

KIT-6 895 1.22 6.0

KIT-6-PEI 10 430 0.68 5.9

KIT-6-PEI 30 251 0.49 5.4

KIT-6-PEI 50 86 0.18 5.3

Table 3.2: Surface properties of a range of materials relevant to solid adsorbents.

40
Literature review August 2012

With this information to hand Sayari et al. developed pore expanded MCM-41 (PE-
MCM-41) which has a lower surface area but triple the average pore size and double the
internal volume of conventional MCM-41 from 4 nm and 0.8 cm3/g up to 25 nm and 3.5
cm3/g respectively.[22]. This PE-MCM-41 was then tested as a low temperature
adsorbent for CO2 capture by wet impregnation of DEA.[13] The materials exhibited
improved capture capacity and kinetics over that of as prepared PEI-MCM-41. The
enhanced capacity of PE-MCM-41-DEA was attributed to the larger pore volume leading
to a higher loading of DEA. A slight loss of capacity was observed with regeneration. It
was assumed to be gradual leeching of DEA although as no DEA was detected by TGA-
MS the loss could be due to uncharacterised amine degradation. Migration outside of the
pores but still onto the sample pan could also lead to the observed reduction in capacity.
Almost all of the work to date has focused on immobilising amines on as
synthesised silica supports or calicined mesoporous materials. Further improvements on
CO2 capacity are challenging because of the limitations on the loading density and amine
efficiency. To further advance this area novel supports have been developed such as
mesoporous capsules developed by Qi et al.[23] These materials featured a hollow
porous silica shell built around nanosized polystyrene latex beads which are then
removed by calcinations (see figure 3.5). The hollow mesoporous capsules have a large
particle size, high interior void volume and large pores to improve the accessibly of
adsorption sites and increase mass transfer. These capsules were then impregnated with
PEI/TEPA leading to a highly successful solid sorbent for CO2 with capacity of up to 7.93
mmolCO2 g-1. [20] These adsorbents had an optimum loading of 50-75 wt%, with higher
capacities leading to pore blockages. These capsules also demonstrate exceptional
stability, with a large retention of capacity observed over repeated adsorption and
regeneration cycles.

Figure 3.5: Process schematic for the formation of mesoporous capsule adsorbents.

The vast majority of amine sorbents use well known silica support as a basis for
the adsorbents. This is undoubtedly due to their ease of synthesis, high porosity,
tuneable pore volume and predictability. Recent contributions from Kuwahara et al. have
exploited a niche area by tuning the properties of the silica support by the introduction
of Zirconium heteroatom into the structure of the support.[24] They successfully
improved the adsorption characteristics of PEI-SBA-15 by incorporating Zr atoms into
the silicate framework creating more effective adsorption sites (see figure 3.6). Similar
modification has been conducted on silicas which interact with CO2. For example work by
Ratnasamy reported Ti-SBA-15 could be used to effectively catalyse the cycloaddition of
CO2 and epoxides.[25] Incorporating Zr into SBA-15-PEI led to a marked increase in
both CO2 adsorption capacity and short term stability as highlighted in figure 4. The
effect of Zr also led to an increase in cyclic stability as the Zr is thought to stabilise the
PEI preventing unwanted degradation/aggregation. These findings underline the
importance of the acid/base properties of the support on CO2 adsorption capacity.

41
Literature review August 2012

Work by Filburn et al. immobilised a series of alcohol amines with the pores of
high surface area PMMA beads.[26] These materials could be regenerated using a
pressure swing adsorption process. Further work by Filburn produced TEPA immobilised
on PMMA, a material which exhibited excellent regeneration and capacity; a dry CO 2
uptake of 10.21 mmolCO2 g-1 at 40°C rising to 14.03 mmolCO2 g-1 under humid conditions
[27] Amines for CO2 capture have also been immobilised upon a range of non silica
supports including but limited to polystyrene, silicon dioxide, activated carbon, solid
resins, zeolites and carbon nanotubes.[28-33] Although not discussed here these
materials have exhibited good CO2 adsorption capacities and are often at much lower
cost than expensive mesoporous supports

Figure 3.6: The enhancement of PEI uptake capacity with the incorporation of Zr into the
silica support. Also note an increase in stability over the 4 cycles.

3.1.6 Regeneration and sorbent stability

For any adsorbent to have a practical application it must be stable over a great
number of adsorption and desorption cycles. Physisorbed amines are regenerated in two
ways either temperature swing adsorption or pressure swing adsorption both acting to
shift the equilibrium in favour of gaseous CO2. Temperature swing adsorption (TSA)
increases the temperature of the adsorbent, providing the activation energy for
desorption. TSA helps shift the equilibrium to desorption as physical adsorption is an
exothermic process. Pressure swing adsorption passes an inert gas through the
adsorbent  shifting  the  equilibrium  towards  desorption  in  accordance  to  Le  Chatlier’s  
principle. Most reported regeneration methods utilise aspects of both increasing the
adsorbent temperature accompanied with an inert gas purge. Table 3 highlights a range
of regeneration conditions as well as the measured stability. From this table it is clear
that the majority of stability tests are a departure from realistic operating conditions.
Most studies utilise pure streams of CO2 , dry conditions or only test for a small number
of cycles.

42
Literature review August 2012

The data in table 3.3 highlights stability issues with immobilised adsorbents with
a large number losing capacity during cycling. This loss of capacity is due to volatilisation
of impregnated amine, amine degradation or changes in adsorbent morphology leading
to a reduction in available sites. Interestingly work by Song using molecular basket
adsorbents registered no change is sorbent stability although a fluctuation of adsorbent
capacity between each cycle was observed. It could be argued that with such a level of
fluctuation and short cycle time sorbent stability would be difficult to measure.

Amine Support Cycles Loss Operating Regeneration Reference


conditions conditions

TEPA (50wt%) SBA-15 7 5% Dry 5% CO2 373 K, He purge [12]


/N2

TEPA (50wt%) MCM-41 6 9% Dry 5% CO2 373 K, He purge [18]


/N2

TEPA (50wt%) SBA-15 4 50% Humid 403 K, Ar purge [6]


(1.26% H2O)
10% CO2 /Ar

PEI (50wt%) MCM-41 10 0% - 12.97% CO2 , 348 K, He purge [7]


oscillation 3.75% O2
between
results 70.16% N2
observed.
13.12 % H2O,
75°C

PEI (50wt%) KIT-6 3 0% 100% CO2 348 K, N2 purge [17]

PEI (83wt%) MC400 50 12 % 100% 373K, N2 purge [20]


CO2 ,348K

PEI (83wt%) MC400 50 5% 100% 348K, N2 purge [20]


CO2 ,348K

TEPA(83wt%) MC400 50 60% 100% 373K, N2 purge [20]


CO2 ,348K

TEPA (83wt%) MC400 50 2% 100% 348K, N2 purge [20]


CO2 ,348K

PEI (50wt%) SBA-15 10 0% - 100% 383K, N2 purge [34]


oscillation CO2 ,348K
between
results
observed.

PEI (50wt%) SBA-15 20 0% - 100% 383K, He purge [19]


oscillation CO2 ,348K
between
results
observed.

Table 3.3:Regeneration performance and conditions of class 1 adsorbents.

43
Literature review August 2012

3.1.7 Real flue gas conditions:

To date only a small handful of sorbents have been studied under either real or
stimulated flue gas conditions. Song studied the molecular basket adsorbent firstly under
stimulated flue gas conditions and then under real natural gas flue gas conditions. [15]
The sorbent performed under dry CO2/N2 gas streams however, the capture capacity was
found to be 50% lower under real flue gas conditions. The differences in stability could
be attributed to the presence of oxygen in the real flue gas feed. Studies under real flue
gas conditions including 60-70 ppm NOx , CO 200-300 ppm highlighted that the PEI-
MCM-41 sorbent adsorbed both CO2 and NOx at a ratio of 3000:1 however the NOx was
not desorbed during regeneration. It is believe that gradual poisoning of any sorbent
would be damaging to its long term stability, especially in coal flue gases where
concentrations of NOx are much greater. Recent work by Liu focused on the performance
of KIT-6 type mesoporous silica impregnated with TEPA under stimulated and humid flue
gas conditions. They reported that the presence of water increased the adsorption
capacity from 2.85mmolg-1 to 3.2 mmolCO2 g-1 at 343K, 37% relative humidity. The
adsorbent was gradually poisoned by SO2 at concentrations below 300 ppm but not
significantly changed by the presence of NO up to 400ppm. It was reported that the
adsorption capacity remained unchanged after 10 cycles exposed to 10% CO2, 100ppm
SO2 , 200ppm NO and 100% RH heated to 393K. This is a promising result as levels this
low are required under the large combustion plant directive.

44
Literature review August 2012

3.2 Class 2: Covalently bound amines

Amines can be attached to solid supports though a covalent bond typically formed
between a surface silanol and a silicon alkoxide. The covalent linker between the support
and amine allows these materials to be fully reversible as leeching due to volatile
physisorbed amines is none existent. These materials are constructed using well known
organic chemistry with a general scheme presented below in figure 3.7. Briefly, the
silanol groups on the surface of the silica react with the silicon alkyoxide releasing an
alcohol dependent on the type of silicon ester used. Shorter esters such as methyl silicon
esters tend to be more reactive than bulkier propyl esters on the basis of steric crowding
of the electrophilic silicon centre. [35] The level of surface of functionalisation is
dependent on the surface concentration of silanols on the support, reaction conditions,
choice of ester and pendant chain. The principal factor here is the availability and
accessibility of silanols on the surface of the support; this is related to the method of
silica preparation and normally falls between the range of one to five silanol groups nm -2.
Silanol density is often difficult to determine at and at times estimated from silane
coverage using 13C Nuclear magnetic spectroscopy (NMR) assuming a given rate of
silanol conversion. [36] Studies using 29Si cross polarization magic angle spinning NMR
were used characterise the 4 possible surface hydroxyl species given below in Figure 3.8.
[37] Free silanols are the most reactive, followed by vincinal silanols. Geminal silanols
and siloxane bridges are unreactive under standard grafting conditions. Fourier-
transform infrared spectroscopy was used to study surface hydroxyl species by Van Der
Voort who reported that the surface silanol density decreased with increasing
calcinations temperature. [38] They also reported an increase in silanol density arising
from ethanol extraction of the templates as oppose to calcinations. The morphology of
the support is highly important as this affects the number of accessible silanols sites.
Work by Chaffee discovered that the pore curvature of the silica support effected the
accessibility of the silanol groups with a smaller extent of curvature leading to a greater
silane loading.[39]

Figure 3.7: A general scheme for the production of covalently tethered amines.

Figure 3.8: The 4 classes of silanol found on the surface of mesoporous silicas.

45
Literature review August 2012

3.2.1 Types of amine

The concept of carbon dioxide adsorption on immobilised amine supports was first
reported by Lea et al., who prepared 3-aminoproply immobilised on silica gel [40]. As is
the case in class 1 adsorbents the amount of amine loading and amine efficiency are
critical variables in the performance of class 2 adsorbents. These are influenced by a
combination of the amine used and the surface of the support. As highlighted in section
X, primary and secondary amines capture CO2 more reversibly than tertiary amines and
are preferred amines for class 2 adsorbents. Yogo studied the difference between amine
order and amine density, concluding that the amine coverage followed the order
mono>di> triamine silane.[41]
The amine density as a function of support weight is another key consideration
and mono, di and tri aminosilanes have been extensively studied. Figure 3.9 highlights
the most commonly used amines is class 2 adsorbents. Monoaminesilanes such as the
frequently utilised 3-aminopropyltrimethoxysilane usually contain 1 primary amine head
group. Diaminesilanes have one primary amine and one secondary amine in their
structure with triaminesilanes featuring two secondary amines and a terminal primary
amine. Although a vast number of usable silanes are available many more can be
envisaged; researchers of solid adsorbents are limited to those commercially available
owing to the desired development of a low cost adsorbent.

Figure 3.9: Commonly used amines for the surface functionalisation of silica supports.

46
Literature review August 2012

The amine loading is normally reported as the number of amines per given mass
of the adsorbent and has a large and contributes to the effectiveness of the adsorbent.
However this number gives no information on the likely amine efficiency, the kinetics or
resultant equilibrium capacity of the adsorbent. From intuition one can assume that
adsorbents with more amine functionalities present will have a greater amine loading.
Evidence taken from the average amine loading confirms this theory although the results
suggests that other factors are influencing the amount of amine loading.[9]

Figure 3.10: Average amine loading normalised for the number of basic nitrogens in each
support.

If the amine loading normalised for the aminosilane stoichiometry is calculated


then the number of silanes bound follows the order of tri<di<monosilanes (figure 3.10).
This is attributed to the increased steric demands of the larger tri and disilanes,
preventing access to potential reactive sites. However, this reduction in surface
functionalisation is offset by the increased stoichiometric ratio of amines.
Owing to the large range of available amine moieties available there are
numerous amine supports that do not fit within the rigid mono, di and triamine
classification. Many amine sorbents may only contain one secondary or tertiary amine.
Zelenak et al. studied the effect of amine basicity on CO2 adsorption for a range of
amines on SBA-12.[42] It was reported that the higher the basicity of the amine the
higher the uptake capacities with the CO2 adsorption following the order of basicity
AP>MAP>PAP(see figure 3.9). In terms of basicity AP and MAP have very similar pkab
values as although the electron donating effect of the methyl group should lead to a
higher basicity the steric demand of the methyl group reduces the observed basicity. In
the instance of PAP the basic lone pair of the nitrogen is delocalised within the aromatic
π system resulting in PAP having a pkab of 8.9 compared to a pkab of ~4 for AP and MAP.
The amine efficiency also echoed this trend in basicity with SBA-12-AP having an
efficiency of 0.49 compared to a theoretical maximum of 0.5 and SBA-12-MAP having a
slightly lower efficiency of 0.45. However, SBA-12-PAP had a lower efficiency of 0.43 and
a capacity ~60% of that of the AP functionalised silica. Assuming the CO2 adsorption
process  follows  the  Dankwert’s  mechanism  the  delocalisation of the N lone pair in PAP
not only reduces the basicity but also stabilises the singly charged intermediate leading
to increased energy barrier towards zwitterions formation; ultimately this leads to a
lower equilibrium adsorption capacity.[43] Given the general similarities in pkab and
hence capture ability, the principle factor effecting capture capacity are the steric
requirements of the silane. Larger silanes such as PAP lead to a reduction in surface
functionalisation which inhibits the uptake capacity of the resultant materials.

47
Literature review August 2012

Alternative strategies towards the immobilisation of amines on silica surfaces


include a two step synthetic strategy; functionalisation with alkyl halide groups then
subsequent substitution reaction with the required amine species. This method allows
the use of more elaborate amines and is often used to produce resulting silanes not
available commercially. Multistep approaches should be approached with caution as each
added step adds complexity and hence cost, issues to be avoided during the
development of low cost scalable adsorbents. This method has been successfully applied
by Tsuda on silica gels and by Guliants who successfully functionalised MCM-48 using the
two step procedure shown in figure 3.11 to functionalise the sterically demanding
pyrrolidine group.[44, 45]

Figure 3.11: The two step immobilisation method used by Guilants.

3.2.2 Kinetics

Class 2 solid amine adsorbents have shown rapid kinetics for the uptake of CO 2
exhibiting very short breakthrough times. However, no detailed studied into the kinetics
of class 2 adsorbents have been completed and little comparable data exists. 3-
aminopropyltriethoxysilane (MAP) on SBA-15 was shown to have an almost
instantaneous breakthrough time, although the material had a low capacity of
0.52mmolCO2 g-1.[41] Using IR spectroscopy and online mass spectroscopy, Khatri
probed the mechanism and kinetics of [N-(2-aminoethyl)-3-aminopropyl]
trimethoxysilane (see figure 3.9) on SBA-15.[46] They also reported rapid breakthrough
curves. Later work by Yogo using IR studies of mono, di and tiramine silanes showed
that the kinetics of adsorption were directly proportional to the amine loading.[47]
Working on SBA-15 they reported that calcination before grafting led to materials with
the highest silanol surface area, fastest kinetics but a lower capacity compared to over
materials. It is clear that an interplay exists between amine loading, capture capacity
and kinetics.
Chaffee investigated the observed similar rapid kinetics on mono silane
functionalised hexagonal mesoporous silica (HMS) reaching saturation within 2 minutes
under a dry 10% CO2 atmosphere. This trend continues with other supports including
MCM-48 and silica xerogel with near equilibrium capacities reached within 30mins for a
dry 5% CO2 gas stream. [48] Sayari found exceeding fast adsorption kinetics for
triamine grafted MCM-48 with a maximum adsorption rate of 1.79mmolCO2 g-1min-1. This
material also featured a large amine loading of 6.0mmolNg-1 and was attributed to the
incorporation of water into the grafting stage.[49] Although several breakthrough curves
have been reported in this area the lack of consistency between measurement
techniques and parameters eludes few conclusions other than empirical observations;
the kinetics are loosely related to the uptake capacity of the amine. In comparison to
class 1 amine adsorbents, class 2 materials general exhibit faster breakthrough curves
but lower levels of equilibrium capacity. One assumes that class 1 adsorbents have
saturated pores with mass transfer being a limiting factor, in the instance of class 2
adsorbents some pore void volume and surface area is retained allowing for rapid mass

48
Literature review August 2012

transit. In this instance the kinetics are controlled by the rapid carbamate formation
between free amines and CO2.

3.2.3 Stability

In principal the covalent bond linking together the amine group and the silica
support should result in an increased level of stability for class 2 amines adsorbents.
Several studies have reported the thermal stability of covalently tethered amines under
inert atmospheres reporting degradation temperatures between 433-623 K. [50-53]
Coal fired power plant flue gas typically contains between 3-10 % oxygen which could
contribute towards the degradation of solid adsorbents akin to that of liquid MEA
solutions. Recent work by Jones et al. explored the oxidative stability of various
covalently tethered amines on silica mesocellular foam.[54] Jones subjected the
adsorbents to the oxygen equivalent of 288 TSA cycles finding that primary and tertiary
amines had superior stability towards oxidative degradation than secondary amines. The
authors observed intermolecular cooperatively within the degradation mechanism. In
aminosilanes with more than one amine both amines degraded at the same time via a
cooperative degradation pathway.
Drage et al. reported that at above temperatures of 135°C amine groups within
PEI reacted with CO2 to form stable urea species.[11] These species were further studied
by Sayari who investigated the stability of primary, secondary and tertiary MAP bound to
a PE-MCM-41 support.[55] Under a dry CO2 atmosphere primary monoamine had the
highest capacity but was the least stable. The primary amines formed urea through
either an isocyanate complex or via carbamate dehydration (see figure 3.12). The
secondary amines were found to be most stable as they were unable to form isocyanate
intermediate and did not undergo carbamate dehydration.

Figure 3.12: Degradation pathways observed for monoaminesilanes by Sayari et al.

The work of Jones and Sayari complements as it is known that urea formation is
inhibited by water and can be reversed by hydrolysis.[56] Differences in behaviour can
be attributed to the differing temperature regimes as urea formation only becomes
significant above 130°C and differing gas compositions; high concentrations of oxygen
could lead to alternative degradation pathways. In a further contribution Sayari et al.
studied a range of mono, di and tri aminosilanes and concluded that under dry CO2 only
secondary monoaminosilanes were stable. [57] All other species studied were reported
to undergo deactivation via urea formation. Building upon this study Sayari et al.
exposed the same adsorbents to unattenuated air at elevated temperature, the
surprising results are shown below in figure 3.13. This findings highlight an apparent
stability of primary amines under oxygenated environments however the original CO2
uptake capacity was poor.[58] Studies under real flue gas conditions may elude clues
regarding the stability of solid adsorbents under real flue gas conditions. Unfortunately,
to date no work has been published on the stability and degradation pathway of class 2
adsorbents under moist flue gas or real flue gas conditions.

49
Literature review August 2012

As described previously real flue gas is comprised of more than CO 2 with O2, SOx
and NOx reducing the effectiveness of liquid amine systems.[59] With this knowledge in
mind it is not until recently that the effect of these minor species has been investigated.
Zheng observed irreversible oxidation of diamines of SBA-15, the rate of which was
dependent on the temperature of the reaction. Furthermore Chuang found that the
adsorbents were irreversibly deactivated by SO2 species.

Figure 3.13: The relative stability of primary amines towards degradation in air.

3.2.4 Sorbent regeneration

The CO2 capture capacity is considered to be fully regenerable for class 2


adsorbents, however the conditions for regeneration and sorbent support influence the
level of regenerability. Regeneration by PSA tends not to fully regenerate the materials
with heating about 373K usually required in accompaniment to the purge gas. [60-62]
Utilising pressure swing adsorption with high temperature is uneconomical hence a
temperature swing regeneration process is preferred. Chaffee et al. reported that
aminoproplysilation of HMS was fully regenerable with an inert purge at 298K under
anhydrous conditions. [39]In the presence of moisture this regenerability was not
observed, it is believed that the water led to the conversion of carbamates of single and
bicarbonates which were not removed with a low temperature purge. As part of their
study on the effects of differing amines Yogo et al. concluded that mono, di and
triaminesilanes on SBA-15 were able to be fully regenerated using a combination of inert
gas purge and heating above 373K.[63] Zheng studied the regeneration of diamine
SBA-15 adsorbents used in humid gas streams and found them to be fully regenerable
with a similar inert purge and addition of heat.[36]

50
Literature review August 2012

Figure 3.14: Fully regenerable behaviour exhibited by MAP on HMS. Regeneration was by
an inert gas purge alone.

Work has shown that the desorption temperature is dependent on the differences
in heat of adsorption of mono, di and triaminosilanes [64] The authors reported that
triaminesilanes had a larger regeneration temperature due to the formation of
intramolecular bidentate species. Related work by Chuang reported the presence of a
range of amine derived species on the surface of the support each having a differing
stabling towards regeneration. [61] During several adsorption/desorption cycles an
increase in the heat of adsorption was noted accompanied with an increase in the
formation of bidentate carbonate species. During these experiments the sorbents were
regenerated with a mixture of He and H2O, the water is believed to displace the CO2. It
was proposed that the elevated concentrations of water led to an increase in carbonate
formation during adsorption.
Steam regeneration was recently investigated by Jones et al. a cost effective
means of providing a simultaneous pressure and temperature swing. Steam regeneration
appears more favourable compared to pressure desorption with an inert gas as the
separation of CO2 and water is facile compared to CO2 /inert gas mixtures [65] However
the amines were found to adsorb substantial amounts of water leading to a loss 17% of
the APS capacity after 3 cycles. More recent work by Veneman focused on the
continuous CO2 capture using a lab scale fluidised be reactor using readily available
adsorbents. [66] They concluded that TSA was more suitable compared to steam
regeneration or pressure swing adsorption. Steam or PSA would lead to a reduced
working capacity as the materials are still adsorbent at CO 2 <0.1bar requiring a
prohibitively large volume of steam or pressure reduction to desorb the CO 2.

3.2.5 Capacity:

The lower amine loading of class 2 adsorbents gives them a lower equilibrium
capacity with the average capacity around 1 mmolCO2 g-1 (See the comprehensive
review by Jones and Gupta for a table of capacities).[9] [67]. This is a result of the
lower amount of amine loading relative to other solid supported amines. Covalently
bound amines form a monolayer of aminosilane species on the pore surface as oppose to
multilayer formation in the case of impregnated amines. The reported capacities for
covalently tethered amines are variable depending on the silanol density of the support
and the size of the functionalising species. The pore size and architecture influence the
loading as well as mass transit properties and the number of accessible amines.

51
Literature review August 2012

Generally, class 2 adsorbents exhibit higher amine efficiencies than impregnated


amines, a result of increased amine availability. To date the largest capacity observed
for class 2 adsorbents in PE-MCM-41 functionalised with 3-[2-(2-
aminoethylamino)ethylamino]propyltrimethoxysilane with a CO2 uptake of 2.65mmolCO2
g-1 under a 0.05bar partial pressure of CO2 .[49] This material had a surprisingly high
amine loading approximately 2.5 times that of the average class 2 adsorbent. This
suggests that if the amine loading can be increased on other covalently bound amines
exceptional capacities could be observed.

3.2.6 Role of Water:

Similarly to impregnated amines, covalently tethered amines exhibit an increase


in CO2 uptake capacity in the presence of water.[68] This increase in capacity is due to
the formation of mono and bicarbonates leading to an increase in amine efficiency. [49,
69] It has been suggested that for covalently tethered amines the rate of carbamate
formation is much faster than the rate of carbonate formation; as is the case with liquid
amine systems.[70] Chaffee observed two stage kinetics for CO2 adsorption on amine
functionalised silica; rapid first stage kinetics followed by sluggish second stage kinetics.
[39] The first stage was similar to that observed under dry conditions and was attributed
to the rapid formation of carbonates with the second slower rate assigned as the
formation of carbonates and bicarbonsates. This helps rationalise the observations of
Yogo who using IR only observed the formation of carbamates within the 5 minute
observation timeframe.[47] With this in mind it could be more efficient to only expose
the flue gas to the adsorbent during this rapid uptake regime and then regenerate before
formation of carbonates which have been shown to exhibit higher stability towards
regeneration.

3.2.7Support effects

As highlighted previously the support material has a heavy influence on the


properties of the sorbent, with many variables applicable to both class 1 and class 2
amines. The number of active silanol groups is a key consideration as in the pore size
and internal volume, without these critical factors covalent functionalisation is difficult to
achieve. Zhao suggests that there are two contributions to carbon dioxide adsorption;
the chemical adsorption based on the active sites and the capillary condensation caused
by nanoscale channels of the mesoporous silica.[71] This argument was supported by
Zelenak et al., who found a positive correlation between pore size and CO2 uptake and
kinetics for a range aminopropyl functionalised silicas .[62] A larger pore offers two
benefits; allowing the grafting of larger molecules and allowing facile mass transport of
CO2 to the functionalised surface. Supports with larger pores and high surface density of
amines have larger capacities. The lower limit for efficient amine adsorbents is 35A
below this the pores are inaccessible to CO2 molecules owing to limited diffusion.

52
Literature review August 2012

3.3 Class 3 adsorbents: covalently bound amine polymers

Recent developments in the area of solid supported amines have led to the
creation of a new class of adsorbent. Class 3 adsorbents combined the favourable
attributes if class 1 and 2 materials by conducting in situ polymerisation constrained
within pores of silica supports. These allow the formation of covalently attached low
molecular weight polymers within support pores. These materials were originally
developed by Kim and Linden who formed polymerisation of aziridine flat silica and
within large pore mesoporous silica respectively but did not report an application.[72, 73]
These materials were found to have good surface areas with the majority of the polymer
formed within the pores allowing the formation of non-sticky particles; a key feature for
the development of practical adsorbents.
Work by Jones et al. who conducted the surface polymerisation of aziridine on
SBA-15 yielding an immobilised silica polymer on a high surface area support, displaying
a substantial CO2 capture capacity.[6] The materials had a high amine loading of
7.0mmolNg-1 and a good capacity of 3.1mmolCO2 g-1 at 25°C. With the amine covalently
bound the material exhibited good thermal stability between 25-130°C and was stable
towards regeneration over 10 cycles. However these materials exhibited retarded
kinetics due to pore blockages a barrier common to both type 1 and type 3 adsorbents.
A more detailed study examining the effect of the amine silica stoichiometry
found that this simple variable could be pivotal in preparing materials with differing
properties.[74] A large amine to silica ration let to a larger degree of polymerisation and
amine loadings of up to 10mmol N g-1 and good working capacities from 2 -5.6mmolCO2
g-1 at 25°C. With an increasing amine loading pore inhibition was present believed to
lead to a reduction in amine efficiencies and kinetics. It was proposed that the
dimensions of the SBA-15 support were the physical boundaries limiting the number of
amines within the adsorbent. Difficulties were reported in controlling the polymerisation
as the degree of branching was unable to be tailored towards biasing primary, secondary
or tertiary amines but was found to be repeatable.
Building upon this work the effect of the support structure on the same
hyperbranched aminosilica (HAS) was exploring using a range of supports with larger
pore sizes.[75] This pore expanded hybranched aminosilicas (PEHAS) did not exhibit
pore blocking but led to lower amine loading that the smaller pore supported analogue.
This reduction in amine loading led to reduction in observed capacity although the
uptake kinetics showed no evidence of hindered diffusion. These materials were testing
under humid 10% CO2 with Ar and it was found that the adsorption capacity correlated
with the amine loading. However smaller pore size supports had a larger uptake see
Figure 3.15. This could be explained by the fact that pore size is inversely proportional to
the surface affinity and can significantly affect the adsorption capacity. [76] As the larger
pore supports give a lower degree of polymerisation, the previous hypothesis suggested
by Jones that termination is a result of steric constraints does not satisfy this
observation. Class 3 materials have large potential for use in carbon capture but are in
their infancy. Methods to improve the amine loading and hence capacity would produce
materials with very favourable properties.

53
Literature review August 2012

Figure 3.15: CO2 uptake capacity for HAS on a range of mesoporous silica supports. Note
the increase in CO2 uptake for the smaller pore SBA-15 support for the same give amine
loading.

Chaffee conducted the stepwise polymerisation of melamine dendrons exhibiting


an excellent level of control not present in aziridine polymerisation. They found that
successive addition level to an increase in amine loading and hence capacity. However a
discrepancy between the theoretical and actual N/H,N/C ratios suggesting incomplete
growth or cross linking between chains. After the fourth stepwise reaction the material
pores are completely filled with an organic loading of 50.3wt %. However the CO2 uptake
capacity was reduced at such high amine loading due to mass transfer limitations arising
from pore blockages rendering many sites inaccessible.
Recently biological derived materials have been bound to porous silica supports
and their use in CO2 capture. Poly (L-lysine) was polymerised on the surface of SBA-15-
APTMS as shown in figure 3.16. [77]This led to the formation of materials with fair
amine loadings and good BET surface areas between 240-360 m2g-1 post
functionalisation. In these materials the CO2 uptake was found to be independent of
amine loading with low capacities of around 1.15mmolCO2 g-1 and stable during 3
temperature swing cycles. These novel materials have yet to be optimised but could be
developed into viable capture materials.

Figure 3.16: Schematic for the production of partially biologically derived solid
adsorbents

54
Literature review August 2012

3.4 Potential for solid supported amine adsorbents for use in carbon capture
systems

The adsorption performance of solid support amine adsorbents have been shown
to be dependent on the composition and morphology of both the amine and the support.
The adsorption behaviour is either aided or unaffected by the presence of water with the
largest capacities recorded under humid conditions. Supported amines are well suited to
the flue gas temperature and pressure required for post combustion applications,
exhibiting high selectivity for CO2 at low partial pressures.
Impregnated amines are simple and cheap to produce with high capacities due to
high levels of amine loading. This high loading results in poor kinetics and low amine
efficiencies. Furthermore as these materials are not covalently bound, leeching of the
volatile amine is a problem towards stability.
Covalently tethered amines have been shown to be fully regenerable with
temperature swing adsorption. Owing to the lower levels of amine loading they exhibit
fast initial kinetics but relatively low adsorption capacities. These materials typically have
high amine efficiencies making significant improvements in capacities challenging. The
additional cost and availability of silane reagents at such scale also poses a potential
problem in the upscaling of these materials
Class 3 adsorbents are still in their infancy and to date appear promising although
little work has been completed in this area. The high capacities and covalent attachment
are notable attributes but understanding how to control polymerisation combined with
cycling and real flue gas tests are required.

55
Literature review August 2012

3.5 References

1. Li, W., et al., Steam-Stripping for Regeneration of Supported Amine-Based CO2


Adsorbents. ChemSusChem. 3(8): p. 899-903.
2. Xu, X., et al., Novel Polyethylenimine-Modified Mesoporous Molecular Sieve of
MCM-41 Type as High-Capacity Adsorbent for CO2 Capture. Energy & Fuels, 2002.
16(6): p. 1463-1469.
3. Xu, X., et al., Preparation and characterization of novel CO2 molecular basket
adsorbents based on polymer-modified mesoporous molecular sieve MCM-41.
Microporous and Mesoporous Materials, 2003. 62(2): p. 29-45.
4. Satyapal, S., et al., Performance and Properties of a Solid Amine Sorbent for
Carbon Dioxide Removal in Space Life Support Applications. Energy & Fuels, 2001.
15(2): p. 250-255.
5. A.L. Chaffee, G.P.K., W.S. Delaney, Preprints of Symposia - American Society,
Division of Fuel Chemistry. 2002. 47(1): p. 65-66.
6. Hicks, J.C., et al., Designing Adsorbents for CO2 Capture from Flue Gas-
Hyperbranched Aminosilicas Capable of Capturing CO2 Reversibly. Journal of the
American Chemical Society, 2008. 130(10): p. 2902-2903.
7. Xu, X., et al., Influence of Moisture on CO2 Separation from Gas Mixture by a
Nanoporous Adsorbent Based on Polyethylenimine-Modified Molecular Sieve MCM-
41. Industrial & Engineering Chemistry Research, 2005. 44(21): p. 8113-8119.
8. Yue, M.B., et al., Promoting the CO2 adsorption in the amine-containing SBA-15
by hydroxyl group. Microporous and Mesoporous Materials, 2008. 114(3): p. 74-
81.
9. Choi, S., J.H. Drese, and C.W. Jones, Adsorbent Materials for Carbon Dioxide
Capture from Large Anthropogenic Point Sources. ChemSusChem, 2009. 2(9): p.
796-854.
10. von Harpe, A., et al., Characterization of commercially available and synthesized
polyethylenimines for gene delivery. Journal of Controlled Release, 2000. 69(2):
p. 309-322.
11. Drage, T.C., et al., Thermal stability of polyethylenimine based carbon dioxide
adsorbents and its influence on selection of regeneration strategies. Microporous
and Mesoporous Materials, 2008. 116(3): p. 504-512.
12. Yue, M.B., et al., CO2 Capture by As-Prepared SBA-15 with an Occluded Organic
Template. Advanced Functional Materials, 2006. 16(13): p. 1717-1722.
13. Franchi, R.S., P.J.E. Harlick, and A. Sayari, Applications of Pore-Expanded
Mesoporous Silica. 2. Development of a High-Capacity, Water-Tolerant Adsorbent
for CO2. Industrial & Engineering Chemistry Research, 2005. 44(21): p. 8007-
8013.
14. Xiang, Z., S. Leng, and D. Cao, Functional Group Modification of Metal Organic
Frameworks for CO2 Capture. The Journal of Physical Chemistry C. 116(19): p.
10573-10579.
15. Xu, X., et al., Adsorption separation of carbon dioxide from flue gas of natural
gas-fired boiler by a novel nanoporous molecular basket adsorbent. Fuel
Processing Technology, 2005. 86(15): p. 1457-1472.
16. G.F. Schutte, E.G.L., J. B. Corss and E. Esen, in Fifth Annual Conference on
Carbon Capture and Sequestration. 2006: Alexandria, Virginia.
17. Son, W.-J., J.-S. Choi, and W.-S. Ahn, Adsorptive removal of carbon dioxide using
polyethyleneimine-loaded mesoporous silica materials. Microporous and
Mesoporous Materials, 2008. 113(3): p. 31-40.
18. Yue, M.B., et al., Efficient CO2 Capturer Derived from As-Synthesized MCM-41
Modified with Amine. Chemistry – A European Journal, 2008. 14(11): p. 3442-
3451.
19. Ma, X., X. Wang, and C. Song, Molecular Basket Sorbents for Separation of CO2
and H2S from Various Gas Streams. Journal of the American Chemical Society,
2009. 131(16): p. 5777-5783.

56
Literature review August 2012

20. Qi, G., et al., High efficiency nanocomposite sorbents for CO2 capture based on
amine-functionalized mesoporous capsules. Energy & Environmental Science.
4(2): p. 444-452.
21. Zeng, P., et al., Nanoparticle sintering simulations. Materials Science and
Engineering: A, 1998. 252(2): p. 301-306.
22. Sayari, A., Unprecedented Expansion of the Pore Size and Volume of Periodic
Mesoporous Silica. Angewandte Chemie International Edition, 2000. 39(16): p.
2920-2922.
23. Qi, G., et al., Facile and Scalable Synthesis of Monodispersed Spherical Capsules
with a Mesoporous Shell. Chemistry of Materials. 22(9): p. 2693-2695.
24. Kuwahara, Y., et al., Dramatic Enhancement of CO2 Uptake by
Poly(ethyleneimine) Using Zirconosilicate Supports. Journal of the American
Chemical Society. 134(26): p. 10757-10760.
25. Srivastava, R., D. Srinivas, and P. Ratnasamy, CO2 activation and synthesis of
cyclic carbonates and alkyl/aryl carbamates over adenine-modified Ti-SBA-15
solid catalysts. Journal of Catalysis, 2005. 233(1): p. 1-15.
26. Filburn, T., J.J. Helble, and R.A. Weiss, Development of Supported Ethanolamines
and Modified Ethanolamines for CO2 Capture. Industrial & Engineering Chemistry
Research, 2005. 44(5): p. 1542-1546.
27. Lee, S., et al., Screening Test of Solid Amine Sorbents for CO2 Capture. Industrial
& Engineering Chemistry Research, 2008. 47(19): p. 7419-7423.
28. Gray, M.L., et al., Parametric Study of Solid Amine Sorbents for the Capture of
Carbon Dioxide. Energy & Fuels, 2009. 23: p. 4840-4844.
29. Przepiórski, J., M. Skrodzewicz, and A.W. Morawski, High temperature ammonia
treatment of activated carbon for enhancement of CO2 adsorption. Applied
Surface Science, 2004. 225(4): p. 235-242.
30. Arenillas, A., et al., CO2 capture using some fly ash-derived carbon materials.
Fuel, 2005. 84(17): p. 2204-2210.
31. Drage, T.C., et al., Preparation of carbon dioxide adsorbents from the chemical
activation of urea formaldehyde and melamine formaldehyde resins. Fuel, 2007.
86(2): p. 22-31.
32. Jadhav, P.D., et al., Monoethanol Amine Modified Zeolite 13X for CO2 Adsorption
at Different Temperatures. Energy & Fuels, 2007. 21(6): p. 3555-3559.
33. Fifield L. S., F.G.E., Addleman R. S., Aardahl C. L., Carbon dioxide capture using
amine-based molecualr anchors on multi wall carbon nanotubes. Div. Fuel Chem.,
Am. Chem. Soc. - Prepr. Symp., 2004. 49: p. 296-297.
34. Wang, X., et al., A solid molecular basket sorbent for CO2 capture from gas
streams with low CO2 concentration under ambient conditions. Physical
Chemistry Chemical Physics. 14(4): p. 1485-1492.
35. Jonathan Clayden, N.G., Sturat Warren, Peter Wothers, Organic Chemistry. 1st
Edition ed. 2001: Oxford University Press.
36. Zheng, F., et al., Ethylenediamine-Modified SBA-15 as Regenerable CO2 Sorbent.
Industrial & Engineering Chemistry Research, 2005. 44(9): p. 3099-3105.
37. Sindorf, D.W. and G.E. Maciel, Silicon-29 NMR study of dehydrated/rehydrated
silica gel using cross polarization and magic-angle spinning. Journal of the
American Chemical Society, 1983. 105(6): p. 1487-1493.
38. Van Der Voort, P., et al., Effect of porosity on the distribution and reactivity of
hydroxyl groups on the surface of silica gel. Journal of the Chemical Society,
Faraday Transactions, 1991. 87(24): p. 3899-3905.
39. Knowles, G.P., et al., Aminopropyl-functionalized mesoporous silicas as CO2
adsorbents. Fuel Processing Technology, 2005. 86(15): p. 1435-1448.
40. Leal, O., et al., Reversible adsorption of carbon dioxide on amine surface-bonded
silica gel. Inorganica Chimica Acta, 1995. 240(1-2): p. 183-189.
41. Hiyoshi, N., K. Yogo, and T. Yashima, Adsorption of carbon dioxide on amine
modified SBA-15 in the presence of water vapor. Chemistry Letters, 2004. 33(5):
p. 510-511.

57
Literature review August 2012

42. Zelenak, V., et al., Amine-modified SBA-12 mesoporous silica for carbon dioxide
capture: Effect of amine basicity on sorption properties. Microporous and
Mesoporous Materials, 2008. 116(1-3): p. 358-364.
43. Danckwerts, P.V., The reaction of CO2 with ethanolamines. Chemical Engineering
Science, 1979. 34(4): p. 443-446.
44. Tsuda, T. and T. Fujiwara, Polyethyleneimine and macrocyclic polyamine silica
gels acting as carbon dioxide absorbents. Journal of the Chemical Society,
Chemical Communications, 1992(22): p. 1659-1661.
45. Kim, S., et al., Tailoring Pore Properties of MCM-48 Silica for Selective Adsorption
of CO2. The Journal of Physical Chemistry B, 2005. 109(13): p. 6287-6293.
46. Khatri, R.A., et al., Carbon dioxide capture by diamine-grafted SBA-15: A
combined Fourier transform infrared and mass spectrometry study. Industrial &
Engineering Chemistry Research, 2005. 44(10): p. 3702-3708.
47. Hiyoshi, N., K. Yogo, and T. Yashima, Adsorption characteristics of carbon dioxide
on organically functionalized SBA-15. Microporous and Mesoporous Materials,
2005. 84(1-3): p. 357-365.
48. Huang, H.Y., et al., Amine-grafted MCM-48 and silica xerogel as superior sorbents
for acidic gas removal from natural gas. Industrial & Engineering Chemistry
Research, 2003. 42(12): p. 2427-2433.
49. Harlick, P.J.E. and A. Sayari, Applications of pore-expanded mesoporous silica. 5.
Triamine grafted material with exceptional CO(2) dynamic and equilibrium
adsorption performance. Industrial & Engineering Chemistry Research, 2007.
46(2): p. 446-458.
50. Knowles, G.P., S.W. Delaney, and A.L. Chaffee, Diethylenetriamine[propyl(silyl)]-
Functionalized (DT) Mesoporous Silicas as CO2 Adsorbents. Industrial &
Engineering Chemistry Research, 2006. 45(8): p. 2626-2633.
51. Wei, J., et al., Adsorption of carbon dioxide on organically functionalized SBA-16.
Microporous and Mesoporous Materials, 2008. 116(3): p. 394-399.
52. Harlick, P.J.E. and A. Sayari, Applications of pore-expanded mesoporous silicas. 3.
Triamine silane grafting for enhanced CO(2) adsorption. Industrial & Engineering
Chemistry Research, 2006. 45(9): p. 3248-3255.
53. Liang, Z., et al., Stepwise growth of melamine-based dendrimers into mesopores
and their CO2 adsorption properties. Microporous and Mesoporous Materials,
2008. 111(3): p. 536-543.
54. Bollini, P., et al., Oxidative Degradation of Aminosilica Adsorbents Relevant to
Postcombustion CO2 Capture. Energy & Fuels. 25(5): p. 2416-2425.
55. Sayari, A., Y. Belmabkhout, and E. Da'na, CO2 Deactivation of Supported Amines:
Does the Nature of Amine Matter? Langmuir. 28(9): p. 4241-4247.
56. Sayari, A. and Y. Belmabkhout, Stabilization of Amine-Containing CO2 Adsorbents:
Dramatic Effect of Water Vapor. Journal of the American Chemical Society.
132(18): p. 6312-+.
57. Sayari, A., A. Heydari-Gorji, and Y. Yang, CO2-Induced Degradation of Amine-
Containing Adsorbents: Reaction Products and Pathways. J. Am. Chem. Soc.: p.
Ahead of Print.
58. Heydari-Gorji, A., Y. Belmabkhout, and A. Sayari, Degradation of amine-
supported CO2 adsorbents in the presence of oxygen-containing gases.
Microporous and Mesoporous Materials. 145(1-3): p. 146-149.
59. Tzimas, E., et al., Trade-off in emissions of acid gas pollutants and of carbon
dioxide in fossil fuel power plants with carbon capture. Energy Policy, 2007.
35(8): p. 3991-3998.
60. Gray, M.L., et al., Improved immobilized carbon dioxide capture sorbents. Fuel
Processing Technology, 2005. 86(14-15): p. 1449-1455.
61. Chang, A.C.C., et al., In-Situ Infrared Study of CO2 Adsorption on SBA-15
Grafted with 3-(Aminopropyl)triethoxysilane. Energy & Fuels, 2003. 17(2): p.
468-473.

58
Literature review August 2012

62. Zelenak, V., et al., Amine-modified ordered mesoporous silica: Effect of pore size
on carbon dioxide capture. Chemical Engineering Journal, 2008. 144(2): p. 336-
342.
63. Hiyoshi, N., K. Yogo, and T. Yashima, Adsorption of Carbon Dioxide on Amine-
modified MSU-H Silica in the Presence of Water Vapor. Chemistry Letters, 2008.
37(12): p. 1266-1267.
64. N. Hiyoshi, K.Y., T. Yashima, J. Jpn. Oet. Inst., 2005. 48.
65. Li, W.C., S.; Drese, J. H. ; Hornbostel, M.; Krishnan, G.; Eisenberger, P. M.;
Jones , C. W. , Steam-stripping for regeneration of supported amine-based CO2
adsorbents ChemSusChem, 2010. 3: p. 889-903.
66. Veneman, R., et al., Continuous CO2 capture in a circulating fluidized bed using
supported amine sorbents. Chemical Engineering Journal, Advance online
article.
67. Samanta, A., et al., Post-Combustion CO2 Capture Using Solid Sorbents: A
Review. Industrial & Engineering Chemistry Research. 51(4): p. 1438-1463.
68. Huang, H.Y., et al., Amine-Grafted MCM-48 and Silica Xerogel as Superior
Sorbents for Acidic Gas Removal from Natural Gas. Industrial & Engineering
Chemistry Research, 2002. 42(12): p. 2427-2433.
69. Serna-Guerrero, R., E. Da'na, and A. Sayari, New Insights into the Interactions of
CO(2) with Amine-Functionalized Silica. Industrial & Engineering Chemistry
Research, 2008. 47(23): p. 9406-9412.
70. Vaidya, P.D. and E.Y. Kenig, CO2-Alkanolamine Reaction Kinetics: A Review of
Recent Studies. Chemical Engineering & Technology, 2007. 30(11): p. 1467-1474.
71. Zhao, H., et al., CO2 Capture by the Amine-modified Mesoporous Materials. Acta
Physico-Chimica Sinica, 2007. 23(6): p. 801-806.
72. Kim, H.J., J.H. Moon, and J.W. Park, A hyperbranched poly(ethyleneimine) grown
on surfaces. Journal of Colloid and Interface Science, 2000. 227(1): p. 247-249.
73. Rosenholm, J.M., A. Penninkangas, and M. Linden, Amino-functionalization of
large-pore mesoscopically ordered silica by a one-step hyperbranching
polymerization of a surface-grown polyethyleneimine. Chemical Communications,
2006(37): p. 3909-3911.
74. Drese, J.H., et al., Synthesis-Structure-Property Relationships for Hyperbranched
Aminosilica CO2 Adsorbents. Advanced Functional Materials, 2009. 19(23): p.
3821-3832.
75. Drese, J.H., et al., Effect of support structure on CO2 adsorption properties of
pore-expanded hyperbranched aminosilicas. Microporous Mesoporous Mater. 151:
p. 231-240.
76. Dubinin, M.M., K.M. Nikolaev, and N.S. Polyakov, Adsorption of gases by
microporous adsorbents. Communication 3. Nitrogen/krypton/xenon-microporous
adsorbent carbon systems. Russian Chemical Bulletin, 1985. 34(2): p. 238-242.
77. Chaikittisilp, W., et al., Poly(L-lysine) brush-mesoporous silica hybrid material as
a biomolecule-based adsorbent for CO2 capture from simulated flue gas and air.
Chem.--Eur. J. 17(38): p. 10556-10561, S10556/1-S10556/4.

59
Literature review August 2012

4.0 Conclusion

The implementation of carbon capture and storage on large point sources of CO 2 requires
advances in both the operation and economics of process. A range of approaches have
been explored for this purpose with solid adsorbents appearing to be a promising way
forward.
In the course of this review the available solid adsorbents were compared against
established benchmark requirements for new materials and promising classes of
adsorbents discussed in depth. On the basis of outgoing flue gas temperatures and
pressures supported amines and metal organic frameworks were the selected classes of
materials to be studied in-depth. In their short research lifetime MOFs related to CO2
capture have been investigated as low temperature, high pressure physisorbents.
However recent work has highlight the potential for MOFs to be used for moderate
temperature, low pressure chemisorbents. The limitations for MOFs are their apparent
lack of stability especially with regard to water and low capacities at proposed operating
temperatures. Before their application in real systems can be considered studies under
real flue gas conditions need to be conducted and the effect on flue gas species on
capacity, stability and kinetics understood.
In contrast to this a great deal of research has been conducted on all classes of
solid supported amines. A large number of solid supported amines have been
investigated for post combustion carbon capture owing to the reliable amine chemistry
and favourable behaviour in humid environments. Class 1 amines exhibit good CO2
capture capacities owing to high levels of amine loading, although amine utilisation can
be low due to a large number in inaccessible sites. High levels of loading are also found
to inhibit the kinetics of adsorption and lead to formation of agglomerates. A principal
drawback of physisorbed amines is the leeching of volatile amine under humid conditions,
limiting the stability. Upon the rare test of class 1 amines under real flue gas conditions
the adsorption capacity was negatively influenced by flue gas species notably the
irreversible adsorption of NOx.
Covalently tethered amines are considered fully regenerable but typically have
lower amine loadings and hence low CO2 uptake capacities. The amount of loading is
dependent on the steric demands of the attached group but the resulting amine
efficiency is usually high. These materials exhibit fast kinetics but require extended
periods of time to reach equilibrium capacity. There have been little work regarding the
testing of these materials under extended cycles or under real flue gas conditions. The
role of the support in class 1 and 2 adsorbents has been show to play a crucial role in
both the kinetics and uptake capacity of the resultant materials. Furthermore activation
of the support via the incorporation on charged ions has also led to significant changes in
adsorption properties.
Although in their infancy, class 3 adsorbents have displayed excellent capacities
and kinetics at low CO2 partial pressure. They successfully incorporate the high amine
loadings of class 1 adsorbents with the covalent stability of covalent binding leading to
good stability. However little is known how to further improve these materials;
increasing the pore volume of the support led to an expected reduction in amine loading
and hence capacity. Little is known how to control the polymerisation to favour a specific
type of amine or to stop at a set chain length to avoid a reduction in mass transfer.

60
Literature review August 2012

At present most researchers focus on improving the adsorption capacity of a


material which is often measured under favourable conditions. The development of a
practical adsorbent is dependent on capacity of the material but also the kinetics,
regenerability, lifetime and process integration are just some of the other crucial
parameters to be considered. An interesting point made by Pirngruber and Llewellyn is
that often capacities are quoted in mmol/g which flatters lightweight MOF capacities
however the volume of the material is an important consideration in process design.
Studies into the performance of these materials under real flue gas conditions are
required to understand the deactivation mechanisms and then adapt to inhibit this
behaviour. To date most cycling studies feature an inert gas purge to regenerate the
sorbent; a luxury which is not practical on the proposed scale of CO 2 capture systems.
The number of cycles after which a material is deemed stable is approximately ten in the
current literature but several thousand in the case of an operational system. This
discrepancy needs to be addressed and rapid aging methods need to be developed to
test the adsorbent stability. The costs of adsorbent materials are rarely mentioned in
the literature with little consideration given to the target cost of less than $10 /kg-1
sorbent.
To integrate solid adsorbents into an operational system a plethora of data is
required before design work can commence. The kinetic data for adsorption, particle
size/shape effects, bed configurations, mechanical strength of particles, adhesive
interactions, required contact time, diffusional effects and regeneration conditions must
all be characterised before implementation into a fully functioning system. The sorbents
are only one part of the capture system and needs to be fully integrated into the overall
power plant process and its limitations.

61
Literature review August 2012

5.0 Outlook

Solid support amines have potential to be excellent adsorbents for CO 2 capture,


exhibiting the required capacity, kinetics and regeneration behaviour. A balance needs to
be found between capture capacity and adsorption kinetics allowing optimal adsorbent
performance. The structure-activity relationships, synergistic effects between support
and adsorbent need to be fully understood allowing iterative improvements in adsorbent
performance. Furthermore work into the behaviour of such materials under realistic
conditions will facilitate the next stages of development.
Metal organic frameworks although new show promise for superb adsorbent
materials. Their modular design and predictable structures allow for rapid optimisation
and tailoring towards specific applications. The application of these materials as low
pressure chemisorbents requires further investigation towards the emerging area of post
synthetic modification and vacant metal coordination sites. Framework stability under
water and other species is a principal limitation of MOFs with few studies conducted
under real flue gas conditions. The potential costs and availability of metals needs to be
carefully considered when designing MOFs as practical solid adsorbents.
With enhanced understanding second and third generation solid adsorbents could
play an active role in future post combustion capture systems. The current research
challenges are to improve adsorbent performance under realistic conditions within the
confines of cost and regeneration limitations. Success in many of the targets outlined in
section 1 would ultimately lead to the development and integration of solid adsorbents
into carbon capture systems.

62

S-ar putea să vă placă și