You are on page 1of 15

Chemical Engineering Journal 171 (2011) 760774

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Flue gas treatment via CO2 adsorption


Abdelhamid Sayari a,b, , Youssef Belmabkhout a , Rodrigo Serna-Guerrero b
a
b

Department of Chemistry, University of Ottawa, 10 Marie Curie, Ottawa, ON, Canada K1N 6N5
Department of Chemical and Biological Engineering, University of Ottawa, 161 Louis Pasteur, Ottawa, ON, Canada K1N 6N5

a r t i c l e

i n f o

Article history:
Received 20 June 2010
Received in revised form 14 January 2011
Accepted 5 February 2011
Keywords:
CO2 adsorption
Zeolite
Carbon
MOFs
Supported amines

a b s t r a c t
Adsorption separation has gained considerable attention as a viable alternative to the currently used,
high energy-demanding aqueous amine scrubbing technologies. This review is a summary of the main
contributions regarding the development of new adsorbents for post-combustion CO2 capture. Emphasis
has been placed on materials evaluated at representative ue gas conditions of CO2 partial pressure (i.e.,
0.050.2 bar) and temperature (2575 C). Whenever possible, the effect of moisture on the adsorbent
stability and CO2 uptake is included, although relatively few studies in the literature have focused on this
issue. This review includes adsorbents produced by modication of existing commercial materials as well
as newly developed materials. These adsorbents were separated in two major classes, namely (i) physical
adsorbents including carbons, zeolites and metal-organic frameworks and (ii) chemical adsorbents, i.e.,
amine-functionalized materials. A critical analysis of the literature is provided with the aim of tracing
the main paths currently pursued toward the development of suitable CO2 adsorbents and to provide a
general overview of the advantages and limitations of each family of adsorbents.
2011 Elsevier B.V. All rights reserved.

1. Introduction
As the concentration of carbon dioxide (CO2 ) in the atmosphere
keeps increasing, serious concerns have been raised with respect
to its impact on the environment. Since it started being monitored
in 1958, the increase of CO2 concentration in the atmosphere has
accelerated from less than 1 ppm/yr prior to 1970 to more than
2 ppm/yr in recent years [1]. As a result, the atmospheric level of
CO2 increased from 315 ppm in 1958 to 385 ppm in 2009 [1,2].
CO2 is considered to be the main anthropogenic contributor to
the greenhouse gas effect, as it is allegedly responsible for 60% of
the increase in atmospheric temperature, commonly referred to as
global warming [2,3]. Among the various sources of CO2 , approximately 30% is generated by fossil fuel power plants, making them
major contributors to global warming [4]. Despite their impact on
the environment, it is acknowledged that fossil fuels will remain the
leading source of energy for years to come, for both power generation and vehicle transportation. Therefore, it is critical to develop
effective methods for the capture and sequestration of CO2 from
post-combustion efuents, such as ue gas. Some reviews dealing
with the main sources of CO2 and potential strategies to prevent
their release to the environment are available in the literature [5,6].
Gas absorption using alkanolamine solutions has been used for CO2

Corresponding author at: Department of Chemistry, University of Ottawa,


10 Marie Curie, Ottawa, ON, Canada K1N 6N5.
E-mail address: abdel.sayari@uottawa.ca (A. Sayari).
1385-8947/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.02.007

scrubbing on industrial scale for decades. However, this process has


a number of shortcomings. For example, it generates severe corrosion of the equipment, and the regeneration of amine solutions
is highly energy intensive [7]. These drawbacks have been widely
documented, prompting a search for alternative technologies. One
viable route is adsorption which, compared to other separation processes, is recognized to be attractive to complement or replace the
current absorption technology due to its low energy requirement
[4,8]. Therefore, the use of appropriate adsorbents may potentially
reduce the cost associated with CO2 separation in the overall carbon
capture and storage (CCS) strategy.
Suitable adsorbents for CO2 removal from ue gas should combine several attributes, including:
(i) High CO2 adsorption capacity: CO2 equilibrium adsorption capacity is one of the main properties used to screen
new adsorbents. Knowledge of the equilibrium adsorption
isotherms is of prime importance for early evaluation of potential adsorbents. Whenever possible, this review will be focused
on adsorption properties measured under conditions relevant
to ue gas treatment, i.e., less than 0.4 bar CO2 partial pressure with a total gas pressure of 12 bar and temperature
below 7080 C. As a rule of thumb, Ho et al. [9] suggested that
an optimum adsorbent for CO2 capture from ue gas, should
exhibit a CO2 adsorption capacity of 24 mmol/g. It is well
established that from the slope of the adsorption isotherm at
low pressure, it is possible to estimate the adsorbate afnity for a given adsorbent. Thus, in terms of CO2 uptake, the

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

(ii)

(iii)

(iv)

(v)

(vi)

(vii)

ideal materials should exhibit a CO2 adsorption isotherm with


steep slope (favorable CO2 adsorption isotherm) corresponding to high uptake at low CO2 partial pressure. A less steep
slope (unfavorable CO2 adsorption isotherm) is indicative of a
lower afnity toward CO2 .
Fast kinetics: Adsorption kinetics affects primarily the working adsorption capacity in dynamic processes such as
adsorption in a xed bed column. A suitable CO2 adsorbent
will have a high rate of adsorption, resulting in a working
capacity close to equilibrium capacity over a wide range of
operating conditions. However, determination of kinetic properties such as diffusion is one of the most challenging issues in
adsorption science, as it involves parameters not always readily available, such as particle size of the adsorbent and use of
adequate experimental set-ups and conditions.
High CO2 selectivity: The adsorbent selectivity toward CO2 has
a direct impact on the degree of purity of the product. This in
turn, plays a major role in the economics of the CO2 adsorption
process [9]. Ideally, an adsorbent for ue gas treatment will not
adsorb any nitrogen.
Mild conditions for regeneration: The ease of regeneration of
the adsorbent is a key property in the selection of materials
for CO2 separation. Depending on the structural and chemical properties of the adsorbent, adsorptiondesorption cycling
may be achieved via temperature, pressure (or vacuum), concentration swing adsorption or a combination thereof. In
practice, incorporation of functional groups can be used to
modify the adsorbentadsorbate interactions (e.g., Van der
Waals, electrostatic, hydrogen bonding or acidbase interactions) and affect the CO2 uptake and selectivity. Optimum
interactions should be neither too weak nor too strong. Too
weak bonding results in low CO2 adsorption capacity at low
pressure, but easy regeneration. Conversely, strong bonding
induces high adsorption capacity but desorption will be difcult and costly.
Stability during extensive adsorptiondesorption cycling: The
lifetime of adsorbents, which determines the frequency of
their replacement, is a critical property of equal importance as
the CO2 adsorption capacity, selectivity and kinetics, because
of its direct impact on the economics of any commercial scale
operation.
Tolerance to the presence of moisture and other impurities in
the feed: In addition to CO2 and N2 , ue gas contains water
vapor and other impurities such as O2 and SO2 . The degree
of tolerance and the afnity of the adsorbent to such impurities may affect signicantly the strategy to be used, with
direct impact on the overall economics of the CO2 separation process. Moisture is known to adversely affect CO2 uptake
in a variety of physical adsorbents such as zeolites and activated carbon. Consequently, the strategies proposed for CO2
adsorption from ue gas is likely to include an upstream drying step. As a result, the overwhelming majority of published
reports dealing with physical adsorbents have not examined
moisture effects. Whenever possible, this review will provide
a general picture on the behavior of the adsorbents in the
presence of moisture. It is also generally established that CO2
adsorbents have high afnity to SO2 and even some afnity
toward NOX , which may adversely affect the CO2 adsorption capability of the material. Thus, abatement of SO2 and
NOX from ue gas prior to CO2 capture is required in most
cases.
Low cost: This is another important parameter to be considered in the development of any potential adsorbent. At
this stage, information on adsorbent cost and other economic
considerations are rather scarce in the literature. Thus, costrelated issues will not be discussed in this review.

761

Because ue gas is generally cooled down to ca. 55 C, to allow


appropriate conditions for SO2 and NOX abatement [10,11], whenever possible, this review will be focused on literature reports
dealing with CO2 adsorption using 520% CO2 -containing mixtures
with a total pressure of 12 bar, and temperature between 25 and
70 C. For a more comprehensive account on CO2 adsorbents in general, the reader may refer to an excellent review by Choi et al. [12].
Similarly, the eld of high temperature CO2 capture has also been
reviewed by Lee et al. [13].
Adsorbents for CO2 capture can be categorized in many ways,
based on their chemical composition, structural characteristics or
according to the adsorption mechanism involved, i.e., physical vs.
chemical. Physical adsorbents for CO2 capture include carbon materials, alumino-silicas such as zeolites, alumino-phosphates (AlPOs)
and alumino-silico-phosphates (SAPOs), and more recently metal
organic frameworks (MOFs). The CO2 chemical adsorbents discussed in this review refer to those obtained through incorporation
of amine groups into solid supports such as mesoporous silica.
Consequently, in this review we distinguished two classes of CO2
adsorbents for stack gas treatment, namely physical and aminefunctionalized adsorbents.

2. Physical adsorbents
2.1. Carbons
Because of their wide availability, low cost and high thermal stability, it is largely established that activated carbons have
advantages over other CO2 adsorbents. Among the carbon based
adsorbents reported in the literature, activated carbons (ACs) and
carbon nanotubes (CNTs) are the most investigated materials. CO2
adsorption on activated carbons has been studied experimentally
and theoretically for a long time [14] and has found commercial
applications [15,16]. There is a wide range of activated carbons with
a large variety of microporous and mesoporous structures. Activated carbon may be produced from many raw materials such as
coal, coke pitch, wood or biomass sources (e.g., saw dust, coconut
shells, olive stones), often via two steps: carbonization and activation [17]. Carbon molecular sieves (CMS), which are a sub-class
of activated carbon with narrow pore size distribution (PSD), are
kinetic-based adsorbents. They have been commercialized mainly
for the separation of air and the production of high purity N2
[18,19]. However, at low CO2 partial pressure, activated carbons
exhibit lower adsorption capacity and selectivity than zeolites due
mainly to their less favorable adsorption isotherms. In spite of
the hydrophobic character of carbon-based adsorbents, their CO2
adsorption ability is adversely affected by the presence of water
vapor [20].
Table 1 shows literature data on CO2 adsorption capacity and
selectivity of activated carbons and carbon nanotubes in the partial pressure range of 0.10.4 bar at 298333 K. Considering 1 and
2 bar as the lowest and highest total pressure of ue gas, the 0.1 and
0.4 bar were chosen arbitrarily as the lowest and the highest CO2
partial pressure relevant to ue gas treatment. Notice that most
studies dealing with CO2 adsorption on activated carbons were
undertaken at high pressure and room temperature.
It is important to notice that, although adsorption capacity
varies considerably for different activated carbons at high pressure
[24,25], the adsorption capacity at low pressure is less sensitive to
the nature of carbons. As seen in Table 1, the typical CO2 equilibrium adsorption capacity for activated carbons at a partial pressure
of 0.1 bar is 1.1 mmol/g at room temperature but decreases rapidly
to 0.25 mmol/g at 328 K. In terms of CO2 adsorption capacity, activated carbons may be particularly interesting for CO2 removal but
only at high pressure. For example, Himeno et al. [24] showed

762

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

Table 1
Literature survey on CO2 adsorption properties of activated carbons and carbon nanotubes at low pressure.
N2 adsorption capacity at
0.91.6 bar (mmol/g)

CO2 /N2 capacity molar


ratio

References

0.61.5

0.5
0.75

1.2
2

[21]

328

0.250.8

0.2
0.35

1.25
2.28

[21]

308
333

0.51.25
0.340.91

[22]
[23]

Carbon material

Temperature (K)

AC

298

AC

SWCNT
MWCNT

CO2 adsorption capacity at


0.10.4 bar (mmol/g)

, not available.

that the adsorption capacity on Maxsorb activated carbon reached


ca. 113 wt% (25.7 mmol/g) at 30 bar and room temperature. The
CO2 vs. N2 selectivity at room temperature and above atmospheric
pressure was reported to be 14 [25]. However, to the best of our
knowledge, CO2 vs. N2 selectivity at low pressure was not reported
in the literature. As shown in Table 1, based on the ratio of CO2 and
N2 uptake (CO2 /N2 ratio), it is inferred that the CO2 vs. N2 selectivity is rather small and weakly dependent on temperature. This
behavior of activated carbons was also pointed out by Kurniwan
et al. [26]. Dreisbach et al. [25] found that CO2 selectivity over N2
for Norit type activated carbon is insensitive to the CO2 pressure
change. In terms of regenerability, in light of the low CO2 adsorption
enthalpy (22 kJ/mol) [24], it is generally established that activated
carbons are easily regenerated allowing their use in pressure swing
adsorption (PSA). However, some activated carbons may contain
functional groups that interact strongly with CO2 , thus preventing
complete desorption. CO2 adsorption kinetics on activated carbons
is generally comparable to zeolites and depends on the type of
diffusion involved, i.e., macropore, micropore or surface diffusion
[2729] and the heterogeneity of the material. A number of investigations dealt with activated carbon performance in terms of process
productivity using PSA [21,30,31].
More recently, carbon nanotubes (CNT) have emerged as a new
class of materials with more homogenous surface properties and
well dened pore shape and size [32]. Although a large number
of theoretical studies dealing with CO2 adsorption on CNT have
been reported, the adsorption capacity and selectivity at low CO2
partial pressure were similar to typical porous carbon materials
such as activated carbons. However, Cinke et al. [22] found that the
adsorption capacity of single-walled carbon nanotubes at 308 K (ca.
1 mmol/g) is twice that of activated carbon in a wide range of CO2
concentration. On the contrary, Su et al. [23] found that at room
temperature, multi-walled CNTs exhibit much lower CO2 adsorption capacity than activated carbons. Using molecular simulation,
Huang et al. [33] showed that regardless of the pore diameter, pressure and temperature, CNTs show slightly higher selectivity toward
CO2 than activated carbons.
In summary, carbonaceous material seem to be interesting only
for CO2 removal at high pressure and low (e.g., room) temperature.
These limitations may not be suitable for low pressure CO2 capture
from ue gas treatment. However, as discussed later, the strategy
to increase the strength of CO2 interactions with such materials at
low partial pressure via surface modication may hold promise.

For instance, the Si/Al ratio and the nature of extraframework


cations, can be varied systematically [43,44], playing a major role
in controlling the CO2 adsorptive properties. Exhaustive studies
[4345] reported that the most promising zeolites for CO2 adsorption are characterized by a low Si/Al ratio, corresponding to high
content of extraframework cations. The presence of aluminum
atoms in these silicate-based molecular sieves introduces negative framework charges that are compensated with exchangeable
cations within the pores. There are often alkali cations such as
sodium or lithium that generate strong electrostatic interactions
with carbon dioxide. The number and nature of extraframework
cations affect the CO2 adsorption properties of zeolites. Maurin
et al. [45] investigated experimentally and theoretically, the CO2
adsorption capacity and enthalpy for several faujasite-type zeolites
in sodium form, with different Si/Al ratios, namely, dealuminated
NaY (DAY, Si/Al = ), NaY (Si/Al = 2.4) and low silica NaX (NaLSX,
Si/Al = 1). At low pressure for both adsorption and desorption, 13X
(or NaX with Si/Al = 1.25) and NaY were claimed to be the most suitable adsorbents with favorable CO2 adsorption isotherms. Purely
siliceous NaY [45], silicalite, ITQ3 and ITQ7 [36] exhibited at least
12 times lower adsorption capacity than NaY and NaLSX at 0.1 bar
and room temperature. Walton et al. [44] showed that substitution of Na cations by Rb, Cs, K and Li has a signicant effect
on the adsorption capacity in a wide range of CO2 concentration
(Fig. 1). The adsorption capacity at 0.1 bar followed the sequence
Cs < Rb K < Li Na for X zeolites. In a more recent contribution
[46], the inverse sequence was observed for Y faujasite. It was
shown that the total or partial substitution of Na cations by Cs
and K cations in Y faujasite led to increased adsorption capacity
at low pressure (Fig. 2), because they are associated with a more

2.2. Zeolites and zeolite-like materials


Zeolites, which are highly ordered microporous crystalline
materials [34], were heavily investigated because they are probably the most promising materials for adsorption and separation
of CO2 [2542]. The great interest of zeolites stems from the fact
that a number of their properties such as pore size and architecture or chemical composition affect their adsorption performance.

Fig. 1. Adsorption isotherms of CO2 on cation exchanged faujasite X [44].

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

763

Table 2
Literature survey on CO2 adsorption properties of some zeolites and zeolite-like materials at low pressure.
Zeolites/Si/Al ratio

CO2 adsorption
temperature (K)

NaX/1
NaX/1
LiX/1
NaY/2.4
CsY/2.4
KY/2.4
Silicalite/
H-ZSM-5/30
Li-MCM-22/15

298
323
303
323
333
333
334
313
333

Adsorption capacity at
0.10.4 bar (mmol/g)
2.83.9
1.432.49
3.14.6
0.451.17
0.861.2
0.751.6
0.160.45
0.71.5
0.681

N2 adsorption capacity
at 0.91.6 bar (mmol/g)

CO2 /N2 capacity molar


ratio

Reference

0.2640.46

0.1
0.23

118.5

1.6
3

[48]
[48]
[44]
[49]
[46]
[46]
[50]
[51]
[52]

, not available.

favorable adsorption isotherm. This was explained by the dominant acidbase (CO2 -framework oxygen atom) interaction over the
polarizing effect in the case of CsY and KY faujasites (particularly
for CsY and to a lesser extent for KY), in contrast to LiY, NaY and X
faujasites.
Table 2 shows the CO2 adsorption properties of different zeolites and zeolite-like materials. As seen, the adsorption capacity
decreased drastically when the temperature increased from 298
to 323 K. Akten et al. [47] showed that the CO2 /N2 selectivity for
Na-4A type zeolite also decreased at increased temperature.
In terms of CO2 adsorption kinetics, zeolites are ranked among
the fastest adsorbents, reaching equilibrium capacity within minutes [12]. Moreover, a large number of studies were devoted to
NaX faujasite using different recycling congurations, including
temperature swing and pressure swing adsorption [9,40,41,53].
Although the CO2 adsorption enthalpy on X and Y zeolites was
found to be dependent on the nature of extraframework cations,
within the range of 3050 kJ/mol, it is low enough to allow
reversible CO2 adsorption. Zeolites generally operate without any
loss in performance, provided that the feed stream is strictly dry.
Although low silica materials exhibit high adsorption capacity and
selectivity at low pressure with favorable isotherms, they are very
sensitive to the presence of water, which strongly inhibits the
adsorption of CO2 [54]. This prompted some investigations on the
ability of hydrophobic high silica zeolites such as MWW zeotype
[52] and NaZSM-5 [55] to remove CO2 . However, because high
silica microporous materials contain less extraframework cations
than faujasite zeolites, they exhibit lower adsorption capacity and
adsorption enthalpy [45]. Moreover, similarly to X and Y zeolites,
they show decreasing selectivity at increasing temperature [47,56].

Fig. 2. Adsorption isotherms of CO2 on cation-exchanged Y faujasites [46].

Alumino-phosphates (AlPO) and silica-alumino-phosphates


(SAPO) are another class of zeolitic materials that were investigated as potential CO2 adsorbents [57,58]. The overall framework
of AlPOs is neutral and is expected to behave as silicalite or dealuminated Y faujasite for CO2 as was shown by Deroche et al. [58]
using a combination of molecular simulation and microcalorimetry.
In fact AlPO-18 (with AEI structure) exhibits unfavorable adsorption isotherm. Similarly to NaX and NaY, the framework of some
SAPOs is negatively charged and the overall charge is balanced by
extraframework cations. In this case, it is expected to obtain a more
favorable CO2 adsorption isotherm with higher adsorption capacity at low pressure of CO2 as reported by Castro et al. [57] for the
proton form of SAPO-34. However, the CO2 adsorption capacity on
SAPO remains lower than X and Y faujasites.
In conclusion, because of their often highly favorable CO2
adsorption isotherms, zeolites and zeolite-like materials with low
Si/Al ratios are among the most promising adsorbents for CO2 capture from ue gas. However, because of their highly hydrophilic
character, the ue gas needs extensive drying prior to CO2 capture.
Notice that among zeolites, 13X is has been the most investigated
material for the purpose of CO2 capture [9,40,41,53,59]. As pointed
out by Ho et al. [9], further work to develop more selective zeolite
adsorbents toward CO2 vs. N2 and O2 may reduce considerably the
cost of CO2 capture.

2.3. MOFs and zeolite-like MOFs


Although an emerging class of porous materials, metal organic
frameworks (MOFs) have attracted a growing interest, motivating
extensive studies on their CO2 adsorption properties, both theoretically and experimentally. MOFs are porous crystalline materials
composed of self-assembled metallic species and organic linkers
[60,61]. Their pore size and shape can be easily tuned by changing either the organic ligands or the metallic clusters. They are
typically rigid materials, but some of them exhibit structural exibility upon adsorption and desorption of gases or liquids [62,63].
The wide range of MOFs combined with their desirable properties such as their remarkably high surface area and controlled
pore size and shape, prompted extensive work on their adsorptive properties, particularly for storage of light gases (H2 , CH4 ) and
storage and separation of CO2 . Although the majority of investigations on CO2 adsorption over MOFs used pure CO2 , as well
as CO2 -containing mixtures, most measurements and simulations
were carried out at high pressure and often at room or subambient
temperature. Seminal contributions in the synthesis of novel MOFs
and their CO2 adsorption properties were reported by Millward and
Yaghi [64]. Their early work was followed by an extensive effort to
develop new types of MOFs for the separation and storage of CO2
[6572]. Millward and Yaghi [64] showed that MOF-117 exhibits
an unprecedented CO2 adsorption capacity at high pressure (e.g.,
ca. 150 wt% at 40 bar), but very small CO2 uptake at subatmo-

764

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

Table 3
Literature survey on CO2 adsorption properties of some MOFs and ZMOFs.
MOFS

Temperature (K)

MOF-508
Cu-BTC
MIL-53
Ni/DOBDC
CO/BOBDC
Mg/DOBDC (Mg-MOF-74)
ZIF-78

323
298
303
296
296
296
298

CO2 adsorption capacity


(mmol/g at 0.10.4 bar)
0.10.7
0.52
0.51.15
2.74.01
2.85.36
5.366.8
0.771.36

N2 adsorption capacity
(mmol/g at 0.91.6 bar)

CO2 /N2 selectivity

Reference

0.60.9
0.25

2
15

50

[75]
[79]
[67]
[72,73]
[72,73]
[68,72]
[70,80]

, not available.

spheric pressure. More recently, Caskey et al. [72] reported a much


higher and reversible adsorption capacity for pure CO2 (23.6 wt%,
5.36 mmol g1 ) at 0.1 bar and room temperature on Mg/DOBDC (or
Mg-MOF-74) (Fig. 3). Although this is an excellent nding, adsorption of CO2 in mixtures with N2 was not reported. Additional data
regarding adsorption of CO2 at 0.1 bar on a number of MOFs may
be found elsewhere [73].
It is important to mention that the majority of MOFs exhibit
unfavorable adsorption isotherms for CO2 in the low pressure
range. Moreover most of these materials adsorb considerable
amounts of N2 , leading to low selectivity toward CO2 . The highest
selectivity toward CO2 vs. N2 was in the range of 530 [7476], and
generally the CO2 adsorption capacity is dramatically reduced at
higher temperature, accompanied by a drop in the CO2 adsorption
selectivity. In terms of kinetics, MOFs are as fast CO2 adsorbents as
zeolites according to some computational studies [6677]. Based on
the aforementioned observations, MOFs seem to be more suitable
for CO2 storage rather than separation.
Table 3 shows the adsorption properties of different types of
MOFs at low pressure. As seen, the adsorption capacity as well as
the CO2 /N2 selectivity for most MOFs, were very low and decreased
drastically when the temperature increased from 298 to 323 K. This
has been documented by Barcia et al. [78], Bastin et al. [75], Bae et al.
[74], Yang et al. [79] who showed decreasing CO2 /N2 selectivity at
increased temperature for MOF-508b and Cu-BTC.
MOFs and ZMOFs structural, chemical and thermal stability has
been hardly addressed in the literature, until recently. It is generally
recognized that by far the most critical issue for the stability of these
materials is their hydrothermal stability. The behavior of MOFs
and their subfamilies in hydrated conditions varies widely, from
materials that irreversibly degrade even under mild conditions to
materials that are highly stable in boiling water. For example, MOF-

117 and IRMOF-1 were reported to be unstable upon exposure to


air in the presence of humidity [8183]. The concerns raised by the
stability of MOFs prompted the discovery of a new class of MOFs
referred to as zeolite-like MOFs (ZMOFs) or zeolitic imidazolate
frameworks (ZIF). ZMOFs are crystalline porous materials which
combine the highly desirable properties of zeolites and MOFs,
such as microporosity, high surface areas, and exceptional thermal and chemical stability [6981]. Because of the strong bonding
between the imidazolate linker and the metal center, many ZMOFs
have high thermal (>673 K) and moisture resistance compared to
other MOF structures [84,85]. Although signicant improvement
was observed in terms of CO2 vs. N2 adsorption selectivity, which
increased up to 50 for ZIF-78 [80], the CO2 adsorption capacity was
still low at low CO2 partial pressure (Table 3).
Zeolites, MOFs and ZMOFs are typically hydrophilic and their
application for CO2 capture from ue gas requires partial or complete drying of the gas stream. To circumvent this limitation,
new materials with no hydrophilic adsorption sites referred to
as covalent organic frameworks (COFs) were developed. COFs are
crystalline organic porous materials without metal ions. Furukawa
and Yaghi [86] and Babarao and Jiang [87] reported high CO2
adsorption capacities for a series of COFs, but adsorption at low
partial pressure of CO2 appeared to be signicantly lower than for
Mg-MOF-74.
In summary, MOFs, ZMOFs and COFs may be promising materials for CO2 removal provided that more favorable CO2 adsorption
isotherms are obtained. Their selectivity and capacity at low partial pressure of CO2 in gas mixtures are quite low and more likely
to be suitable for CO2 storage rather than CO2 separation from ue
gas. Although in their early stages of development, MOFs, ZMOFs
and COFs are promising materials for CO2 adsorption showing very
interesting and adjustable properties.
3. Amine-functionalized adsorbents
The technology currently used in industry for CO2 capture is
absorption with liquid amine solutions. The removal of CO2 by
amines occurs via the widely accepted formation of carbamate and
bicarbonate species, as represented in Scheme 1 [88]. These are
reversible reactions that permit the regeneration of amines, typically by heating the CO2 -rich solution.
2(RNH2 ) + CO2 RNHCO2 RNH3 +
carbamate

RNH

2
2
+
RNH2 + CO2 + H2 O RNH3 + HCO3 (RNH
3 )2 CO3

bicarbonate

Fig. 3. CO2 adsorption isotherms (296 K, 01 atm) for M/DOBDC materials. Inset is a
close-up of the low pressure region. Filled and open symbols represent adsorption
and desorption data, respectively [72].

carbonate

The liquid amine absorption process inspired researchers to use


amine-modied solid materials as adsorbents for CO2 capture. As
far as ue gas treatment is concerned, it was anticipated that supported amines will maintain a high selectivity toward CO2 with a
negligible uptake of other components, particularly N2 , but without the aforementioned drawbacks associated with aqueous amine
solutions. According to a study on amine-functional adsorbents by
Gray et al. [89], a capacity of at least 3 mmol/g is required for this

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

765

Scheme 1. Typical reaction pathway between CO2 and amines [88].

process to be competitive against absorption technologies, corroborating the previously mentioned range of 24 mmol/g proposed
by Ho et al. [9]. Although the early efforts to produce aminefunctionalized adsorbents were not particularly successful in terms
of adsorption capacity, the collective effort of several research
groups resulted in signicant performance improvements, leading
to increasing interest in this subject matter. Based on ISI Web of
Knowledge, Fig. 4 illustrates the remarkable increase in the number
of publications related to CO2 adsorption on amine-functionalized
materials, with ca. 70% of these contributions published in the last
6 years.
We have broadly organized the present section according to
the type of interactions between amine groups and the support,
namely (i) amine-impregnated materials where mostly weak interactions occur, and (ii) covalently bonded amine-containing species,
obtained typically via surface-grafting of aminosilanes. The rationale behind such classication is that materials with either strong
or weak interactions exhibit a number of common characteristics.
An example is that grafted materials offer comparatively higher
rate of adsorption than amine-impregnated adsorbents [91] and,
in some cases even higher than commercial adsorbents such as
13X [90]. However, the organic content of amine-grafted adsorbents depends on the surface density of hydroxyl groups, needed
to anchor the aminosilane. As for impregnated amines, higher loadings may be achieved, but often accompanied by increasingly strong
diffusion limitations.
3.1. Amine-impregnated materials
3.1.1. Ordered mesoporous supports
Xu et al. [92] were rst to report on polyethyleneimine (PEI)impregnated mesoporous materials for CO2 adsorption, coining
the term molecular basket. It was found that the adsorption
capacity of PEI-impregnated MCM-41 improved at increased loading. The highest value of adsorption capacity, corresponding to
3.02 mmol/g was obtained under a stream of pure CO2 at 75 C
using a sample with 75 wt% PEI. However, the maximum efciency, i.e., CO2 /PEI molar ratio was obtained in the presence of

Fig. 4. Number of publications related to CO2 capture on amine-functional materials


according to ISI-Web of Science database.

a material containing 50 wt% PEI, and decreased steadily at higher


loadings. As discussed later, this behavior was conrmed by other
researchers. Further work under conditions relevant to ue gas
treatment showed that the 50% PEI on MCM-41 silica exhibits an
adsorption capacity of ca. 2.1 mmol/g in the presence of 10% CO2 /N2
at 75 C. A particularly interesting behavior of PEI-impregnated
MCM-41 materials was the fact that, unlike other adsorbents,
adsorption capacity improved as temperature increased from 25 to
75 C. Since the actual adsorption event is exothermic in nature, the
increasing adsorption capacity with temperature was attributed
to the occurrence of a bulk-like state of PEI inside the mesopores, with amine groups not readily accessible at low temperature,
resulting in a diffusion-limited process. Since then, other authors
working on PEI-impregnated materials reported similar ndings
and the idea of a diffusion-limited process has been generally
accepted.
Following their early study on PEI-impregnated MCM-41, Xu
et al. [93] analyzed the adsorption of CO2 in humid streams. A positive effect of moisture was observed in terms of increased capacity,
particularly when the molar concentration of water was equal or
lower to that of CO2 , providing support to the bicarbonate formation mechanism. No further increase in adsorption capacity was
observed for streams with higher moisture content. Accordingly,
its capacity was enhanced from 2.01 mmol/g in a simulated dry ue
gas containing 15% CO2 to 2.84 mmol/g in a stream containing 10%
moisture and 13% CO2 , balance air. In a later contribution, the same
group [94] impregnated PEI on SBA-15 under the assumption that
the structural characteristics of the support would affect the performance of the aminated adsorbents. Allegedly, due to the larger
pore size and volume of SBA-15 compared to MCM-41, PEI-SBA-15
used the amine groups more efciently under the same loading of
50 wt% PEI. Indeed, as mentioned above, while PEI-MCM-41 exhibited an adsorption capacity of 2.1 mmol/g, PEI-SBA-15 showed a
capacity of 3.18 mmol g1 under a ow of 15% CO2 balance air at
75 C.
Ahns group [95,96] has also investigated the adsorptive properties of PEI-impregnated on a variety of ordered mesoporous silicas.
It was found that at constant PEI loading, the use of various supports afforded different adsorption capacities, and that supported
PEI had a higher capacity than its pure liquid counterpart. Interestingly, under otherwise the same conditions, the adsorption capacity
appeared to be dependent on pore diameter (dp ). When PEI was dispersed on a KIT-6 type silica with 6 nm dp at a loading of 50%, the
material adsorbed 3.07 mmol/g in a stream of pure CO2 at 75 C,
vs. 2.52 mmol/g when using MCM-41 with 2.8 nm pores as support. The adsorption capacity of 50% PEI-loaded KIT-6 in conditions
closer to ue gas was 1.95 mmol/g in the presence of 5% CO2 /N2
at 75 C. The pore size can also affect the rate of adsorption as the
time required to achieve 90% of the total capacity was in the order of
KIT-6 < SBA-16 = SBA-15 < MCM-48 < MCM-41. Following the rationale that using large pore sizes afford better adsorption capacity,
PEI and tetraethylenepentamine (TEPA) were impregnated on a silica monolith with hierarchical pore structure [96]. Due to its larger
dp , with mean values at 3, 17 and 120 nm, the optimum loading
of PEI was 65 wt%, with a capacity of 3.75 mmol/g for a stream of
5% CO2 in N2 at 75 C, a capacity much higher than that obtained
using conventional MCM-41 mesoporous silica as support. TEPA-

766

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

loaded samples were not particularly attractive as their adsorption


capacity deteriorated after only 5 adsorptiondesorption cycles.
Another approach used by Yue et al. [97] consisted of impregnating TEPA on as-synthesized SBA-15, as opposed to calcined support.
The reported adsorption capacity for a sample with 50 wt% loading in the presence of 10% CO2 in N2 was ca. 3.25 mmol/g at 75 C.
The materials prepared with as-synthesized support consistently
performed better than those prepared using the corresponding calcined supports, with up to 10% higher capacity with the added
advantage that no steps are required to remove the organic template. The proposed explanation was that the polymeric template
in the pores of as-synthesized supports interacts with the TEPA,
forming a more even distribution of the functional groups, and preventing TEPA from aggregating into a micellar-like form, which
is believed to be its naturally occurring form. The use of assynthesized supports was further explored by Yue et al. [98] using
MCM-41-type silica. A TEPA-MCM-41 sample loaded with 50 wt%
PEI had a capacity of 4.54 mmol/g for 5% CO2 /N2 at 75 C, outperforming the above-mentioned TEPA-SBA-15.
In addition, Yue et al. [99] impregnated as-synthesized SBA-15
using a mixture of TEPA and diethanolamine (DEA). In this case, the
maximum CO2 /N ratio (adsorption efciency) was found to be ca.
0.4 at a loading of ca. 30% TEPA and 20% DEA. In the case of TEPADEA-SBA-15 at 75 C, its adsorption capacity ranged from 3.77 to
3.61 mmol/g throughout 6 adsorptiondesorption cycles for 5% CO2
in N2 . The good performance of this adsorbent was attributed to the
hydroxyl groups present in DEA. This is analogous to the effect of
water vapor, which is associated with a more favorable CO2 to N
stoichiometry, as shown in Scheme 1.
Franchi et al. [100] impregnated DEA on a variety of supports whose adsorption capacity and stability were compared to
a benchmark adsorbent, i.e., 13X zeolite. The most promising support used in this work consisted of MCM-41 silica with pores
enlarged by post-synthesis treatment (PE-MCM-41, dp = 9.7 nm),
which afforded an adsorption capacity of 3 mmol/g at 25 C in the
presence of 10% CO2 in N2 , a value comparatively higher than
that reported for 13X (i.e., 2.8 mmol g1 ) under the same conditions. Similarly to supported PEI, the adsorption capacity increased
with DEA loading, but the adsorption efciency (CO2 /N) decreased,
suggesting that loading more than 6 mmol of DEA per gram of
adsorbent is unattractive.
A series of amine-impregnated mesoporous aluminas (MA),
using diisopropanolamine, triethanolamine, 2-amino-2-methyl1,3-propanediol, diethylenetriamine (DETA) and PEI, were investigated by Plaza et al. [101]. The most attractive materials were
those containing PEI and DETA with 40 wt% loading. Not only did
these adsorbents offer higher adsorption capacity at room temperature but, unlike the other materials in this study, their capacity
increased at higher temperature. DETA-MA exhibited a capacity of
ca. 1 mmol/g at 25 C for pure CO2 , increasing to ca. 1.4 mmol/g
at 57 C. For PEI-MA, the adsorption capacity at 57 C was ca.
1.14 mmol/g in the presence of pure CO2 . It is worth noting that
the adsorption capacity was enhanced at higher temperature only
for samples with high amine loadings, supporting the hypothesis
that this behavior is associated with diffusion limitations within
the amine phase at low temperature.
A number of literature reports explored the regeneration behavior of PEI-impregnated mesoporous materials. Drage et al. [11] used
a proprietary mesoporous silica impregnated with PEI (40 wt%).
The adsorbent showed a capacity of 2.4 mmol/g at 70 C in the
presence of 15% CO2 in N2 . This work analyzed the effect of regeneration temperature using pure CO2 as stripping gas. It was observed
that desorption at a temperature of less than 140 C resulted in
an incomplete regeneration. However, some concerns were raised
regarding the use of such high temperatures, mainly because of
the following problems: (i) evaporation of PEI may occur and (ii)

a secondary reaction between CO2 and amine groups formed a


stable product, most likely urea, resulting in a decreasing number of adsorption sites. The suggested alternative was to use a
different stripping gas or lower desorption temperature although
sacricing some working adsorption capacity. As discussed later, a
strategy to prevent the formation of urea during extensive cycling
even at high temperature has been proposed recently by Sayari
and Belmabkhout [102]. A PSA strategy was explored by Dasgupta
et al. [103] using 50% PEI-impregnated SBA-15. The highest capacity
reported at 75 C was 1.36 mmol/g for 12% CO2 in N2 . A steady state
was obtained after 1520 cycles, and the productivity was better
compared to similar PSA procedure using 13X zeolite at 75 C.
3.1.2. Ordered microporous supports
In addition to mesoporous materials, zeolites have also been
used as supports. Jadhav et al. [104] dispersed monoethanolamine
(MEA) on 13X zeolite producing materials with different loadings.
Quite interestingly, the adsorbent with the highest capacity at low
temperature (i.e., 35 C), with 1.96 mmol g1 for 15% CO2 in N2 ,
contained only 2.9 wt% MEA, while the best capacity at 75 C (i.e.,
0.45 mmol/g) was obtained on a sample with the highest loading (i.e., 25 wt%). These capacities were comparatively higher than
unmodied 13X, which adsorbed 0.64 and 0.36 mmol/g at 35 and
75 C, respectively. An interesting advantage of amine-containing
13X was a signicant improvement in its tolerance to moisture.
While it is generally accepted that preferential adsorption of water
on 13X results in a drastic reduction of CO2 uptake, the adsorption
capacity in the presence of 100% RH decreased by only ca. 13% with
respect to dry conditions.
Another type of zeolite, namely beta-zeolite, was used by Fisher
et al. [105] to support TEPA and compared with TEPA-impregnated
on amorphous alumina and silica. The results clearly showed the
advantages of using a support with good structural properties since
beta-zeolite was loaded with up to 38.4 wt% compared to only 14.6
and 8.3 wt% for SiO2 and Al2 O3 , respectively, most likely as a result
of a comparatively higher pore volume. Such loading translated in a
signicantly higher adsorption capacity for TEPA-beta zeolite over
the other samples, being 2.08 mmol/g for 10% CO2 balance nitrogen
at 30 C, while it was 0.19 and 0.68 mmol/g for TEPA-Al2 O3 and
TEPA-SiO2 , respectively.
3.1.3. Other supports
While ordered mesoporous supports are suitable substrates
for the dispersion of amines, other supports were also explored.
Extensive work performed by Filburns group [106108], dealt
with impregnation of a variety of amines, such as PEI,
monoethanolamine, diethanolamine, triethanolamine, and TEPA
on high surface area polymeric supports, mainly polymethylmethacrylate (PMMA). Although their original purpose was to
produce adsorbents for air purication in conned environments,
the results have proven to be of interest for other applications
such as ue gas treatment as shown in a later contribution [108],
where a TEPA-impregnated PMMA exhibited capacities of 21.45
and 13.88 mmol/g at 20 and 70 C, respectively in the presence of
15% CO2 and 2.6% H2 O balance N2 . It is also worth noting that the
reported adsorption capacities were remarkably higher than any
other data reported in the literature. Moreover, contrary to amineimpregnated mesoporous inorganic supports, adsorption capacity
decreased at higher temperature for the polymer-based adsorbents. In the same work, TEPA was reacted with acrylonitrile to
selectively transform primary amines into secondary amines before
impregnation, under the premise that secondary amines are advantageous. However, this was not the case, since the sample produced
by impregnation of the modied TEPA, referred to as TEPAN, underperformed TEPA-PMMA in terms of adsorption capacity and rate
with adsorption capacity values of 14.22 and 4.01 mmol/g at 20 and

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

767

Table 4
Literature data on CO2 adsorption capacity of amine-impregnated adsorbents.
Support

Amine

MCM-41
MCM-41
SBA-15
KIT-6
Monolith
As-synthesized SBA-15
As-synthesized MCM-41
As-synthesized SBA-15

PEI
PEI
PEI
PEI
PEI
TEPA
TEPA
TEPA + DEA

PE-MCM-41
Mesoporous Al2 O3
Mesoporous SiO2
SBA-15
PMMA
PMMA

DEA
DETA
PEI
PEI
TEPA
Ethyleneamine +
acrylonitrile
DBU
PEI
PEI
PEI
MEA
TEPA

PMMA
SiO2 (CARiACT)
PMMA (Diaion)
AOS carbon
13X
Beta-zeolite

Amine
loading (wt%)

Capacity
(mmol/g)

CO2 /N

50
50
50
50
65
50
50
50 (30% TEPA,
20% DEA)
76
40
40
50
41
Proprietary
information
30
40
40
5
25
38

2.1
2.84
3.18
1.95
3.75
3.25
4.54
3.77
3
1.4
2.4
1.36
13.88
4.18
2.34
3.95
3.60
1.98
0.45
2.08

70 C, respectively. Furthermore, TEPAN-PMMA was not stable as


its capacity decreased considerably, e.g., from 4.01 to 1.68 mmol/g
after only three cycles with adsorption at 25 C and regeneration at
70 C.
The hypothesis that secondary amines perform better than
primary amines was also proposed by Gray et al. [109] in a
comparative study between ethyleneamine-impregnated polymer
and aminopropyl-grafted SBA-15. The reported capacity of the
latter was 2.01 mmol/g under a humid stream of 10% CO2 in
N2 at 25 C, a higher value than the ethyleneamine containing
adsorbent, i.e., 1.92 mmol/g. Another sample, produced by loading
ethyleneamine after reaction with acrylonitrile, had an adsorption
capacity of 4.18 mmol/g, although it exhibited poor stability, as
capacity decreased to 2.69 mmol/g after regeneration.
In an attempt to improve the CO2 /N ratio, Gray et al. [91]
used a tertiary amine, i.e., 1,8 diazabicyclo-[5.4.0]-unedec-7-ene
(DBU) on PMMA beads, following the rationale that tertiary amines
react with CO2 to form bicarbonate with a more favorable CO2 /N
stoichiometry of 1, compared to the formation of carbamate
(Scheme 1). Since this reaction can take place only in the presence
of humidity, the simulated ue gas used contained 10% CO2 and 2%
H2 O (i.e., 100% RH at 25 C), balance N2 . The sample with the highest loading (ca. 30% DBU) was also the best in terms of adsorption
capacity, with an average of 3.02 mmol/g over four cycles at 25 C
and of 2.34 mmol/g at 65 C, corresponding to a CO2 /N ratio of 0.76
and 0.59, respectively.
Gray et al. [110] explored other supports for PEI impregnation, including polystyrene (Macronet), silicon dioxide (CARiACT)
and PMMA (Diaion). While the polystyrene-supported amines performed poorly, the other two adsorbents were promising, with
PEI-CARiACT and PEI-Diaion adsorbing 2.55 and 2.40 mmol/g,
respectively in a dry stream containing 10% CO2 balance He at
40 C. The enhancement of adsorption capacity due to moisture was
quite evident in this work, with capacity increasing to 3.65 and
3.53 mmol/g for PEI-CARiACT and PEI-Diaion, respectively when
the stream contained 7% water vapor (100% RH).
Carbon-based supports have also been studied. Plaza et al.
[111,112] impregnated PEI on supports produced from sewage
sludge and air-oxidized olive stones (AOS). The resulting adsorbents however, did not offer any particular advantage with an
adsorption capacity of 1.98 mmol/g for PEI-AOS in the presence of

Experimental conditions

Reference

CO2 concentration (%)

T ( C)

0.18
0.27
0.27
0.17
0.25
0.28
0.34
0.38

10
13 (13% H2 O)
15
5
5
10
5
5

75
75
75
75
75
75
75
75

[92]
[93]
[94]
[95]
[96]
[97]
[98]
[99]

0.41
0.12
0.26
0.12
1.28
Proprietary
information
0.59
0.42
0.39
1.70
0.11
0.21

5
100
15
12
15 (2.6% H2 O)
10 (humid)

25
57
70
75
70
25

[100]
[101]
[101]
[103]
[108]
[109]

10 (2% H2 O)
10 (2% H2 O)
10 (2% H2 O)
100
15
10

65
40
40
25
75
30

[89]
[104]
[110]
[111]
[104]
[105]

pure CO2 at 25 C, a value lower than other non-PEI containing activated carbons included in the same work. It should be noted that
the PEI-AOS sample had a much smaller loading of 5 wt% compared
to other amine-impregnated materials.
As summarized in Table 4, the above review of literature
data provides evidence of the great variety of materials produced
by impregnation of amine-containing species on solid supports.
Amine-impregnated adsorbents were produced using a variety of
materials as support, including polymers, zeolites and mesoporous
oxides, although supports with large dp seem to be more appropriate. They exhibited the highest adsorption capacities reported
so far. Despite their diversity, some general properties can be
mentioned about these adsorbents. It was found that high amine
loadings result in enhanced adsorption capacity, however this was
usually accompanied by a decrease in rate of adsorption and CO2 /N
ratio. When the amine content is high, the optimum adsorption
capacity may occur at high temperature, making them inappropriate for applications where lower temperature is required.
Furthermore, because of the weak interactions between the active
phase and the support, amine-impregnated adsorbents can be
unstable. Thus, regeneration of such materials has to be performed
under a strict control of temperature, since low temperatures may
result in incomplete regeneration, but higher temperatures induce
evaporation of the supported amines or degradation of the aminebearing molecules if CO2 is present.
3.2. Grafted materials
To the best of our knowledge, Leal et al. [113] published the rst
study dealing with amine-grafted materials for adsorption of CO2 .
In this pivotal study, silica gel was decorated with amine groups via
grafting of (3-aminopropyl)triethoxysilane (AP) under anhydrous
conditions, resulting in an adsorbent with a capacity of 0.41 mmol/g
(CO2 /N = 0.33) under a stream of pure CO2 at 23 C. Although such
adsorption capacity was very low compared to other benchmark
adsorbents such as zeolites and activated carbons, the relevance of
this work lies in the novel ideas put forth that were subsequently
adopted by other researchers. They used infrared spectroscopy to
substantiate the proposed reaction mechanisms between CO2 and
supported amine groups producing carbamate and bicarbonate in
dry and humid streams, respectively (Scheme 1). Thus, a more

768

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

Fig. 5. Amine loading (left) and adsorption capacity (right) vs. TRI/SiO2 ratio on MCM-41 (TRI-M41C) and PE-MCM-41 (TRI-M41EC) [117].

favorable stoichiometry is expected in the presence of moisture,


which was corroborated by the experimental value of adsorption
capacity, increasing from 0.41 to 0.89 mmol/g in a wet stream (100%
RH). For several years thereafter, no further CO2 adsorption studies
on amine-grafted materials appeared, most likely because of the
lack of interest in this topic. However, with the rapid development
of ordered mesoporous materials, and the increasing awareness
of the greenhouse gas effect, new studies began to appear in the
literature starting in 2003.
In a comparative investigation on propylamine-grafted MCM48 and silica xerogel, Huang et al. [114] provided evidence for the
advantages of ordered mesoporous supports. Using 10% CO2 in N2 ,
a signicantly higher capacity of ca. 1.42 mmol/g was obtained for
AP-MCM-48 at room temperature vs. ca. 0.58 mmol/g for aminegrafted silica xerogel under the same conditions. Since the amine
loading was 2.3 and 1.7 mmol/g for MCM-48 and silica xerogel,
respectively, it is inferred that the signicantly higher capacity
of AP-MCM-48 was accompanied by an improved efciency. Further, using a CO2 -containing stream with 100% RH gave rise to an
adsorption capacity twice as high with a CO2 /N = 1 corresponding
to quantitative transformation of amine groups into ammonium
bicarbonate.
Further evidence for the suitability of periodic mesoporous supports was reported by Knowles et al. [115] using AP-grafted hexagonal mesoporous silica (HMS) and amorphous silica gel. A higher
amine loading of 2.3 mmol/g was obtained vs. only 1.1 mmol/g
for amorphous silica. These loadings mirrored the difference in
surface areas of HMS (1198 m2 /g) compared to amorphous silica
(567 m2 /g). The adsorption capacity at 20 C in the presence of 90%
CO2 /Ar was 1.59 mmol/g for AP-HMS compared to 0.68 mmol/g
for AP-grafted amorphous support. In further work, Knowles
et al. [116] used (3-trimethoxysilylpropyl)diethylenetriamine (TRI)
grafted on HMS and obtained a capacity of 1.34 mmol/g in the
presence of 90% CO2 balance Ar at 20 C. They also found that the
material is thermally stable up to 170 C under pure N2 or mildly
oxygenated environments, a comparatively higher temperature
than amine-impregnated adsorbents.
Sayaris group made signicant contributions to the area of CO2
capture by amine-containing nanoporous materials. They demonstrated the benecial effect of using materials with larger pore
diameter and pore volume than typical MCM-41 silica [90,117,118].

To do so, they used the post-synthesis pore-expansion method


developed earlier [119,120]. Based on as-synthesized MCM-41 as
starting material, they generated PE-MCM-41 with pore size and
pore volume up to 20 nm and 3.5 cm3 /g, vs. typically ca. 34 nm
and ca. 0.71 cm3 /g for regular MCM-41, with hardly any change
in surface area. As shown in Fig. 5, grafting MCM-41 and PE-MCM41 with TRI led to comparable amine loadings, because of similar
surface areas. However, as shown in Fig. 5, using 5% CO2 in N2
at 25 C, the CO2 uptake was ca. 50% higher for TRI-PE-MCM-41
than TRI-MCM-41, at all amine loadings. Moreover, TRI-PE-MCM41 adsorbed CO2 about 30% faster than MCM-41-based material,
showing the importance of pore size and volume.
Another contribution of Sayaris group was the optimization
of the grafting conditions, leading to dramatic improvement of
amine loading and adsorptive properties. Grafting is traditionally
practiced under reux, in dry solvent (typically toluene at 110 C)
with large excess of silane. Harlick and Sayari [90] found that
the optimum grafting conditions of TRI on PE-MCM-41 in toluene
were as follows: T = 85 C; water added: 0.3 mL per gram of support; aminosilane added: 3 mL per gram of support. Under such
conditions, the amine content increased by ca. 30% (i.e., 7.98 vs.
6.11 mmol/g for conventional dry grafting), whereas the adsorption capacity using 5% CO2 /N2 at 25 C increased by ca. 70% from
1.55 mmol/g for conventional dry grafting to 2.65 mmol/g. Thus,
under these CO2 adsorption conditions, the combination of pore
expansion and optimization of grafting conditions improved the
adsorption capacity by close to 300% compared to the adsorbent
produced via anhydrous grafting on conventional MCM-41, in addition to a signicant increase in the rate of adsorption. The advantage
of using amine-functionalized mesoporous materials was further
evidenced when a stream of humid CO2 was used. In the presence of 5% CO2 in N2 with 27% RH, the adsorption capacity for
TRI-PE-MCM-41 increased to 2.94 mmol/g in contrast to a dramatic
decrease observed for 13X, down to 0.09 mmol g1 . This work also
provided evidence of the advantage of amine-grafted adsorbents
in terms of adsorption kinetics, as the CO2 rate of adsorption on
TRI-PE-MCM-41 was found to be higher than 13X zeolite. Later, it
would be corroborated that TRI-PE-MCM-41 is also comparatively
faster than its PEI-impregnated counterpart [91].
Further studies on TRI-PE-MCM-41 [118] demonstrated that
enhanced capacity was not the only advantage of TRI-PE-MCM-

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

769

Fig. 6. Working adsorption capacity of TRI-PE-MCM-41 over various adsorptiondesorption cycles in dry (TRI-70/70-d) and humid (TRI-70/70-h) streams with adsorption
and desorption at 70 C [102].

41, since incorporation of amines signicantly increased the


selectivity toward CO2 over N2 . Using conditions directly related
to ue gas, i.e., 10% CO2 balance N2 at 50 C, Serna-Guerrero
et al. [121] obtained a stable capacity of 1.59 mmol/g over 100
adsorptiondesorption cycles with regeneration under vacuum at
90 C.
To further address the long-term stability of aminefunctionalized adsorbents for CO2 capture, Sayari and Belmabkhout
[102] carried an in-depth investigation using extensive
adsorptiondesorption cycling under different conditions. As
shown in Fig. 6, they found that under dry conditions, the adsorbent will ultimately deactivate even under mild conditions. The
degree of deactivation depended on the nature of the adsorbent
and the adsorptiondesorption conditions. The adsorbent deactivation was clearly associated with the formation of urea groups,
which are stable under the desorption conditions. To prevent the
formation of urea and drastically improve the stability of aminefunctionalized adsorbents, the use of humid streams was proposed.
As illustrated in Fig. 6, the adsorbent underwent more than 700
cycles without any loss in adsorption capacity when the adsorption
and desorption gases contained 7% RH at 70 C. Another interesting
nding was that by treating deactivated AP-grafted PE-MCM-41
in the presence of water vapor at ca. 200 C, it was possible to
hydrolyze the urea groups and fully regenerate the adsorbent.
The advantages of using TRI were also discussed by Hiyoshi
et al. [122] in a thorough comparative analysis of monoamine,
diamine and triamine-bearing molecules grafted on SBA-15. The
higher amine density achieved through the use of TRI resulted in
the best performing adsorbent. The reported adsorption capacity
for TRI-SBA-15 at 60 C and 15% CO2 was of 1.58 mmol/g under dry
conditions and 1.80 mmol/g in a stream containing 60% RH. Furthermore, they reported that TRI-SBA-15 was stable over 50 cycles
of adsorption at 60 C and desorption at 100 C.
Kim et al. [123] made a comparative study between mesoporous
silica (MS) functionalized with molecules containing 13 amine
groups produced by anhydrous grafting and co-condensation.
In general, samples prepared by co-condensation presented
higher amine contents, that reportedly promoted a better amineefciency and adsorption capacity. In line with the observations
mentioned above, the adsorbent with the highest capacity was
TRI-MS, with 1.74 mmol/g for pure CO2 at 25 C, while the most
efcient under the same conditions was the AP-MS sample with
a CO2 /N ratio of 0.43 and a capacity of 1.14 mmol/g. This is consistent with ndings by Serna-Guerrero et al. [124] who obtained
the maximum efciency of CO2 /N = 0.5 using AP-grafted PE-MCM41, whereas the CO2 /N ratio for TRI-PE-MCM-41 never exceeded

0.34 [121]. In addition, Kim et al. [123] compared their aminegrafted materials with PEI-impregnated KIT-6 silica. Although a
higher capacity was obtained on the PEI-containing sample (i.e.,
1.79 mmol/g), its CO2 /N efciency at room temperature was only
0.1.
The use of diamine-bearing molecules was investigated under
the hypothesis that the occurrence of two amine groups in
close proximity will lead to enhanced formation of carbamate,
thus higher CO2 /N efciency. Knofel et al. [125] grafted N-[3(trimethoxysilyl)propyl] ethylenediamine (EDA) on SBA-16 silica.
Although this work was mainly focused on CO2 adsorption at high
pressure, it clearly showed that incorporation of amine groups
resulted in an improved capacity at CO2 partial pressures below
1 bar. The reported capacity for pure CO2 at 1 bar was of 1.4 mmol/g
at 27 C for the best performing EDA-SBA-16. It was observed however, that at high pressure (ca. 4 bar or more), the non-aminated
samples exhibited higher adsorption capacity. A possible explanation was that physical adsorption predominates at high pressure
and so, the higher pore volume of the unmodied support offers a
comparative advantage in terms of adsorption capacity.
In recent years, efforts to further improve the grafting process have been pursued, with the aim of improving the efciency
and capacity of aminated silicas. Wang et al. [126] incorporated
AP-functionality by simultaneous extraction of structure directing agent and grafting on as-synthesized SBA-15. The adsorbent
obtained by the proposed approach outperformed a sample synthesized using the typical grafting procedure on calcined SBA-15.
The sample using as-synthesized support produced a material with
an adsorption capacity of ca. 0.45 mmol/g at 65 C at a CO2 partial
pressure of 0.1 bar, representing a CO2 /N efciency of 0.44, close to
the stoichiometric ratio of 0.5. It was suggested that, unlike calcination, the extraction of surfactant template performed with ethanol
preserved the surface silanol groups, which translated into a better
distribution of surface amines with a subsequent improvement of
adsorption capacity.
The drawback of surface silanol groups removal during calcination of the support was also addressed by Wei et al. [127]. They
proposed rehydrating SBA-15 by soaking it in water at 97 C, before
grafting with EDA. The obtained material had an amine content
of 3.06 mmol/g, and a capacity of 0.73 mmol g1 for 0.15 bar CO2
at 60 C. A similar material prepared using non-hydrated SBA-15
had an amine loading of 2.59 mmol/g and an adsorption capacity of
0.59 mmol/g.
Zelenak et al. reported on the effect of pore size [128] and
the basicity of the functional groups [129] on the performance of
amine-functionalized adsorbents for CO2 capture. It was suggested

770

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

Fig. 7. Schematic representation of the synthesis of hyperbranched aminosilica according to Jones et al. [136].

that large pore sizes are associated with an efcient use of amine
groups. Indeed, when grafting AP on MCM-41 with a pore size of
3.3 nm, a high amine content was obtained (i.e., 3 mmol/g), but the
adsorption capacity was only 0.57 mmol/g for 10% CO2 at 25 C.
In contrast, a capacity of 1.54 mmol/g was obtained when SBA-15
with a pore size of 7.1 nm was used as support, despite a slightly
lower amine loading of 2.7 mmol/g. The lower efciency of MCM41-based material, however, may not be only a result of a difference
in pore sizes. The amine surface density reported for AP-MCM41 was lower, having only 1.1 amine group per nm2 compared to
2.4 amines per nm2 for AP-SBA-15. Since admittedly two amine
molecules in close proximity are required for reaction with CO2 ,
this would be a disadvantage for the lower amine density sample.
Some interesting studies were devoted to the effect of the supports on the adsorbent performance. Knofel et al. [130] compared
AP-grafted mesoporous silica (MS) and mesoporous titania (MT).
The highest adsorption capacity of ca. 0.24 mmol/g for 10% CO2
at 30 C was obtained with AP-MS. This is a low capacity compared to other materials reported in the literature, but the main
nding of this work was that the properties of the support may
inuence the behavior of the functionalized adsorbent. While no
interactions were detected between CO2 and the silica support,
interactions occurred in the presence of MT. This is reected in
a higher capacity when expressed in terms of surface area, i.e.,
1 mol/m2 and 0.6 mol/m2 for AP-MT and AP-MS, respectively.
Another approach explored by Lu et al. [131] was the use of particles
with dened geometry, by grafting EDA on mesoporous spherical
particles. The adsorption capacity at 60 C in the presence of 10%
CO2 in air was of 0.73 mmol/g. These adsorbents showed a remarkable stability when regenerated using a TSA procedure at 120 C
or under VSA, although their adsorption capacity decreased in the
rst VSA cycle. In addition, it was suggested that the combination of heat and vacuum resulted in an improvement in desorption
rate.
Looking for an inexpensive source of silica, Bhagiyalakshmi
et al. [132] grafted tris(2-aminoethyl)amine (TREN) and TEPA onto
chloropropyl-modied mesoporous supports produced from rice
husk. The highest capacities were obtained with TREN-grafted
MCM-48 with values of ca. 1.59 and 1.36 mmol/g at 25 and 50 C,
respectively in the presence of pure CO2 .
The only contribution dealing with amine-grafted zeolites used
ITQ-6 [133], which offers attractive characteristics such as high concentration of surface silanol groups and a pore size in the nanometer
range. The most promising adsorbent had a capacity of 0.67 mmol/g
for 12% CO2 at 20 C.

3.3. Hyperbranched aminosilicas


A different method of functionalization, introduced recently,
consisted in iterative building of amine-containing dendrimers
inside the porous supports. Liang et al. [134,135] produced highly
branched dendrimers by step-wise reaction between diisopropylethylamine and cyanuric chloride inside the pores of SBA-15 [134]
or mesocellular siliceous foams [135]. The optimum adsorbent produced with this approach was obtained after 3 reaction steps, with
a capacity of ca. 1 mmol/g for 90% CO2 in Ar at 20 C. However, as
a larger number of reaction steps were performed to obtain higher
generation dendrimers, the adsorbent lost its structural properties,
allegedly due to space limitations, which negatively impacted the
adsorptive properties.
Jones group [136,137] proposed an innovative amine polymerization approach inside SBA-15 channels with promising results.
In this case, aziridine was polymerized by ring opening inside the
pores of SBA-15 producing a covalently tethered hyperbranched
aminosilica, as represented in Fig. 7. This material exhibited a
capacity of 3.11 mmol/g under a ow of water saturated 10% CO2 /Ar
at 25 C. The CO2 /N efciency was as high as 0.44 at room temperature, close to the theoretical value of 0.5. With respect to its
performance at 75 C, and 10% CO2 /Ar, the hyperbranched-SBA15 was stable, presenting an average adsorption capacity of ca.
1.98 mmol/g over 12 cycles with regeneration at 130 C. In a later
contribution [138], it was shown that higher loading of hyperbranched amines afforded a better capacity. The best reported
adsorbent had an amine loading of 9.78 mmol/g and adsorbed ca.
4 mmol/g at 10% CO2 /N2 at 75 C in the presence of humidity.
As summarized in Table 5, Similarly to amine-impregnated
adsorbents, the covalently bonded aminated adsorbents span
materials with a wide variety of characteristics and performances.
However, a number of common advantages and limitations of
amine-grafted materials can be outlined. Only supports that exhibit
surface hydroxyl groups can be used to produce amine-grafted
materials. It was observed that high amine loading is a result
of high surface area and availability of surface silanol groups,
but the efcient use of functional groups is observed mainly
in supports with large pores. While the equilibrium adsorption
capacities are certainly not as high as those reported with some
amine-impregnated adsorbents, properly designed amine-grafted
materials do not exhibit the strong diffusion limitations observed
in impregnated adsorbents. Therefore high adsorption rates are not
restricted upon operating at high temperature. A particular advantage offered by amine-grafted adsorbents is their high stability over

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

771

Table 5
Literature data on CO2 adsorption capacity of amine-grafted adsorbents.
Support

Silica gel
MCM-48
HMS
HMS
PE-MCM-41
SBA-15
MS
SBA-16
SBA-15
SBA-16
SBA-15
SBA-12
MS
MSP
MCM-48
ITQ-6
SBA-15
SBA-15

Amine

AP
AP
AP
TRI
TRI
TRI
TRI (co-cond)
EDA
AP
EDA
AP
AP
AP
EDA
TREN
AP
Amine-dendrimers
Aziridine polymer

Capacity
(mmol/g)

Amine
loading
(mmol/g)
0.89
2.3
1.59
1.34
1.59
1.80
1.74
1.4
0.45
0.727
1.54
1.04
0.24
0.73
1.36
0.67
1
4

CO2 /N

1.26
2.3
2.29
4.57
7.9
5.80
5.18
0.76
2.56
3.06
2.72
2.13
1.6
0.99
4
1.26
1.25
9.78

hundreds, most likely thousands, of adsorptiondesorption cycles


[102].
Although this review focused on CO2 capture, as mentioned in
the Introduction, there are other impurities in ue gas. Of particular
concern with respect to amine-functional materials is the presence of SO2 , as it was found that it affects negatively their cyclic
performance. For example, it was recently reported that after exposure to SO2 , the working adsorption capacity of TRI-PE-MCM-41
decreased from 1.57 mmol/g to 0.89 mmol/g for a mixture of 10%
CO2 /N2 at 50 C [139]. Supported by gravimetric measurements and
FTIR spectroscopy, it was proposed that SO2 reacts irreversibly with
the primary amines of the triamine functional molecules. Consequently, it might be necessary to engineer processes to remove
SO2 prior to CO2 capture to preventing its contact with aminefunctionalized adsorbents.

4. Conclusions
Major advances have been achieved toward the development of
a CO2 capture technology based on adsorption. Physical adsorbents
such as zeolites, carbon-based materials and MOFs were found to be
suitable, mostly at low temperature and high pressure. These adsorbents, however, often adsorb water vapor preferentially over CO2 ,
and their CO2 adsorption capacity at low pressure is not sufciently
high. Although these materials may provide elegant solutions for
CO2 sequestration and storage, they are not particularly suitable for
post-combustion gas treatment. Nevertheless, a continuous effort
is being deployed to circumvent such drawbacks. The strategies
being used include surface modication to enhance the interactions
with CO2 , thus increasing the adsorption capacity at low pressure.
Another route is to design completely new materials such as ZMOFs
and COFs with increased tolerance to moisture in the gas feed, thus
improved CO2 selectivity.
Likewise, tremendous progress has been achieved in the development of novel chemical adsorbents such as amine-modied
materials with large surface area. By optimizing the synthesis conditions and using supports with adequate structural properties, it
was possible to develop materials with superior CO2 adsorptive
properties, particularly suitable for ue gas treatment. Typically,
these materials exhibit large CO2 adsorption capacity even at low
pressure, high rate of adsorption and desorption, and excellent tolerance to moisture in the feed. Furthermore, contrary to physical
adsorbents, the selectivity of amine-functionalized materials is not
signicantly affected by temperature, at least within the range of

0.71
1
0.69
0.29
0.20
0.31
0.34
1.84
0.18
0.24
0.57
0.49
0.15
0.73
0.34
0.53
0.40
0.41

Experimental conditions

Reference

CO2 concentration

T ( C)

100% (100% RH)


10% (100% RH)
90%
90%
10%
15% (humid)
100%
100%
10%
15%
10%
10%
10%
10%
100%
12%
90%
10% (humid)

50
25
20
20
50
60
25
27
65
60
25
25
30
60
50
20
20
75

[113]
[114]
[115]
[116]
[121]
[122]
[123]
[125]
[126]
[127]
[128]
[129]
[130]
[131]
[132]
[133]
[134]
[138]

interest for ue gas treatment. While the stability of this kind of


adsorbents has been questioned, it was recently demonstrated that
their stability may be dramatically enhanced during thousands of
adsorptiondesorption cycles, provided that the feed and purge
gases contains moisture. The role of moisture is to prevent the
formation of urea linkages, which is the main source of material
deactivation.
This review clearly showed a steady improvement in the CO2
adsorptive properties of novel materials. The course followed so
far has resulted in major achievements that may well pave the way
for an alternative CO2 capture technology in the near future.
References
[1] D.J. Hofmann, J.H. Butler, P.P. Tans, A new look at atmospheric carbon dioxide,
Atmos. Environ. 43 (2009) 20842086.
[2] R. Monastersky, A burden beyond bearing, Nature 458 (2009) 10911094.
[3] A.J. Yamakasi, An overview of CO2 mitigation options for global warming
emphasizing CO2 sequestration options, J. Chem. Eng. Jpn. 36 (2003) 361375.
[4] D. Aaron, C. Tsouris, Separation of CO2 from ue gas: a review, Sep. Sci. Technol. 40 (2005) 321348.
[5] S.M. Benson, F.M. Orr Jr., Carbon dioxide capture and storage, MRS Bull. 33
(2008) 303305.
[6] K.M.K. Yu, I. Curcic, J. Gabriel, S.C.E. Tsang, Recent advances in CO2 capture
and utilization, ChemSusChem 1 (2008) 893899.
[7] H. Audus, Greenehouse gas mitigation technology: an overview of the CO2
capture and sequestration studies and further activities of the IEA greenhouse
gas R&D programme, Energy 22 (1997) 217221.
[8] A. Sjostrom, H. Krutka, Evaluation of solid sorbents as a retrot technology
for CO2 capture, Fuel 89 (2010) 12981306.
[9] M.T. Ho, G.W. Allinson, D.E. Wiley, Reducing the cost of CO2 capture from
ue gases using pressure swing adsorption, Ind. Eng. Chem. Res. 47 (2008)
48834890.
[10] S. Majumdar, A. Sengupta, J.S. Cha, K.K. Sirkarst, Simultaneous SO2 /NOx separation from ue gas in a contained liquid membrane permeator, Ind. Eng.
Chem. Res. 33 (1994) 667675.
[11] T.C. Drage, A. Arenillas, K.M. Smith, C.E. Snape, Thermal stability of
polyethylenimine based carbon dioxide adsorbents and its inuence on
selection of regeneration strategies, Micropor. Mesopor. Mater. 116 (2008)
504512.
[12] S. Choi, J.H. Drese, C.W. Jones, Adsorbent materials for carbon dioxide capture
from large anthropogenic point sources, ChemSusChem 2 (2009) 796854.
[13] K.B. Lee, M.G. Beaver, H.S. Caram, S. Sircar, Reversible chemisorbents for carbon dioxide and their potential applications, Ind. Eng. Chem. Res. 47 (2008)
80488062.
[14] C.J. Major, B.J. Sollami, K. Kammermeyer, Carbon dioxide removal from air by
adsorbents, Ind. Eng. Chem. Process Des. Dev. 4 (1965) 327333.
[15] D.M. Ruthven, Principles of Adsorption and Adsorption Processes, WileyInterscience, New York, 1984.
[16] R.T. Yang, Gas Separation by Adsorption Processes, Imperial College Press,
Boston, 1997.
[17] D. Lozano-Castello, D. Cazorla-Amoros, A. Linares-Solano, Powdered activated
carbons and activated carbon bers for methane storage: a comparative study,
Energy Fuels 16 (2002) 13211328.

772

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

[18] C.G. Coe, Structural effects on the adsorptive properties of molecular sieves
for air separation, in: T.J. Pinnavaia, M.F. Thorpe (Eds.), Access in Nanoporous
Materials, Plenum Press, New York, 1995, pp. 213229.
[19] S. Hayashi, M. Kawai, T. Kaneko, Dynamics of high purity oxygen PSA, Gas Sep.
Purif. 10 (1996) 1923.
[20] V.S. Filipe, C.A. Lopes, A.M. Grande, J.M. Ribeiro, E. Loureiro, V. Oikonomopoulos, A.E. Nikolakis, Rodrigues, Adsorption of H2 , CO2 , CH4 , CO, N2 and H2 O in
activated carbon and zeolite for hydrogen production, Sep. Sci. Technol. 44
(2009) 10451073.
[21] B.K. Na, I.K. Koo, H.M. Eum, H. Lee, H.K. Song, CO2 recovery from ue gas by
PSA process using activated carbon, Korean J. Chem. Eng. 18 (2001) 220227.
[22] M. Cinke, J. Li, C.W. Baushlicher, A. Ricca, M. Meyyappan, CO2 adsorption in single walled carbon nanotubes, Chem. Phys. Lett. 376 (2003)
761766.
[23] F. Su, C. Lu, W. Chen, H. Bai, J.F. Hwang, Capture of CO2 from ue gas via
multiwalled carbon nanotubes, Sci. Total Environ. 407 (2009) 30173023.
[24] S. Himeno, T. Komatsu, S. Fujita, High-pressure adsorption equilibria of
methane and carbon dioxide on several activated carbons, J. Chem. Eng. Data
50 (2005) 369376.
[25] F. Dreisbach, R. Staudt, J.U. Keller, High pressure adsorption data of methane,
nitrogen, carbon dioxide and their ternary mixture on activated carbon,
Adsorption 5 (2005) 215227.
[26] Y. Kurniwan, S.K. Bathia, V. Rudolph, Simulation of binary mixture adsorption
of methane and CO2 at supercritical conditions in carbons, AIChE J. 52 (2006)
957967.
[27] D.D. Do, K.A. Wang, New model for the description of adsorption kinetics in
heterogeneous activated carbon, Carbon 36 (1998) 15391554.
[28] F. Dreisbach, R. Staudt, J.U. Keller, Experimental investigation of the kinetics
of adsorption of pure gases and binary gas mixtures on activated carbon,
in: F. Meunier (Ed.), Proceedings of Fundamental of Adsorption 6, Elsevier,
Amsterdam, 1998, pp. 12191224.
[29] Y. Belmabkhout, R. Serna-Guerrero, A. Sayari, Adsorption of CO2 from dry
gases on MCM-41 silica at ambient temperature and high pressure. 1: pure
CO2 adsorption, Chem. Eng. Sci. 64 (2009) 37213728.
[30] B.K. Na, H. Lee, I.K. Koo, H.K. Song, Effect of rinse and recycle methods on the
pressure swing adsorption process to recover CO2 from power plant ue gas
using activated carbon, Ind. Eng. Chem. Res. 41 (2002) 54985503.
[31] K.T. Chue, J.N. Kim, Y.J. Yoo, S.H. Cho, Comparison of activated carbon and
zeolite 13X for CO2 recovery from ue gas by pressure swing adsorption, Ind.
Eng. Chem. Res. 34 (1995) 591598.
[32] W. Zhou, X. Bai, E. Wang, S. Xie, Synthesis, structure, and properties of singlewalled carbon nanotubes, Adv. Mater. 21 (2009) 45654583.
[33] L. Huang, L. Zhang, Q. Shao, L. Lu, X. Lu, S. Jiang, W. Shen, Simulations of
binary mixture adsorption of carbon dioxide and methane in carbon nanotubes: temperature, pressure, and pore size effects, J. Phys. Chem. C 111
(2007) 1191211920.
[34] Ch. Baerlocher, L.B. McCusker, D.H. Olson, Atlas of Zeolite Structure TypesEDN
6th revised ed., Elsevier, Amsterdam, 2007.
[35] R.V. Siriwardane, M.S. Shen, E.P. Fisher, J.A. Poston, Adsorption of CO2 on
molecular sieves and activated carbon, Energy Fuels 15 (2001) 279284.
[36] A. Goj, D.S. Sholl, E.D. Akten, D.J. Kohen, Atomistic simulations of CO2 and N2
adsorption in silica zeolites: the impact of pore size and shape, Phys. Chem.
B 106 (2002) 83678375.
[37] P.J.E. Harlick, F.H. Tezel, An experimental adsorbent screening study for CO2
removal from N2 , Micropor. Mesopor. Mater. 76 (2004) 7179.
[38] C. Lu, H. Bai, B. Wu, F. Su, J.F. Hwang, Comparative study of CO2 capture
by carbon nanotubes, activated carbon, and zeolites, Energy Fuels 22 (2008)
30503056.
[39] N. Konduru, P. Lindner, N.M. Assaf-Annid, Curbing the greenhouse effect by
carbon dioxide adsorption with zeolite 13X, AIChE J. 53 (2007) 31373143.
[40] J. Zhang, P.A. Webley, P. Xiao, Effect of process paramaters on power requirements of vacuum swing adsorption technology for CO2 capture from ue gas,
Energy Convers. Manage. 49 (2008) 346356.
[41] G. Li, P. Xiao, P.A. Webley, J. Zhang, R. Singh, M. Marshal, Capture of CO2 from
high humidity ue gas by vacuum swing adsorption with 13X, Adsorption 14
(2008) 415422.
[42] J. Zhang, P.A. Webley, Cycle development and design for CO2 capture from ue
gas by vacuum swing adsorption, Environ. Sci. Technol. 42 (2008) 563569.
[43] R.M. Barrer, R.M. Gibbons, Zeolitic carbon dioxide: energetics and equilibrium
in relation to exchangeable cations in faujasite, J. Chem. Soc., Faraday Trans.
61 (1965) 948961.
[44] K.S. Walton, M.B. Abney, M.D. LeVan, CO2 adsorption in Y and X zeolites modied by alkali metal cation exchange, Micropor. Mesopor. Mater. 91 (2006)
7884.
[45] G. Maurin, P.L. Llewellyn, R.G. Bell, Adsorption mechanism of carbon dioxide
in Faujasites: Grand Canonical Monte Carlo simulations and microcalorimetry
measurements, J. Phys. Chem. B 109 (2005) 1608416091.
[46] G.D. Pirngruber, P. Raybaud, Y. Belmabkhout, J. Cejka, A. Zukal, The role of
extraframework cations in the adsorption of CO2 on faujasite Y, Phys. Chem.
Chem. Phys. 12 (2010) 1353413546.
[47] E.D. Akten, R. Siriwardane, D.S. Sholl, Monte Carlo simulation of single- and
binary-component adsorption of CO2 , N2 and H2 in zeolite Na-4A, Energy
Fuels 17 (2003) 977983.
[48] S. Cavenati, C.A. Grande, A.E. Rodrigues, Adsorption equilibrium of methane,
carbon dioxide, and nitrogen on zeolites 13X at high pressures, J. Chem. Eng.
Data 49 (2004) 10951101.

[49] G. Maurin, Y. Belmabkhout, G. Pirngruber, L. Gaberova, P.L. Llewellyn, CO2


adsorption in LiY and NaY at high temperature: molecular simulations compared to experiments, Adsorption 13 (2007) 453460.
[50] J.A. Dunne, R. Mariwala, M. Rao, S. Sircar, R.J. Gorte, A.L. Myers, Calorimetric
heats of adsorption and adsorption isotherms. 1. O2 , N2 , Ar, CO2 , CH4 , C2 H6 ,
and SF6 on silicalite, Langmuir 12 (1996) 58885895.
[51] P.J.E Harlick, F.H. Tezel, Adsorption of carbon dioxide, methane and nitrogen:
pure and binary mixture adsorption by ZSM-5 with SiO2 /Al2 O3 ratio of 30,
Sep. Purif. Technol. 37 (2002) 3360.
[52] A. Zukal, J. Pawlesa, J. Cejka, Isosteric heats of adsorption of carbon dioxide
on zeolite MCM-22 modied by alkali metal cations, Adsorption 15 (2009)
264270.
[53] J. Merel, M. Clausse, F. Meunier, Experimental investigation on CO2 postcombustion capture by indirect thermal swing adsorption using 13X and 5A
zeolites, Ind. Eng. Chem. Res. 47 (2008) 209215.
[54] F. Brandani, D. Ruthven, The effect of water on the adsorption of CO2 and C3 H8
on Type X zeolites, Ind. Eng. Chem. Res. 43 (2004) 83398344.
[55] A. Hirotani, R. Mizukami, H. Miura, T. Takaba, A. Miya, A. Fahmi, A. Stirling, M.
Kubo, A. Miymoto, Grand Canonical Monte Carlo of the adsorption of CO2 on
silicalite and NaZSM-5, Appl. Surf. Sci. 120 (1997) 8184.
[56] J.M. Leyssale, G.K. Papadopoulos, D.N. Theodoru, Sorption thermodynamics
of CO2 , CH4 , and their mixtures in the ITQ-1 zeolite as revealed by molecular
simulations, J. Phys. Chem. B 110 (2006) 2274222753.
[57] M. Castro, S.J. Warrender, P. Wright, D.C. Apperley, Y. Belmabkhout, G. Pirngruber, H.K. Min, M.B. Park, S.B. Hong, Silicoaluminophosphate molecular
sieves STA-7 and STA-14 and their structure-dependent catalytic performance in the conversion of methanol to olens, J. Phys. Chem. C 113 (2009)
1573115741.
[58] I. Deroche, L. Gaberova, G. Maurin, P.L. Llewellyn, M. Castro, P. Wright,
Adsorption of carbon dioxide in SAPO STA-7 and AlPO-18: Grand Canonical
Monte Carlo simulations and microcalorimetry measurements, Adsorption
14 (2008) 207213.
[59] D. Bonenfant, M. Kharoune, P. Niquette, M. Mimeault, R. Hausler, Advances in
principal factor inuencing carbon dioxide adsorption in zeolites, Sci. Technol.
Adv. Mater. 9 (2008) 013001013007.
[60] O.M. Yaghi, M. OKeeffe, M.W. Ockwing, M. Chae, M. Eddaoudi, J. Kim, Reticular
synthesis and the design of new materials, Nature 423 (2003) 705714.
[61] G. Ferey, Hybrid porous solids: past, present, future, Chem. Soc. Rev. 37 (2008)
191214.
[62] S. Bourrelly, P.L. Llewellyn, C.V. Serre, F. Millange, T. Loiseau, G. Ferey, Different adsorption behaviours of methane and carbon dioxide in the isotypic
nanoporous metal terephthalates MIL-53 and MIL-47, J. Am. Chem. Soc. 127
(2005) 1351913521.
[63] P.L. Llewellyn, S. Bourrrelly, C. Serre, Y. Filinchuk, G. Ferey, How hydration drastically improves adsorption selectivity for CO2 over CH4 in the
exible chromium terephthalate MIL-53, Angew. Chem. Int. Ed. 45 (2006)
77517754.
[64] A.R. Millward, O.M. Yaghi, Metal-organic frameworks with exceptionally high
capacity for storage of carbon dioxide at room temperature, J. Am. Chem. Soc.
127 (2005) 1799817999.
[65] P.L. Llewellyn, S. Bourrelly, C. Serre, A. Vimont, M. Daturi, L. Hamon, G. De
Weireld, J.S. Chang, D.Y. Hong, Y.K. Hwang, S.H. Jhung, G. Ferey, High uptakes
of CO2 and CH4 in mesoporous metal-organic frameworks MIL-100 and MIL101, Langmuir 24 (2008) 72457754.
[66] Q. Yang, C. Zhong, J.F. Chen., Computational study of CO2 storage in metalorganic frameworks, J. Phys. Chem. C 112 (2008) 15621569.
[67] V. Finsy, L. Ma, L. Alaert, D.E. De Vos, C.V. Baron, J.F.M. Denayer, Separation
of CO2 mixtures with the MIL-53(Al) metal-organic framework, Micropor.
Mesopor. Mater. 120 (2009) 221227.
[68] D. Britt, H. Furukawa, H.B. Wang, T.G. Glover, O.M. Yaghi, Highly efcient
separation of carbon dioxide by metal-organic framework replete with open
metal sites, PNAS 106 (2009) 2063720640.
[69] R. Banerjee, A. Phan, P. Wang, C. Knobler, H. Furukawa, M. OKeeffe, O.M.
Yaghi, High throughput synthesis of zeolitic imidazolate frameworks and
application to CO2 capture, Science 319 (2008) 939943.
[70] R. Banerjee, H. Furukawa, D. Britt, C. Kobler, M. OKeeffe, O.M. Yaghi, Control
of pore size and functionality in isoreticular zeolitic imidazolate frameworks
and their carbon dioxide selective capture properties, J. Am. Chem. Soc. 131
(2009) 38753877.
[71] B. Wang, A. Cote, H. Furukawa, M. OKeeffe, O.M. Yaghi, Colossal cages in
zeolitic imidazolate framework as selective carbon dioxide reservoirs, Nature
453 (2008) 207211.
[72] S.R. Caskey, A.G. Wong-Foy, A.J. Matzger, Dramatic tuning of carbon dioxide uptake via metal substitution in a coordination polymer with cylindrical
pores, J. Am. Chem. Soc. 130 (2008) 1087010871.
[73] A.O. Yazaydin, R.Q. Snurr, T.H. Park, K. Koh, J. Liu, M.D. LeVan, A.I. Benin,
P. Jakubczak, M. Lanuza, D.B. Galloway, J.J. Low, R.R. Willis, Screening of
metal-organic frameworks for carbon dioxide capture from ue gas using a
combined experimental and modeling approach, J. Am. Chem. Soc. 131 (2009)
1819818199.
[74] Y.S. Bae, O.K. Farha, A.M. Spokoyny, C.A. Mirkin, S. Punatahnam, J.T. Hupp,
Q.R. Snurr, Carborane-based metal-organic frameworks as highly selective
sorbents for CO2 over methane, Chem. Commun. (2008) 41354137.
[75] L. Bastin, P.S. Barcia, E.J. Hurtado, J.A.C. Silva, A.E. Rodrigues, B. Chen, A Microporous metal-organic framework for separation of CO2 /N2 and CO2 /CH4 by
xed-bed adsorption, J. Phys. Chem. C 112 (2008) 15751581.

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774


[76] B. Liu, B. Smit, Comparative Molecular simulation study of CO2 /N2 and CH4 /N2
separation in zeolites and metal-organic frameworks, Langmuir 25 (2008)
59185926.
[77] A.I. Skoulidas, D.S. Sholl, Self-diffusion and transport of light gases in
metal-organic framework materials assessed using molecular dynamics simulations, J. Phys. Chem. B 109 (2005) 1576015768.
[78] P.S. Barcia, L. Bastin, E.J. Hurtado, J.A.C. Silva, A.E. Rodrigues, B. Chen, Single and
multicomponent sorption of CO2 , CH4 and N2 in a microporous metal-organic
framework, Sep. Sci. Technol. 43 (2008) 34943521.
[79] Q. Yang, C. Xue, C. Zhong, J.F. Chen, Molecular simulation of separation of
CO2 from ue gas in Cu-BTC metal-organic framework, AIChE J. 53 (2007)
28322840.
[80] A. Phan, C.J. Doonan, F.J. Uribe-Romo, C. Knobler, M. OKeeffe, O.M. Yaghi,
Synthesis, structure, and carbon dioxide of zeolitic imidazolate frameworks,
Acc, Chem. Res. 43 (2010) 5867.
[81] Y. Li, R.T. Yang, Gas adsorption and storage in metal-organic framework MOF177, Langmuir 23 (2007) 1293712944.
[82] D.F. Bahr, J.A. Reid, W.M. Mook, C.A. Bauer, R. Stumpf, A.J. Skulan, N.R. Moody,
B.A. Simmons, M.M. Shindel, M.D. Allendorf, Mechanical properties of cubic
zinc carboxylates IRMOF-1 metal-organic framework crystals, Phys. Rev. B 76
(2007) 184106/1184106/7.
[83] J.A. Greathouse, M.D. Alendorf, The interaction of water with MOF-5 simulated by molecular dynamics, J. Am. Chem. Soc. 128 (2006) 1067810679.
[84] K.S. Park, Z. Ni, A.P. Ct, R. Choi, R. Huang, F.J. Unribe-Romo, H.K. Chae, M.
OKeeffe, O.M. Yaghi, Exceptional chemical and thermal stability of zeolitic
imidazolate frameworks, PNAS 103 (2006) 1018610191.
[85] A. Demessence, D.M. DAlessandro, M.L. Foo, J.R. Long, Strong CO2 binding in
a water-stable, triazolate-bridged metalorganic framework functionalized
with ethylenediamine, J. Am. Chem. Soc. 131 (2009) 87848786.
[86] H. Furukawa, O.M. Yaghi, Storage of hydrogen, methane, and carbon dioxide
in highly porous covalent organic frameworks for clean energy applications,
J. Am. Chem. Soc. 131 (2009) 88758883.
[87] R. Babarao, J. Jiang, Molecular screening of metal-organic frameworks for CO2
storage, Langmuir 24 (2008) 62706278.
[88] E.F. da Silva, H.F. Svendsen, Computational chemistry study of reactions,
equilibrium and kinetics of chemical CO2 absorption, Int. J. Greenhouse Gas
Control 1 (2007) 151157.
[89] M.L. Gray, K.J. Champagne, D. Fauth, J.P. Baltrus, H. Pennline, Performance of
immobilized tertiary amine solid sorbents for the capture of carbon dioxide,
Int. J. Greenhouse Gas Control 2 (2008) 38.
[90] P.J.E. Harlick, A. Sayari, Applications of pore-expanded mesoporous silicas.
5. Triamine grafted material with exceptional CO2 dynamic and equilibrium
adsorption performance, Ind. Eng. Chem. Res. 46 (2007) 446458.
[91] R. Serna-Guerrero, A. Sayari, Modeling adsorption of CO2 on aminefunctionalized mesoporous silica. 2: kinetics and breakthrough curves, Chem.
Eng. J. 161 (2001) 182190.
[92] X. Xu, C. Song, J.M. Andersen, B.G. Miller, A.W. Scaroni, Novel
polyethylenimine-modied mesoporous molecular sieve of MCM-41
type as high-capacity adsorbent for CO2 capture, Energy Fuels 16 (2002)
14631469.
[93] X. Xu, C. Song, B.G. Miller, A.W. Scaroni, Inuence of moisture on
CO2 separation from gas mixture by a nanoporous adsorbent based on
polyethylenimine-modied molecular sieve MCM-41, Ind. Eng. Chem. Res.
44 (2005) 81138119.
[94] X. Ma, X. Wang, C. Song, Molecular basket sorbent for separation of CO2 and
H2 S from various gas streams, J. Am. Chem. Soc. 131 (2009) 57775783.
[95] W.J. Son, J.S. Choi, W.S. Ahn, Adsorptive removal of carbon dioxide using
polyethylenimine-loaded mesoporous materials, Micropor. Mesopor. Mater.
113 (2008) 3140.
[96] C. Chen, S.T. Yang, W.S. Ahn, R. Ryoo, Amine-impregnated silica monolith with
a hierarchical pore structure: enhancement of CO2 capture capacity, Chem.
Commun. 24 (2009) 36273629.
[97] M.B. Yue, Y. Chun, Y. Cao, X. Dong, J.H. Zhu, CO2 capture by as-prepared SBA-15
with an occluded organic template, Adv. Funct. Mater. 16 (2006) 17171722.
[98] M.B. Yue, L.B. Sun, Y. Cao, Y. Wang, Z.J. Wang, J.H. Zhu, Efcient CO2 capturer
derived from as-synthesized MCM-41 modied with amine, Chem. Eur. J. 14
(2008) 34423451.
[99] M.B. Yue, L.B. Sun, Y. Cao, Z.J. Wang, Y. Wang, Q. Yu, J.H. Zhu, Promoting the
CO2 adsorption in the amine-containing SBA-15 by hydroxyl group, Micropor.
Mesopor. Mater. 114 (2008) 7481.
[100] R.S. Franchi, P.J.E. Harlick, A. Sayari, Applications of pore-expanded mesoporous silica. 2. Development of a high-capacity, water-tolerant adsorbent
for CO2 , Ind. Eng. Chem. Res. 44 (2005) 80078013.
[101] M.G. Plaza, C. Pevida, B. Arias, J. Fermoso, A. Arenillas, F. Rubiera, J.J. Pis,
Application of thermogravimetric analysis to the evaluation of aminated solid
sorbents for CO2 capture, J. Therm. Anal. Calorim. 92 (2008) 601606.
[102] A. Sayari, Y. Belmabkhout, Stabilization of amine-containing CO2 adsorbents:
dramatic effect of water vapor, J. Am. Chem. Soc. 132 (2010) 63126314.
[103] S. Dasgupta, A. Nanoti, P. Gupta, D. Jena, A.N. Goswami, M.O. Garg, Carbon
dioxide removal with mesoporous adsorbents in a single column pressure
swing adsorber, Sep. Sci. Technol. 44 (2009) 39733983.
[104] P.D. Jadhav, R.V. Chatti, R.B. Biniwale, N.K. Labhsetwar, S. Devotta, S.S. Rayalu, Monoethanol amine modied zeolite 13X for CO2 adsorption at different
temperatures, Energy Fuels 21 (2007) 35553559.
[105] J.C. Fisher, J. Tanthana, S.S.C. Chuang, Oxide-supported tetraethylenepentamine for CO2 capture, AIChE J. 28 (2009) 589598.

773

[106] S. Satyapal, T. Filburn, J. Trela, J. Strange, Performance and properties of a solid


amine sorbent for carbon dioxide removal in space life support applications,
Energy Fuels 15 (2001) 250255.
[107] T. Filburn, J.J. Helble, R.A. Weiss, Development of supported ethanolamines
and modied ethanolamines for CO2 capture, Ind. Eng. Chem. Res. 44 (2005)
15421546.
[108] S. Lee, T.P. Filburn, M. Gray, J.W. Park, H.J. Song, Screening test of solid amine
sorbents for CO2 capture, Ind. Eng. Chem. Res. 47 (2008) 74197423.
[109] M.L. Gray, Y. Soong, K.J. Champagne, H. Penniline, J.P. Baltrus, R.W. Stevens,
R. Khatri, S.S.C. Chuang, T. Filburn, Improved immobilized carbon dioxide
capture sorbents, Fuel Process. Technol. 86 (2005) 14491455.
[110] M.L. Gray, J.S. Hoffman, D.C. Hreha, D.J. Fauth, S.W. Hedges, K.J. Champagne,
H.W. Pennline, Parametric study of solid amine sorbents for the capture of
carbon dioxide, Energy Fuels 23 (2009) 48404844.
[111] M.G. Plaza, C. Pevida, B. Arias, J. Fermoso, M.D. Casal, C.F. Martin, F. Rubiera, J.J.
Pis, Development of low-cost biomass-based adsorbents for postcombustion
CO2 capture, Fuel 88 (2009) 24422447.
[112] M.G. Plaza, C. Pevida, B. Arias, M.D. Casai, C.F. Martin, J. Fermoso, F. Rubiera,
J.J. Pis, Different approaches for the development of low-cost CO2 adsorbents,
J. Environ. Eng. 135 (2009) 426431.
[113] O. Leal, C. Bolivar, C. Ovalles, J.J. Garcia, Y. Espidel, Reversible adsorption of carbon dioxide on amine surface-bonded silica gel, Inorg. Chim. Acta 240 (1995)
183189.
[114] H.Y. Huang, R.T. Yang, D. Chinn, C.L. Munson, Amine-grafted MCM-48 and
silica xerogel as superior sorbents for acidic gas removal from natural gas,
Ind. Eng. Chem. Res. 42 (2003) 24272433.
[115] G.P. Knowles, J.V. Graham, S.W. Delaney, A.L. Chaffee, Aminopropylfunctionalized mesoporous silica as CO2 adsorbents, Fuel Process. Technol.
86 (2005) 14351448.
[116] G.P. Knowles, S.W. Delaney, A.L. Chaffee, Diethylenetriamine[propyl(silyl)]functionalized (DT) mesoporous silicas as CO2 adsorbents, Ind. Eng. Chem.
Res. 45 (2006) 26262633.
[117] P.J.E. Harlick, A. Sayari, Applications of pore-expanded mesoporous silicas. 3.
Triamine silane grafting for enhanced CO2 adsorption, Ind. Eng. Chem. Res.
45 (2006) 32483255.
[118] Y. Belmabkhout, A. Sayari, Effect of pore expansion and aminefunctionalization of mesoporous silica on CO2 adsorption over a wide
range of conditions, Adsorption 15 (2009) 318328.
[119] A. Sayari, M. Kruk, M. Jaroniec, I.L. Moudrakovski, New approaches to
pore size engineering of mesoporous silicates, Adv. Mater. 10 (1998)
13761379.
[120] A. Sayari, Unprecedented expansion of the pore size and volume of periodic
mesoporous silica, Angew. Chem. Int. Ed. 112 (2000) 30423044.
[121] R. Serna-Guerrero, Y. Belmabkhout, A. Sayari, Further investigations of CO2
capture using triamine-grafted pore-expanded mesoporous silica, Chem. Eng.
J. 158 (2010) 513519.
[122] N. Hiyoshi, K. Yogo, T. Yashima, Adsorption characteristics of carbon dioxide
on organically functionalized SBA-15, Micropor. Mesopor. Mater. 84 (2005)
357365.
[123] S.N. Kim, W.J. Son, J.S. Choi, W.S. Ahn, CO2 adsorption using aminefunctionalized mesoporous silica prepared via anionic surfactant-mediated
synthesis, Micropor. Mesopor. Mater. 115 (2008) 497503.
[124] R. Serna-Guerrero, E. Dana, A. Sayari, New insights into the interactions
of CO2 over amine-functionalized silica, Ind. Eng. Chem. Res. 47 (2008)
47614766.
[125] C. Knofel, J. Descarpenteries, A. Benzaouia, V. Zelenak, S. Mornet, P.L.
Llewellyn, V. Hornebecq, Functionalized micro-/mesoporous silica for the
adsorption of carbon dioxide, Micropor. Mesopor. Mater. 99 (2007) 7985.
[126] L. Wang, L. Ma, A. Wang, Q. Liu, T. Zhang, CO2 adsorption on SBA-15 modied
by aminosilane, Chin. J. Catal. 28 (2007) 805810.
[127] J. Wei, J. Shi, H. Pan, W. Zhao, Q. Ye, Y. Shi, Adsorption of carbon dioxide
on organically functionalized SBA-16, Micropor. Mesopor. Mater. 116 (2008)
394399.
[128] V. Zelenak, M. Badanicova, D. Halamova, J. Cejka, A. Zukal, N. Murafa, G.
Goerigk, Amine-modied ordered mesoporous silica: effect of pore size on
carbon dioxide capture, Chem. Eng. J. 144 (2008) 336342.
[129] V. Zelenak, D. Halamova, L. Gaberova, E. Bloch, P. Llewellyn, Amine-modied
SBA-12 mesoporous silica for carbon dioxide capture: effect of amine basicity
on sorption properties, Micropor. Mesopor. Mater. 116 (2008) 358364.
[130] C. Knofel, C. Martin, V. Hornebecq, P.L. Llewellyn, Study of carbon dioxide
adsorption on mesoporous aminopropylsilane-functionalized silica and titania combining microcalorimetry and in situ infrared spectroscopy, J. Phys.
Chem. C 113 (2009) 2172621734.
[131] C. Lu, F. Su, S.C. Hsu, W. Chen, H. Bai, J.F. Hwang, H.H. Lee, Thermodynamics
and regeneration of CO2 adsorption on mesoporous spherical-silica particles,
Fuel Process. Technol. 90 (2009) 15431549.
[132] M. Bhagiyalakshmi, L.J. Yun, R. Anuradha, H.T. Jang, Utilization of rice husk ash
as silica source for the synthesis of mesoporous silicas and their application
to CO2 adsorption through TREN/TEPA grafting, J. Hazard. Mater. 175 (2010)
928938.
[133] A. Zukal, I. Dominguez, J. Mayerova, J. Cejka, Functionalization of delaminated zeolite ITQ-6 for the adsorption of carbon dioxide, Langmuir 25 (2009)
1031410321.
[134] Z. Liang, B. Fadhel, C.J. Schneider, A.L. Chaffee, Stepwise growth of melaminebased dendrimers into mesopores and their CO2 adsorption properties,
Micropor. Mesopor. Mater. 111 (2008) 536543.

774

A. Sayari et al. / Chemical Engineering Journal 171 (2011) 760774

[135] Z. Liang, B. Fadhel, C.J. Schneider, A.L. Chaffee, Adsorption of CO2 on mesocellular siliceous foam iteratively functionalized with dendimers, Adsorption 15
(2009) 429437.
[136] C.W. Jones, J.C. Hicks, D.J. Fauth, M. Gray, Structures for capturing CO2 , methods of making the structures and methods of capturing CO2 , US Patent
Application No. US2007/0149398, 2007.
[137] J.C. Hicks, J.D. Drese, D.J. Fauth, M.L. Gray, G. Qi, C.W. Jones, Designing adsorbents for CO2 capture from ue gas hyperbranched aminosilicas capable of
capturing CO2 reversibly, J. Am. Chem. Soc. 130 (2008) 29022903.

[138] J.H. Drese, S. Choi, R.P. Lively, W.J. Koros, D.J. Fauth, M.L. Gray, C.W. Jones,
Synthesisstructureproperty relationships for hyperbranched aminosilica
CO2 adsorbents, Adv. Func. Mater. 19 (2009) 38213832.
[139] Y. Belmabkhout, A. Sayari, Isothermal versus non-isothermal
adsorptiondesorption cycling of triamine-grafted pore-expanded MCM-41
mesoporous silica for CO2 capture from ue gas, Energy Fuels 24 (2010)
52735280.

You might also like