You are on page 1of 16

Adsorbents for the post-combustion capture of CO

2
using rapid temperature
swing or vacuum swing adsorption
Niklas Hedin
a,
, Linna Andersson
a
, Lennart Bergstrm
a
, Jinyue Yan
b,c
a
Department of Materials and Environmental Chemistry and The Berzeli Center EXSELENT, Arrhenius Laboratory, Stockholm University, SE-106 91 Stockholm, Sweden
b
Chemical Engineering and Technology/Energy Processes, Royal Institute of Technology, Teknikringen 50, SE-100 44 Stockholm, Sweden
c
Sustainable Development of Society and Technology, Mlardalen University, SE-721 23 Vsters, Sweden
a r t i c l e i n f o
Article history:
Received 4 September 2012
Received in revised form 11 November 2012
Accepted 15 November 2012
Available online 20 December 2012
Keywords:
Zeolites
Metal organic frameworks
Amine-modied silica
Carbon capture and storage
Temperature swing adsorption
Vacuum swing adsorption
a b s t r a c t
In general, the post-combustion capture of CO
2
is costly; however, swing adsorption processes can reduce
these costs under certain conditions. This review highlights the issues related to adsorption-based pro-
cesses for the capture of CO
2
from ue gas. In particular, we consider studies that investigate CO
2
adsor-
bents for vacuum swing or temperature swing adsorption processes. Zeolites, carbon molecular sieves,
metal organic frameworks, microporous polymers, and amine-modied sorbents are relevant for such
processes. The large-volume gas ows in the gas ue stacks of power plants limit the possibilities of using
regular swing adsorption processes, whose cycles are relatively slow. The structuring of CO
2
adsorbents is
crucial for the rapid swing cycles needed to capture CO
2
at large point sources. We review the literature
on such structured CO
2
adsorbents. Impurities may impact the function of the sorbents, and could affect
the overall thermodynamics of power plants, when combined with carbon capture and storage. The heat
integration of the adsorption-driven processes with the power plant is crucial in ensuring the economy of
the capture of CO
2
, and impacts the design of both the adsorbents and the processes. The development of
adsorbents with high capacity, high selectivity, rapid uptake, easy recycling, and suitable thermal and
mechanical properties is a challenging task. These tasks call for interdisciplinary studies addressing this
delicate optimization process, including integration with the overall thermodynamics of power plants.
2012 Elsevier Ltd. All rights reserved.
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
2. Adsorbents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 420
2.1. Microporous physisorbents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421
2.1.1. Zeolites and aluminum phosphates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421
2.1.2. Carbon molecular sieves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424
2.1.3. Metal organic frameworks, covalent organic frameworks, and microporous polymers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424
2.2. Mesoporous physisorbents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424
2.3. Amine-modified mesoporous silica chemisorbents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 425
3. Swing adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 426
4. Structured and supported adsorbents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
4.1. Structured adsorbents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428
4.2. Supported adsorbents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
5. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430
0306-2619/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.apenergy.2012.11.034

Corresponding author. Tel.: +46 8 16 24 17; fax: +46 8 15 21 87.


E-mail address: niklas.hedin@mmk.su.se (N. Hedin).
Applied Energy 104 (2013) 418433
Contents lists available at SciVerse ScienceDirect
Applied Energy
j our nal homepage: www. el sevi er. com/ l ocat e/ apenergy
1. Introduction
The combustion of fossil fuel supplies the world with 81% of its
commercial energy, and releases 30 10
12
kg of CO
2
annually [1].
This release of CO
2
to the atmosphere has triggered changes in the
climate that call for urgent cuts in the emission of greenhouse
gases. Carbon capture and storage (CCS) is one of the viable ap-
proaches to reducing such emissions. In contrast with many other
approaches, CCS can be implemented in the existing energy infra-
structure. The difculties associated with its introduction lie in the
high cost of investment, and the high penalties from energy de-
ciencies. These costs are large compared with the economic bene-
ts from emission trading, as the cost of emitting one tonne of CO
2
(as of July 17, 2012) is 7.66, according the spot price of the Euro-
pean Union Emission Trading System.
CCS is conceptually straightforward: CO
2
is collected, trans-
ported, and nally stored for a long time. The capture of CO
2
is
probably the most expensive part of CCS, and its cost must be re-
duced signicantly if it is to be implemented in the current energy
system [1]. Established technologies exist for the transportation of
CO
2
, originally developed for enhanced oil recovery. The long-term
storage of CO
2
does not appear to be particularly expensive, but
eventual leaks or catastrophic failures are delicate issues that have
been only partly researched. On a global scale, CCS could reduce
the overall cost of mitigating climate change and increase the ex-
ibility of approaches used to reduce the emission of greenhouse
gases (GHG) [2]. According to a joint report by the Organisation
for Economic Co-operation and Development (OECD) and the Inter-
national Energy Agency (IEA), in a predicted scenario for 2050, CCS
will be the second-largest contributor to the mitigation of green-
house gas emissions, after energy efciency improvements [3]
(Fig. 1).
CCS can be implemented in various ways, using methods
including post-combustion, pre-combustion, and oxyfuel capture
[2,4]. The fact that the current energy infrastructure is based on
the combustion of fossil fuels means that the post-combustion cap-
ture of CO
2
represents a straightforward approach to CCS in the
short-term. Pre-combustion and oxyfuel capture have distinct
advantages in terms of the ease and cost of CO
2
capture; however,
it is unlikely that these methods can be introduced on a global
scale with the speed necessary to replace the introduction of
post-combustion capture. End-of-pipe solutions seem to be the
most viable in the short-term. The structure of the global energy
system has a massive inertia, and it cannot be changed rapidly.
However, it could be argued that oxyfuel retrots for supercritical
pulverized coal power plants have a signicantly lower capital cost
than post-combustion retrots [5].
A number of large-scale demonstration plants (equipped with a
capture capacity of 1 MtCO
2
/year) for the post-combustion capture
of CO
2
from coal-red power plants are under development, and it
is planned that they will be in operation within the next decade
[4]. Large demonstration plants are also planned for the post-
combustion capture of CO
2
from natural gas-red power plants.
The overall technical challenges stem from the additional energy
consumption associated with CCS. The estimated costs for the
post-combustion capture of CO
2
from coal-red power plants vary,
and are not reviewed here, but the cost of reducing CO
2
emissions
using CCS is expected to reduce signicantly over time [4]. Current
developments in the post-combustion capture of CO
2
have focused
mainly on the power and gas industry sector, but CCS technologies
could potentially be applied in other industrial processes. For
example, the cement and steel industries emit CO
2
on a globally
signicant scale. These other processes produce gas emissions with
varying partial pressures of CO
2
, and the costs associated with mit-
igating the release of CO
2
vary widely [6].
Liquid amine-based processes for the post-combustion capture
of CO
2
are closest to being commercialized. These absorption-
reaction processes have large thermal losses, which add large
parasitic contributions to CCS. Most of todays technologies for
separating CO
2
from N
2
-rich gases rely on absorption in aqueous
solutions of alkanolamines [7]. Alkanolamines have a saturated
hydrocarbon back bond with amine and hydroxyl groups. In the
absorptionprocesses, CO
2
acts as a solute andreacts chemicallywith
alkanolamines at low temperatures, whereas the N
2
does not. Rela-
tively pure CO
2
is recovered when the temperature is raised [8]. The
alkanolamine solutions are corrosive and must be diluted [9], which
leads to a large thermal loss on heating and cooling of the solvent.
The well-researched (and for other CO
2
separation processes, com-
mercialized) solvent monoethanol amine (MEA) has among the best
properties [10], but we refrain from reviewing the extensive litera-
ture on liquid amines for CCS, and instead focus on other reviews.
A technological breakthrough will depend on innovative methods
for CO
2
-over-N
2
separation, which could involve solvents, sorbents,
and membranes [2,11], and will involve system integration. Baxter
et al. [12] identied technologies that might lower the cost of CCS.
Many of these technologies involve functions and structures on
the nanometer length scale [12]. For CCS, they argued for an in-
creased focus (within 5 years) on developing newliquid absorbents
that could reduce costs and energy penalties. They concludedthat in
the mid-term (510 years), new nanostructured solids (adsorbents
or membranes) must be developed to reduce the cost of capturing
CO
2
fromue gases ona large scale [12]. Swing adsorptionprocesses
or membrane separation processes could be used. Ho et al. [13]
showed that adsorption-based processes can be used to separate
CO
2
from ue gases at a reduced cost compared with current tech-
nologies, if more CO
2
-selective sorbents could be developed.
Flue gases have various amounts and types of impurities that
may affect the separation of CO
2
over N
2
. The concentration of
the impurity components in the ue gas depends on the various
trade-offs that were made during the system design and optimiza-
tion, and are consequences of technical standards. The understand-
ing of how such impurities affect CO
2
capture, transport, and
Fig. 1. Contributions to greenhouse gas mitigation for six different scenarios according Philibert et al. [3].
N. Hedin et al. / Applied Energy 104 (2013) 418433 419
storage is indispensable [14]. Li et al. [1518] discussed and ana-
lyzed the effects of impurities by calculating the thermodynamic
properties of CO
2
in the presence of impurities. We expect the
presence of impurities to be at least equally as important in
adsorption processes as they are in absorption processes.
This review focuses on sorbents for the post-combustion cap-
ture of CO
2
, and the related swing adsorption, materials engineer-
ing, and system integration aspects. Other reviews present
information on the additional engineering aspects [7,19,1921].
We review the chemistry-related literature on adsorbents for the
years 20102011, and parts of 2012; for older studies, see Choi
et al. [22] and Hedin et al. [23].
2. Adsorbents
Gas molecules that are close to a surface achieve a reduced free
energy, as they are attracted to the electronic environment of the
surface [24]. The attractive gassolid interactions and the associ-
ated loss of entropy lead to an increase in the number density of
gas molecules close to the surface. For a gassolid system, gas mol-
ecules spend a much longer time on the surface than in the gas.
This surface excess of molecules is called adsorption. The thermo-
dynamic behavior of the gassolid system is described by adsorp-
tion isotherms (p,T), which capture how the surface excess
depends on temperature and pressure (see Fig. 2 for a representa-
tion of two such adsorption isotherms). The adsorption of CO
2
can
occur with or without the formation of chemical bonds. Molecules
physisorb via a range of physical interactions. For the physisorp-
tion of CO
2
, the electric quadrupole moment-electric eld gradient
interaction often dominates the interactions of CO
2
with a solid
surface. The magnitude of the electric quadrupole moment of
CO
2
is roughly three times larger than that for N
2
[30]; hence, it
interacts more strongly than N
2
with the electrical eld gradients
in porous materials. The partitioning of CO
2
on surfaces is higher
than that of N
2
in a gas mixture of CO
2
and N
2
[7].
The amount of molecules adsorbed increases with increasing
pressure, up to a maximum capacity. Because of the exothermic
nature of adsorption, the amount of gas adsorbed increases with
decreasing temperature. Since the uptake of gas depends strongly
on both temperature and pressure, gas separation processes based
on cyclic temperature or pressure variations have been devised. In
temperature swing adsorption (TSA), the adsorbent is regenerated
by raising the temperature. In pressure swing adsorption (PSA), the
gas components are captured at somewhat elevated pressures, and
the adsorbent is regenerated by lowering the pressure (it should be
noted that the pressures used are always higher than atmospheric
pressure, even after the pressure is decreased). Vacuum swing
adsorption (VSA) is similar to PSA, and is another non-cryogenic
gas separation process. The regeneration is conducted at reduced
pressures, and the adsorption occurs at ambient or near-ambient
pressures. In conventional PSA, the gas mixture is pressurized dur-
ing certain periods in the adsorption cycle. However, the mass
ows of N
2
and CO
2
are enormous in the ue gas stacks of power
plants, and it is difcult to imagine how such ows could be com-
pressed. In addition to the technical difculties with compression,
Chaffee et al. [25] indicated that regular PSA might also be too
expensive for the post-combustion capture of CO
2
. A 1000 MW
coal-red plant produces 1000 tonnes of dilute-stream CO
2
per
hour. Hence, we imagine that either TSA or VSA, or combinations
thereof, would be the optimal choice for the capture of CO
2
under
such conditions.
Adsorbents are divided into classes dened by their pore sizes.
According to the International Union of Pure and Applied Chemis-
try (IUPAC), micropores are smaller than 2 nm, mesopores are be-
tween 2 and 50 nm, and mesopores are larger than 50 nm. A high
capacity for the adsorption of CO
2
is essential for CO
2
-capture
adsorbents. In this review, we describe the microporous sorbents
as adsorbents, but for such microporous solids the difference be-
tween adsorption and absorption is somewhat arbitrary. Derouane
[26] discussed how the cases of volume-lling and surface cover-
age overlap when the pores become extremely small. Adsorbents
may have pores in the form of slits, cylinders, or cages intercon-
nected by pore windows. When the sizes of the pores, slits, or pore
windows are molecule-sized they can separate gas molecules by
equilibrium, kinetic, or molecular sieving mechanisms (Fig. 3).
Gas molecules have different effective kinetic diameters within
solids and in gases. CO
2
appears to have a smaller kinetic diameter
than N
2
in microporous solids. In the gas phase, CO
2
has a larger
kinetic diameter than N
2
[27]. The exact values appear to be sub-
strate-dependent, but values of 3.3 and 3.8 have been suggested
for CO
2
and N
2
in zeolites, respectively [28,29]. When the pore win-
dows are much larger than the sizes of CO
2
and N
2
, the potential for
A
B
C
Fig. 3. When the sizes of the pores, slits, or pore windows are molecule-sized they
can separate CO
2
from N
2
gas molecules by: (A) equilibrium, (B) kinetic, or (C)
molecular sieving mechanisms. Here, blue represent N
2
molecules and yellow CO
2
.
(For interpretation of the references to colour in this gure legend, the reader is
referred to the web version of this article.)
Low T
High T
n
1
n
2
High p
Low p
Pressure
A
m
o
u
n
t

Fig. 2. Two typical adsorption isotherms, which illustrate how the thermodynamic
behavior of the gassolid system depends on temperature and pressure.
420 N. Hedin et al. / Applied Energy 104 (2013) 418433
the separation of CO
2
and N
2
is determined by the differential equi-
librium partitioning of these molecules, as illustrated in Fig. 3A.
When the aperture of the pore openings approaches the size of
the somewhat larger N
2
, the CO
2
-over-N
2
selectivity of the adsor-
bent is enhanced by the difference between the diffusivities. The
diffusivity of N
2
is reduced signicantly when the dimensions of
the pore window aperture approach 3.8 ; in contrast, the diffusiv-
ity of CO
2
still matches well with the pore window (Fig. 3B). If the
pore window is smaller than 3.8 , the mixture of CO
2
and N
2
can
be molecularly sieved (Fig. 3C).
For VSA, the working capacity of an adsorbent is largely dened
by the differential between the uptake at a pressure at 0.1 bar and
the uptake under vacuum conditions. Table 1 gives details for dif-
ferent sorbents, with the temperature at which adsorption occurs,
the amount of CO
2
adsorbed at a partial pressure of 0.1 bar of CO
2
,
and the selection mechanism for the adsorption. To compare the
uptake levels with a hypothetical working capacity, one must as-
sume that the uptake under dynamic vacuum conditions is negligi-
ble. The capture of CO
2
would most likely be implemented after
ue gas condensation, when the temperature of the gas mixture
is 330 K. The adsorption of CO
2
is seldom studied at this temper-
ature: most of the values documented in the literature were deter-
mined at temperatures of 273298 K. The uptake of CO
2
is higher
at low temperatures than at high temperatures, due to the exother-
mic nature of adsorption. The low temperature and the absence of
a subtraction of the uptake under dynamic vacuum condition
means that the values in Table 1 are overestimates of the working
capacity. However, the table should still be helpful for a compari-
son of various sorbents.
In addition to a working capacity for CO
2
sorption, there is a
range of other physical, chemical, chemical engineering, and solid
mechanics criteria that must be fullled for a successful adsor-
bent/desorbent. Composition, particle size, pore size, pore volume,
and pore connectivity are all essential characteristics that affect
the function of a CO
2
sorbent. It appears that the heat of adsorption
for CO
2
should, ideally, have an intermediate value. If it is too high,
the sorbent will be difcult to regenerate economically, and if it is
too low the capacity for adsorbing CO
2
and the CO
2
-over-N
2
selec-
tivity will be too small [47]. An ideal CO
2
adsorbent should have a
high capacity for the uptake of CO
2
and a high CO
2
-over-N
2
selec-
tivity, and it should be easy to regenerate, have excellent mass
transport mechanical properties, be tolerant to water vapor, and
be robust to the presence of other gases.
2.1. Microporous physisorbents
Physical adsorption (physisorption) denotes cases with rela-
tively weak interactions at the gassolid interfaces. These interac-
tions are similar to those appearing in real gases, in terms of their
physical mechanisms. When adsorbed molecules undergo elec-
tronic recongurations at the surface, similar to a chemical bond,
we typically refer to the process as chemical adsorption (chemi-
sorption); chemisorption is dealt with in Section 2.2.
Microporous adsorbents have interconnected pores with mole-
cule-sized pore windows, and can be used to separate gas mole-
cules via equilibrium, kinetic, or molecular sieving mechanisms.
These adsorbents can be crystalline or amorphous. Crystalline
adsorbents have remarkably narrow pore-size distributions, due
to their well-dened atomic positions, whereas amorphous adsor-
bents have broader pore-size distributions, due to the more ran-
dom distribution of bond angles and bond lengths among their
atoms. Zeolites and metal organic frameworks (MOFs) are crystal-
line microporous adsorbents, whereas carbon molecular sieves
(CMS) and microporous polymers are amorphous microporous
adsorbents. Both classes are being studied extensively for their
properties as CO
2
adsorbents, and for their potential use as CO
2
sorbents in adsorption-based gas separation.
2.1.1. Zeolites and aluminum phosphates
Zeolites, which are crystalline microporous aluminosilicates
with large internal specic surface areas and volumes, have net-
works of interconnecting channels, or cages. Zeolites are described
by the number of oxygen atoms encircling their pore-window
apertures or channels (which form a percolating structure), and
they are referred to as 8-, 10-, and 12-membered ring zeolites; typ-
ical such zeolites are displayed in Fig. 4AC, respectively.
2.1.1.1. Hydrophilic zeolites. The zeolite NaX, which is the most
studied adsorbent for the capture of CO
2
[22,23], has a structural
code of FAU [48]. The cages of zeolite X are interconnected by large
windows, see Fig. 4C. The adsorption of CO
2
on zeolite NaX is large
even at relatively small pressures of CO
2
. Its large cages are con-
Table 1
Different sorbents, temperature at which adsorption occurs, amount of CO
2
adsorbed at a partial pressure of 0.1 bar of CO
2
, and the selection mechanism of the adsorption.
Sorbent T (K) n (mmol/g) Selection Ref.
Zeolites
NaX 273 4.0 Physisorption [31]
NaA 298 3.1 Physisorption [32]
Zeoltie-rho 303 3 Physisorption [33]
AlPO
4
-53 273 0.7 Physisorption [34]
SAPO
4
-56 273 2.5 Physisorption [31]
MOFs
Mg-MOF-74 273 5.8 Chem (?)/Phys [35]
rht-MOF-7 273 1.2 Physisorption [36]
CD-MOF-2 273 1.6 Physisorption [37]
SUMOF-3 273 1.1 Physisorption [38]
Microporous carbons and polymers
Carbon molecular sieve: VR-93 298 1.3 Physisorption [39]
Microporous polymer: PI-1 273 0.7 Physisorption [40]
Microporous polymer: BILP-.4 273 1.3 Physisorption [41]
Microporous polymer: Cs-N1 273 1.2 Physisorption [42]
COF-103 273 0.8 Physisorption [43]
Amine-based sorbents
Triamine modied silica (PE-MCM-41) 293 2.3 Chemisorption (Phys) [44]
Propylamine modied silica (AMS-6) 273 1.6 Chemisorption (Phys) [45]
Amine-modied MOF (CAU-1) 273 1.6 Chem (?)/Phys [46]
Note: For abbreviations see the text or the references.
N. Hedin et al. / Applied Energy 104 (2013) 418433 421
nected by 12-membered ring windows that do not hinder the local
diffusion of CO
2
or N
2
. Hence, the CO
2
-over-N
2
selection on NaX
operates via equilibrium mechanisms. CO
2
mainly physisorbs on
zeolite NaX, but a non-negligible amount of chemisorbed CO
2
is
present at low pressures of CO
2
. The Si-to-Al ratio of zeolite X is
1.3, and the aluminum gives the framework a negative charge
that is balanced by cations such as Na
+
. Zhang et al. [49] studied
and compared the capacity and kinetics of CO
2
adsorption on zeo-
lite NaX. The capacity to adsorb CO
2
was much higher on zeolite
NaX than on activated carbon at low pressures of CO
2
, despite
the fact that the activated carbon had a much higher specic sur-
face area. da Silva et al. [50] presented microcalorimetric data on
the adsorption of CO
2
on zeolite NaX, and showed that the differ-
ential enthalpy for adsorption decreased from 60 kJ/mol at low
loadings to 30 kJ/mol at intermediate and high loading of CO
2
.
These ndings are consistent with those of other studies [51,52].
Zeolite NaXs potential as a sorbent for CO
2
capture may be lim-
ited by its exceptionally hydrophilic nature. Wang et al. [53]
showed that the uptake of CO
2
on zeolite NaX was reduced from
3 mmol/g (0.1 bar) for pure CO
2
to 1.2 mmol/g when the sorbent
was simultaneously loaded with 3.4 mmol of H
2
O/g. The binary
adsorption data were described well by a model that used virial ex-
cess mixing coefcients (VECM), but not by the ideal adsorbed
solution theory (IAST). Earlier, Brandani and Ruthven [54] recog-
nized that small amounts of water reduced the uptake of CO
2
on
different versions of zeolite X; specically, NaX, CaX, LiLSX, and
NaLSX. Water adsorbed strongly to the cations. CO
2
was less prone
to adsorb when water had adsorbed, as water reduced the electri-
cal eld gradients within the zeolite X. Water vapor in ue gases
could therefore be a serious problem for zeolite NaX and other
hydrophilic zeolites. Rendering NaX hydrophobic or using other
means to moderate the competitive adsorption of CO
2
and H
2
O
could be benecial [53,54].
Zeolite Y has a structure similar to that of zeolite X, and has also
been investigated in detail as an adsorbent for CO
2
. It has the same
structural framework as zeolite X, but with an Si-to-Al ratio of 2
3 and fewer cations; this renders zeolite Y somewhat less hydro-
philic. Dragan [55] studied the adsorption of CO
2
(and SO
2
) on zeo-
lite Y at different temperatures and pressures of CO
2
, and Yan et al.
[56] studied the adsorption of CO
2
on zeolite Y, varying the cation
type from Na
+
to La
3+
and Ba
2+
. The temperature-programmed
desorption of CO
2
displayed a desorption signal at a temperature
of 325 C for zeolite NaY. This signal disappeared on zeolite LaY
and BaY, reecting the weaker chemisorption of CO
2
on zeolite
LaY and zeolite BaY than on zeolite NaY. Pirngruber et al. [57] stud-
ied the adsorption of CO
2
on alkali metal-exchanged zeolite Y,
using IR spectroscopy and DFT modeling. They proposed that a
simultaneous interaction of CO
2
with the cations and the frame-
work oxygens could explain the strong adsorption of CO
2
observed
on zeolite CsY and zeolite KY at low pressures of CO
2
. Bendenia
et al. [58] ion-exchanged Na
+
in zeolite NaX for Ni
2+
and Cr
3+
ions,
and studied the adsorption of N
2
and CO
2
, revealing that the ion-
exchanged forms displayed reduced adsorption of CO
2
and N
2
.
Zeolite A is another hydrophilic zeolite that has been investi-
gated in great detail as an adsorbent for CO
2
. The cages of zeolite
A (Fig. 4A) are interconnected by small windows, and the aper-
tures of these 8-membered ring windows are comparable in size
with the kinetic diameters of N
2
and CO
2
. The size of the aperture
Fig. 4. Illustrations of the structures of three zeolites and a metal organic framework (MOF): (A) zeolite A, (B) ZSM-5, (C) zeolite X and (D) MOF-5.
422 N. Hedin et al. / Applied Energy 104 (2013) 418433
can be further tuned using the extra-framework cations that bal-
ance the charge of the framework, as certain cation sites are pres-
ent near the pore window. Hence, zeolite A is a suitable molecular
sieve for small gas molecules. As discussed above, CO
2
appears to
have a smaller kinetic diameter than N
2
in microporous solids, and
values of 3.3 and 3.8 have been suggested for CO
2
and N
2
in zeo-
lites, respectively [28,29]. Zeolites CaA, NaA, and KA are called
zeolite 5A, 4A, and 3A, respectively, with the number denoting
the approximate size of the pore window aperture. Breck et al.
[59] showed that the capacity for the adsorption of CO
2
varied
with the Na
+
-to-K
+
ratio on zeolite NaKA, and Yeh and Yang [60]
extended this study to report enhanced CO
2
-over-N
2
selectivity
of CO
2
on zeolite NaKA, using percolation theory. Previous studies
examined aspects of this kinetic selection of CO
2
over N
2
on
zeolite NaKA [32,61], and Liu et al. [32] showed a high CO
2
-
over-N
2
selectivity by tuning the Na
+
-to-K
+
ratio in zeolite NaKA.
For zeolite NaA, the pore window is large enough for both N
2
and CO
2
to adsorb; however, this is not the case for zeolite KA.
For a carefully selected Na
+
-to-K
+
ratio, zeolite NaKA shows a
signicantly enhanced CO
2
-over-N
2
selectivity, compared with
zeolite NaA. K
+
preferably replaces Na
+
in the site close to the
pore-window aperture; hence, K
+
affects the size of the pore win-
dow even at small levels of exchange. The uptake kinetics of CO
2
on structured versions of zeolite NaKA have been examined previ-
ously, and the zeolite showed high selectivity for CO
2
-over-N
2
selection, and a rapid uptake of CO
2
[61]. It seems that the CO
2
selectivity of zeolite NaKA sorbents can be signicantly enhanced
by creating a differential between the CO
2
and N
2
diffusion rates,
without signicantly affecting the uptake kinetics, and that the
zeolite can be structured into monoliths that can be used in rapid
swing cycles. Zeolite NaKA exhibits advantageous properties for
the removal of CO
2
from dry mixtures of CO
2
and N
2
. It is likely
that an additional water-removal step would have to be imple-
mented for this adsorbent to be used for the removal of CO
2
from
ue gases. Of note, in the zeolite NaX family, zeolite NaKA is an
extremely hydrophilic sorbent.
Adsorbents for CO
2
are also researched and used for the pro-
cessing of natural gas and biogas. We do not provide an extensive
review of the literature related to this work, but we do mention
some studies that are important for the post-combustion capture
of CO
2
from ue gases. In the context of the upgrading of biogas
from landlls, Montanari et al. [62] studied the coadsorption of
CO
2
with methane on the zeolites X and 4A. Similarly as for meth-
ane, the capacity for the sorption of CO
2
was lower on zeolite A
than on zeolite 13X; however, the regeneration with N
2
was faster
for zeolite 4A than for 13X (zeolite 13X is another name for zeolite
NaX). They concluded that the capacity for the adsorption of CO
2
was enhanced in the presence of water on zeolite 13X, but not
on zeolite NaA. Such enhancement of the adsorption of CO
2
contra-
dicts the earlier ndings discussed above. It would be advanta-
geous if the effects of water on the adsorption of CO
2
on zeolites
NaX and Na(K)A could be claried further, as it would clearly be
advantageous if water enhanced the adsorption of CO
2
. Tomadakis
et al. [63] studied the separation of H
2
S from CO
2
on zeolites 4A,
5A, and 13X with PSA, and concluded that almost pure CO
2
could
be produced using zeolites 5A and 13X.
Bulnek et al. [64] studied the combined physisorption and
chemisorption of CO
2
on zeolite ferrierite, where the charge-
balancing ions on the zeolite were exchanged for alkali metal
cations (FER) (Li
+
, Na
+
, K
+
). They concluded that the energy of
adsorption with a low coverage of CO
2
was much higher for the
versions of zeolite ferrierite with a low Si-to-Al ratio than for those
with a high ratio. Ridha et al. [6567] investigated the feasibility of
the low-pressure encapsulation of CO
2
within zeolite NaKCHA,
arguing that the hysteresis loop observed in the low-pressure re-
gime could allow for the encapsulation of CO
2
. They studied the
adsorption of CO
2
and N
2
on zeolites LiCHA, NaCHA, and KCHA.
We believe that zeolite NaKCHA could potentially be applied for
the kinetic separation of CO
2
over N
2
. It would also be interesting
to compare the properties of zeolite NaKA and NaKCHA. Hudson
et al. [68] recently documented an enhanced CO
2
-over-N
2
selectiv-
ity in zeolite SSZ-13 with counterions exchanged for Cu
2+
; this zeo-
lite has a CHA structure and molecule-sized pore window
apertures.
In this subsection, we highlighted a study that investigated a
microporous titanium silicate [69]. This material is not strictly a
zeolite, per denition, but it is hydrophilic and related to zeolites.
Park et al. [69] studied the adsorption of CO
2
on a microporous tita-
nium silicate, ETS-10, which had its counterions exchanged for al-
kali cations. Among the studied forms, Na
+
-ETS-10 showed the
highest specic surface area, while Li
+
-ETS-10 showed the highest
uptake of CO
2
at 298 K. Li-ETS-10 showed the highest base
strength, and Cs
+
-ETS-10 showed the lowest.
2.1.1.2. Hydrophobic zeolites. Zeolites can be rather hydrophobic,
and such zeolites could be relevant for the capture of CO
2
in the
presence of H
2
O. Zeolites with a high Si-to-Al ratio are more hydro-
phobic than those with a low ratio. Lively et al. [70] showed that
such hydrophobic zeolites could be introduced as an adsorbent
within hollow bers. They showed that these ber-adsorbent con-
structs could, in principle, separate CO
2
at an energy cost of
0.13 GJ/ton-CO
2
captured, when certain heat integration prob-
lems were solved. In another study, Lively et al. [71] examined hol-
low polymeric bers with embedded adsorbent particles, and
showed how this composite could be used for rapid TSA processes.
Such hollow bers with adsorbents appear to be relevant for rapid
TSA, which could be regenerated by steam. Membranes that are
combined with sorbents are called mixed matrix membranes
(MMMs), and many studies have investigated these materials.
Chung et al. and Bernardo et al. reviewed this literature in detail
[72,73]. Yang et al. [74] studied ion-exchanged zeolite beta, and
their capacities as adsorbents for CO
2
. Zeolite beta typically has a
large Si-to-Al ratio, and is consequently a relatively hydrophobic
zeolite. They concluded that zeolite beta with K
+
ions showed the
best performance. Zukal et al. [75] measured and compared the up-
take of CO
2
on six high-silica zeolites: TNU-9, IM-5, SSZ-74, ferrie-
rite, ZSM-5, and ZSM-11. The channels of ZSM-5 (Fig. 4B) are
encircled by 10-membered rings. IM-5 had the highest capacity
to adsorption of CO
2
and could be highly relevant for further stud-
ies. Further developments in hydrophobic adsorbents with a signif-
icant capacity for CO
2
adsorption are expected.
2.1.1.3. Aluminum phosphates. The structures of microporous alu-
minophosphates (AlPO or AlPO
4
) and silicoaluminophosphates
(SAPO or SAPO
4
) are closely related to those of zeolites, and very
closely related to those of microporous silicates [76,77]. Liu [34]
studied a range of AlPO
4
structures with small pore apertures,
and concluded that certain structures exhibited an enhanced
CO
2
-over-N
2
selectivity. This enhancement was hypothesized to
be related to a kinetic effect. The overall uptake of CO
2
was some-
what lower than expected, possibly due to a non-perfect calcina-
tion. In SAPO
4
, the framework carries a negative charge, which is
balanced by cations. SAPO
4
-34 is an interesting adsorbent that
has the structural character of chabazite (CHA). Zhang et al. [78]
prepared Sr-SAPO
4
-34 and showed that it adsorbed more CO
2
than
did Na-SAPO
4
-34. In the context of inorganic membranes for
CO
2
-over-N
2
separation, Li and Fan [79] showed that SAPO
4
-34
membranes exceeded what was previously held to be the upper
bound for CO
2
-over-N
2
separation with membranes. This sorbent
could also be relevant to the study of adsorption-driven separation,
and could potentially be applied as the sorbent component in MMM
constructs. Cheung et al. [31] showed that the H-form of SAPO
4
-56
N. Hedin et al. / Applied Energy 104 (2013) 418433 423
exhibited a large capacity for CO
2
adsorption, similar to the uptake
on zeolite NaX. This large capacity for CO
2
adsorption was
combined with an uptake of H
2
O lower than that on zeolite X at
lowvapor pressures of water. These ndings suggest that the capac-
ity for the adsorption of CO
2
is signicantly higher on SAPO
4
-56
than on SAPO
4
-34.
2.1.2. Carbon molecular sieves
Carbon molecular sieves (CMSs) are microporous carbon-based
sorbents with molecule-sized pores. They are prepared in a multi-
step procedure that uses the carbonization of biomass (coconut
shell granules or similar), activation, deposition (typically using
chemical vapor deposition (CVD)), and the subsequent carboniza-
tion of aromatic molecules. These adsorbents have found applica-
tions in the production of N
2
from air. Ultrapure N
2
can be
produced via kinetically enhanced separation, using the differences
in the diffusion coefcients for N
2
and O
2
in a certain CMS [28,80].
Ruthven and Reyes [81] showed that in the kinetically enhanced
separation of gases, the particle size distribution (PSD) affects the
selectivity.
Alcaniz-Monge et al. [82,83] studied the adsorption of CO
2
on
CMSs in the form of particles and monoliths, and made a detailed
investigation of the effects of pore blockage. Bikshapathi et al.
[84] prepared CMSs from carbon bers, and studied the break-
through of adsorbed CO
2
. Silvestre-Albero et al. [39] studied CMSs,
and concluded that they are at least equivalent to, if not better
than, certain MOFs (zeolites 13X and 5A) in terms of several as-
pects relating to the adsorption of CO
2
. Wahby et al. [85] studied
similar CMSs and their potential use as selective adsorbents for
CO
2
. We believe that the possibilities for a kinetic selection of
CO
2
-over-N
2
on CMSs are worthy of more detailed study.
2.1.3. Metal organic frameworks, covalent organic frameworks, and
microporous polymers
Metal organic frameworks (MOFs) have been studied as adsor-
bents for CO
2
. MOFs are crystalline compounds with intercon-
nected pores, and are formed by metal ions coordinated with
rigid organic linkers. Fig. 4D shows a structural model of the
well-studied MOF-5. MOFs have been studied in some detail in re-
cent years for the separation of CO
2
in the upgrading of natural gas,
biogas, and landll gas, and for the capture of CO
2
from ue gas.
These applications were reviewed recently by Bae and Snurr [86],
who evaluated different MOFs in terms of their uptake of CO
2
, their
working capacity as a sorbent, their regenerability, their adsorptive
selectivity, and a combined adsorptive and desorptive selection
parameter. In their screening, they did not consider the kinetics
of CO
2
or any heat-related aspects. In their analysis, the following
MOFs displayed various characteristics that would be benecial in
adsorption-driven ue gas capture based on vacuum swing pro-
cesses: zeolitic imidazole frameworks (ZIFs) ZIF-78, -79, -81, -82
[87], and Co-carborane MOF-4b [88], Ni-MOF-74 [89], ethylenedi-
amine-H
3
[(Cu
4
Cl)
3
-(BTTri)
8
] [90], [ZN
2
(tcpb){p-(CF
3
)NC
5
H
4
}
2
] [91].
ZIFs have topologies identical to zeolites, and derive their structure
from heterocyclic and nitrogen-containing linkers [86]. In addition
to the research reviewed by Bae and Snurr [86], we highlight a few
recent studies. Debatin et al. [92] showed that MOFs based on the
in situ synthesis of an imidazolate-4-amide-5-imidate ligand dis-
played a signicant uptake of CO
2
. MOFs can form interpenetrated
structures with smaller pore sizes, compared with their non-inter-
penetrated equivalent structures. These interpenetrated structures
are promising for the separation of CO
2
from gas mixtures, because
the electrical eld gradients are larger in small pores than in large
pores. Yao [38] studied the uptake of CO
2
and N
2
on three such
interpenetrated MOFs based on Zn
4
O clusters and rigid dicarboxyl-
ates, and McDonald et al. [93] showed remarkably high capacities
and selectivities for the adsorption of CO
2
in an Mg-based MOF
modied with bifunctional amines. Functionalized MOFs can also
exhibit enhanced interactions with CO
2
, via either enhanced elec-
trical eld gradients, or specic interaction patterns induced by
the functionalization. Devic et al. [94,95] indicated that such inter-
actions can be of the donoracceptor type between certain hydro-
xyl groups and CO
2
. It is technically difcult to differentiate
between the contributions from enhanced electrical eld gradients
and donoracceptor interactions for functionalized MOFs, com-
pared with their non-modied equivalents. Spectroscopy, and in
particular IR spectroscopy, can be helpful for such differentiation.
Mg-MOF-74 displays very promising properties [35]; however,
it is unclear if the parasitic contributions for an adsorption-driven
separation of CO
2
-over-N
2
would be lower for MOFs than for zeo-
lites or similar adsorbents. A recent study by Lin et al. [47] indi-
cated that MOFs might have greater parasitic contributions than
certain zeolites. This indication, together with the potentially high
cost of production and the often water-sensitive nature of MOFs,
may limit their suitability as adsorbents for the post-combustion
capture of CO
2
from ue gases.
Recently, covalent organic frameworks and microporous poly-
mers have been studied in detail for various applications, including
the separation of CO
2
from ue gases. These microporous organics
can exhibit ultrahigh specic surface areas, and other relevant
properties [96]. Studies have shown that certain members of this
class of solids can adsorb signicant amounts of CO
2
and simulta-
neously display a high CO
2
-over-N
2
selectivity, under conditions
relevant for the post-combustion capture of CO
2
from ue gas
[4143,97100].
2.2. Mesoporous physisorbents
Mesoporous adsorbents have pores in the size regime of 2
50 nm, which allows for rapid mass transport. Belmabkhout et al.
[101,102] reported the high adsorption capacity of mesoporous sil-
ica for CO
2
, CH
4
, N
2
, H
2
, and O
2
, and Ma et al. [103] showed that the
adsorption of CO
2
on a MCM41 sorbent with both micro- and mes-
opores could be improved by tailoring the mesopores using either
hexadecyltrimethylammonium bromide (CTAB) as a soft template,
or mesoporous carbon as a hard template. These solids displayed a
high capacity for the adsorption of CO
2
at pressures of CO
2
higher
than those used in VSA or TSA. We doubt that mesoporous physi-
sorbents could be applied for the capture of CO
2
from ue gases;
however, the inclusion of mesoporosity within microporous sor-
bents could be particularly benecial for the mass transport kinet-
ics, and could allow for more rapid swing cycles.
Activated carbons are well-known adsorbents, and have micro-
and mesopores. Choi et al. [22] and Hedin et al. [23] compared the
performance of these materials for the separation of CO
2
, and re-
viewed the related literature. Their capacity for the adsorption of
CO
2
is typically large only at high pressures of CO
2
. Siriwardane
et al. [104] studied and compared the adsorption of CO
2
on zeolites
and activated carbons, revealing that at a CO
2
pressure of 4 bar and
25 C, the activated carbons had adsorbed approximately 50% of
their equilibrium capacity. These activated carbons showed signif-
icantly higher adsorption of CO
2
compared with zeolites at high
pressures of CO
2
, but displayed a lower uptake of CO
2
at pressures
of CO
2
relevant for the removal of CO
2
from ue gases using VSA
and TSA.
Dantas et al. [105] studied breakthrough curves for CO
2
, and
concluded that amine-modied active carbons did not display lar-
ger capacities than unmodied carbons. An et al. [106] fabricated
activated carbons fromphenolic resins. These materials had mainly
micropores, and the researchers concluded that small pores are
superior to large pores when the partial pressure of CO
2
is low in
the mixture from which the CO
2
is to be separated. Activated car-
bons have a weaker physical interaction with CO
2
than aluminum-
424 N. Hedin et al. / Applied Energy 104 (2013) 418433
rich zeolites such as zeolite NaX and NaA. This means that
although the specic surface area of activated carbons can be large,
they typically display a rather small uptake of CO
2
at low pres-
sures, which renders them less relevant as adsorbents for VSA.
2.3. Amine-modied mesoporous silica chemisorbents
Chemisorbents display properties that give them the potential
to be applied for the separation of CO
2
from ue gases, and as sor-
bents for the chemical looping cycle (CLC). Current research fo-
cuses on weak chemisorbents such as amines, and strong
chemisorbents such as strong bases. For strong chemisorbents that
are relevant for CLC, we refer to the review by Choi et al. [22]. Here,
we review recent studies on amine-based silica. Substrates other
than silica have been used as supports for amines, including car-
bons, zeolites, and MOFs [107112].
Alkanolamines in water and chilled ammonia have been studied
for the capture of CO
2
from ue gases. In the liquid phase, the CO
2
-
amine chemistry is largely known [113,114]. Amines react with
CO
2
and form various compounds. Masuda et al. [115] showed that
carbamates were formed in protophobic solvents, carbamic acids
were formed in protophilic solvents, and HCO

3
formed in propa-
nol/water mixtures. The local chemistry at the silica interface is ex-
pected to be somewhat different from that in water, which could
open up different energetics and kinetics for the sorption of CO
2
.
Additional benets could result from tethering the amine on solid
substrates, in terms of the potential environmental and health con-
cerns regarding byproducts from the liquid-based processes. The
high volatilities of the amines and byproducts could be reduced
by tethering the amines. Certain alkanolaminewater mixtures
are particularly corrosive, and tethering the amines to substrates
could reduce such corrosion.
In analogy with the chemistry of amines and CO
2
in the liquid
phase, amines and polyamines have been tethered to silica sup-
ports, or have been introduced simply by coating or lling the
pores. While silica-amine sorbents can be loaded with large
amounts of amines, there is a tradeoff between amine loading
and the remaining surface area. When the pores of the substrate
are lled with the amine, the capacity for the adsorption of CO
2
is large, but there is no room for gas transport within the compos-
ite sorbents, and the advantage of rapid gas-phase transport is lost.
Hence, the amine-lled porous materials act primarily as absor-
bents, rather than adsorbents.
Recently, there has been discussion regarding which silica sub-
strate should be used. Mesostructured silica with ordered pores
has typically been used as a substrate for amines. Although the
adsorption data may be easier to interpret from such well-ordered
substrates, there are both economic and technical advantages in
using unordered and mesoporous silica supports [116]. However,
mesostructured silica supports are commonly rather unstable un-
der steam conditions [117].
The pores of the silica should not be too small. It has been
shown that it is advantageous to expand the pores of MCM-41
and similar silica substrates before tethering the amines [44,118
123]. MCM-48 is a silica substrate with unusually large specic
surface areas, and a three-dimensionally interconnected structure.
Bacsik et al. [45] observed a higher uptake of CO
2
on AMS-6 than
on MCM-48 when both were modied with tethered amine groups.
AMS has larger pores, but the same structure as MCM-48.
The chemistry governing the reactions of CO
2
with amines on
silica supports is not fully understood. Molecular spectroscopy
techniquesmost often Fourier transform infrared (FTIR) spectros-
copy, but also
13
C nuclear magnetic resonance spectroscopyhave
been used to study the chemistry of these reactions [44,124129].
When tethered amines react with CO
2
they mainly form carba-
mateammonium ion pairs, when the amine density is large en-
ough. Two amine groups react with one CO
2
molecule.
Carbamateammonium ion pairs are the main control on whether
water vapor is present. At low amine densities, carbamic acid and
silylcarbamates tend to form in parallel with each other. Such sily-
lcarbamates appear to form only under dry conditions. The uptake
of CO
2
typically increases under moist conditions [44,130,131].
Danon et al. [125] excluded the formation of HCO
3
-
during the
adsorption of CO
2
on SBA-15 mesoporous silica modied with (3-
aminopropyl) triethoxysilane, even with water present. Numerous
authors have speculated about the formation of HCO

3
, perhaps in
analogy with reactions with liquid amines. The uptake of CO
2
on
amines tethered to sorbents at low pressures of CO
2
tends to have
a critical dependence on the amine density. The low-pressure up-
take becomes signicant only when the local amine density is high
enough for ammoniumcarbamate ion pairs to form. Previous
studies systematically studied this effect for monoamines tethered
to silica [131133].
It is important to quantify the contributions from the physi-
sorption and chemisorption of CO
2
on amines, because this balance
determines the regeneration conditions that are necessary for the
recovery of CO
2
from the sorbent. Thermodynamic methods (such
as conventional adsorption measurements) are not ideal for the
differentiation of such contributions, especially for cases when
the heats of adsorption are similar between physisorption and
chemisorption. Aziz et al. [127] quantied these contributions by
using infrared spectroscopy to determine the chemisorption and
physisorption of CO
2
on amine-modied silica, and showed that
CO
2
mainly chemisorbed, even at low amine densities.
Mesoporous silicas have also been modied with amine or
amine-like functionalities other than (3-aminopropyl)triethoxysi-
lane. For example, Kim et al. [134] recently studied derivatives of
hydroxylated amidine in their solid state for the capture of CO
2
,
Belmabkout et al. [135,136] studied silica modied with triamines,
which showed a signicant uptake of CO
2
, and Serna-Guerrero
[137,138] modeled the kinetics and thermodynamic of the uptake
of CO
2
on related sorbents.
Hicks et al. [139] studied tethered and hyperbranched polyeth-
yleneimine (PEI) on mesoporous silica; this system displayed effec-
tive and reversible capture of CO
2
[140]. These composites
displayed high amine loading, and chemical anchoring of amines,
which made them attractive for the capture of CO
2
. Drese et al.
[141] measured an extraordinary CO
2
-uptake capacity of
5.6 mmol/g for hyperbranched aminosilica (HAS) on SBA-15 sup-
ports, and Choi et al. [142] investigated related adsorbents for their
potential use in the capture of CO
2
from dilute mixtures, which
could be useful for applications in submarines or space shuttles,
or, speculatively, for the capture of CO
2
from air. Satyapal et al.
[143] discussed these hybrid sorbents, which were based on a por-
ous polymer, oligomeric amines, and a polymeric modier that en-
hanced the rate of uptake of CO
2
. Using a similar approach, several
groups have studied mesoporous silica lled or coated with
amines, revealing that tetraethylenepentamine (TEPA) in particular
showed signicant potential for the capture of CO
2
[144147]. In
such composite sorbents, polyethylene glycol enhanced the CO
2
sorption kinetics of TEPA [148]. Using impregnation from solution,
Ebner et al. [149] introduced up to 40 wt.% of polyethylene(imine)
into a support of porous silica in the form of beads. During the
application of up to 76 consecutively applied adsorptiondesorp-
tion cycles, the working capacity for the sorption of CO
2
under iso-
thermal conditions ranged from 0.25 to 2.8 mol/kg, with a
maximum observed at a temperature of 80 C. Jiang et al. [150]
studied the CO
2
adsorption capacity of porous spheres of poly-
methylmethacrylate covered with multilayers of amine-based
polyelectrolytes, revealing that up to 10 layers of PEI and polysty-
rene sulfonate (PSS; which were added using a layer-by-layer pro-
cess) resulted in a capacity for CO
2
adsorption of 1.7 mol/kg
N. Hedin et al. / Applied Energy 104 (2013) 418433 425
sorbent. Wang et al. [151] improved the CO
2
adsorption capacity of
such lms (containing PEI) by introducing surfactants to create
additional pathways for the CO
2
to diffuse into the lms. At
30 C, the porous lm with PEI adsorbed up to 142 mg/g of CO
2
,
and >50% of the amines reacted with CO
2
. Films with PEI and
surfactants showed improved CO
2
sorption kinetics compared with
those without surfactants, and the regeneration performances
were comparable for lms with and without surfactants. The PEI
lms were deposited on three types of mesoporous supports: hier-
archically porous silica, mesoporous carbon aerogels, and meso-
porous silica of the type SBA-15. Qi et al. [152] studied the
capture of CO
2
on nanocomposite sorbents based on capsules of
mesoporous silica functionalized with PEI or TEPA. Under simu-
lated ue gas conditions, the TEPA-functionalized capsules showed
a maximum capacity of 7.9 mmol/g for the adsorption of CO
2
. The
kinetics for the CO
2
uptake were fast: within the rst few minutes,
90% of the total CO
2
uptake capacity was lled. The amine-
functionalized capsules, with a thin and mesoporous silica shell
(12 nm) and a large interior void volume, showed the highest
capacity to adsorb CO
2
. The use of larger particles increased both
the optimal amine content and the capacity for the adsorption of
CO
2
. Composite sorbents impregnated or lled with oligomeric
amines could potentially be used for the large-scale capture of
CO
2
; however, their longevity has been questioned [140].
Oxidation in the presence of oxygen could degrade amine-based
sorbents. Flue gases can contain signicant amounts of oxygen gas,
and monoethanolamine undergoes such oxidative degradation.
Bollini et al. [153] showed that tethered primary and tertiary
amines were stable in an oxidizing environment with a somewhat
elevated temperature, but secondary amines degraded. They con-
cluded that it would be difcult to use secondary amines in the
post-combustion capture of CO
2
. This limits the potential of teth-
ered secondary diamines, triamines, and TEPA and PEI supported
on silica for the adsorption-driven post-combustion capture of
CO
2
in the case of a high oxygen content in the ue gas.
The parasitic losses related to the chemical reactions of amines
with certain impurities have received little attention. Many amines
tend to form heat-stable salts with SO
2
and carbonyl sulde (COS)
[154], and one would expect amine-modied silica to react in a
similar way, possibly with an added parasitic contribution. We
have not found any studies investigating how irreversible reactions
between COS and amine groups could affect amine-based sorbents.
3. Swing adsorption
Swing adsorption processes make use of the differences be-
tween the tendencies for adsorption of the individual components
in multicomponent gaseous mixtures (such as ue gases from
combustion). The most common swing adsorption methods use a
cyclic variation of the pressure or temperature. These methods
are known as pressure swing adsorption (PSA), vacuum swing
adsorption (VSA), thermal swing adsorption (TSA), and the related
electrically induced thermal swing adsorption (ESA) [28,155].
Here, we review the extensive literature on the optimization
and variation of these swing adsorption processes, focusing on
the removal and concentration of a minority component of CO
2
(heavy) from N
2
(light), due to its potential use in post-combustion
capture. Industrial PSA processes are closely related to the original
Skarstrom cycle [156], and nd uses in, for example, renery tech-
nology [78]. This cycle includes two beds, which are subjected to
pressurization, adsorption, blow down, and purge, with a pattern
that maintains a continuous stream of product. Fig. 5 illustrates
these developments for one of the beds in this cycle. The bed is
pressurized with CO
2
-lean N
2
-rich gas in step A, as illustrated in
Fig. 5A. The ue gas is fed to the bed in step B, displayed in
Fig. 5B and CO
2
-lean gas is produced. CO
2
-rich gas is produced in
step C (blow down) and step D (purge), as illustrated in Fig. 5C
and D.
These PSA processes typically purify weakly adsorbing species
(e.g., H
2
) from gaseous mixtures with strongly adsorbing species.
CO
2
typically adsorbs more strongly to surfaces than N
2
, due to
its larger electric quadrupolar moment. To recover CO
2
from a mix-
ture with a large proportion of N
2
, other PSA processes have been
developed, modeled, and evaluated. The methods used to recover
heavy components using PSA are known as heavy-reux PSA
processes. A rinse step is introduced in the Skarstrom cycle after
the feed, which equilibrates the beds and ushes out the light
product (N
2
) before the heavy component (CO
2
) is recovered. A
so-called dual-reux PSA process has been developed, and models,
descriptions, and experimental demonstrations for this process
have been presented [157159]. In this particular PSA cycle, a rinse
step was introduced to increase the purity of the strongly adsorbed
component (CO
2
) over the weakly adsorbed component (N
2
);
however, Park et al. [160] showed that the rinsing dramatically
increased the power consumption. Fiandaca et al. [161] presented
a preliminary investigation of an algorithm for fast-cycle PSA
operation. Although their study focused on the production of N
2
from air, it was also helpful for the optimization of CO
2
separation.
Thakur et al. [162] recently argued that the duplex PSA
proposed by Leavitt [163] could be modied to achieve a compact
process.
It is likely that conventional PSA processes will be difcult to
implement for the capture of CO
2
in ue gases from large point
sources. Indeed, Chaffee et al. [25] indicated that this process is
too expensive [25]. We believe that the technical difculties
mainly lie in the pressurization steps in PSA, and in the integration
of the enormous ow in the ue gas stack for a normal power sta-
tion. However, there are advantages in combusting fuel at some-
what elevated pressures [164], and for such power plants the
application of rapid PSA could be more straightforward. Rapid
VSA and rapid TSA are more promising for the capture of CO
2
from
ue gases. Note that the CO
2
capture step in VSA and TSA must be
integrated with the complex thermodynamics of the power gener-
ation plant itself. The separation performance can be challenging,
due to the different compositions of the ue gases and the varia-
tions in the conguration of the VSA/TSA with the power plant. Re-
cent studies have investigated the integration of CCS with the
various components in a power generation system[165172]. Such
integration may result in a reduction in the energy consumption
N
2
CO
2
N
2
CO
2
:N
2
(15:85)
N
2
CO
2
N
2
CO
2
CO
2
N
2
N
2
N
2
CO
2
A B C D
Fig. 5. One of the beds in a Skarstrom PSA cycle, which is subjected to: (A)
pressurization, (B) adsorption of CO
2
and production of N
2
, (C) blow down with
production of CO
2
and (D) purge with production of CO
2
.
426 N. Hedin et al. / Applied Energy 104 (2013) 418433
and the costs of the overall system. The composition of the ue
gases depends on the fuel and type of combustion used, but it is
generally dominated by N
2
, 3%15% CO
2
by volume, H
2
O, O
2
, Ar,
and small amounts of SOx and NOx [5,173]. Many amines tend
to form heat-stable salts with SO
2
and COS [153], and such salts in-
crease the parasitic contributions to the CO
2
separation and
decrease the lifetime of the sorbent. Besides the thermal perfor-
mance, it is essential to optimize the economic performance by
considering external factors, which could include policy incentives
and regulations devised for the implementation and commerciali-
zation of CCS [5].
In VSA, the strongly adsorbed component is recovered at sub-
atmospheric pressures. Such vacuum recovery of adsorbed CO
2
is
promising for the separation of CO
2
from ue gases in a dual-bed
conguration [157159,173176]. Ho et al. [13] showed that a
combined process using zeolite 13X with both high-pressure and
vacuum desorption steps resulted in a predicted cost for CO
2
re-
moval that was comparable to that of CO
2
capture using MEA
absorption. The selectivity of Zeolite X is high, but not high enough
to give pure CO
2
with a single step. However, if the CO
2
-over-N
2
selectivity could be enhanced, the purity could be increased, and
the cost of the separation could be reduced. Recently, Delgado
et al. [177] used theoretical modeling to evaluate the feasibility
of CO
2
recovery using activated carbon in a novel VSA cycle. Their
model described the structure of the setup using the bed void frac-
tion and an axial dispersion coefcient, and the particle diameter
in Erguns equation [177,212]. The theoretical results indicated
that it would be possible to recover CO
2
with high purity (>93%)
from a mixture containing 13% CO
2
at 40 C, with higher recovery
(>90%) and lower power consumption (<0.12 kW h/kgCO
2
) than
other processes that use rinse steps to reach the same purity. The
specic power consumption would be somewhat higher for a pur-
ity of >95%, which appears to be a specic requirement [178]. Aar-
on and Tsouris [7] suggested that PSA and VSA likely cannot be
used (with a standard sorbent) as a stand-alone processes for the
capture of CO
2
; however, they suggested that they could be used
as a pre-step before an absorption-driven process. This adds to
the existing motivation for the further development of adsorbents
designed for processes with small parasitic contributions to the
separation of CO
2
from ue gases.
In TSA, steam can be used for the regeneration of adsorbents,
thereby giving TSA an advantage over VSA, which requires a vac-
uum. Merel et al. [179] proposed, and demonstrated experimen-
tally, the use of TSA for the capture of CO
2
from ue gases.
Recently, the prospects for the economic implementation of TSA
were strengthened by Grande et al. [180,181] in their evaluation
of CO
2
capture using an ESA process. They estimated the energy
consumption associated with the separation of CO
2
to be 2.04 GJ/
tonne CO
2
, for an adsorbent consisting of 70 wt.% zeolite X and
30 wt.% of a conducting binder. Moate et al. [182] performed a
numerical study of the material energy balances for the capture
of CO
2
from N
2
/O
2
/CO
2
mixtures using a TSA system, concluding
that detailed heat management is necessary to maximize the regen-
eration efciency. Combining vacuum temperature swing cycles
with steam may offer advantages, as argued by Dutcher et al. [183].
The sheer size of typical power plants places certain constraints
on the cycle times that can be used in capturing CO
2
using swing
adsorption processes. The ow of gas through the stack of a power
plant is enormous: a 1000 MW coal-red plant produces
1000 tonnes of CO
2
/h in a dilute stream. A rough calculation gives
a relation between the mass of the sorbent needed and the cycle
time as follows: (mass adsorbent) (relative capacity) = (ow
rate) (cycle time) (number of beds). A typical cycle of 10 min
would demand 3000 tonnes of adsorbents. These quantities of
adsorbents are impractically large, and would lead to the need
for huge amounts of processing equipment. By reducing the cycle
times, smaller quantities of adsorbents could be used, and the cost
for separation could be reduced. Methods to achieve rapid swing
cycles, and technologies associated with these processes, are cur-
rently under development [184]. We strongly believe that the
use of structured or supported adsorbents will be crucial for such
fast-cycling of VSA, VPSA, or TSA. Section 4 reviews current devel-
opments related to such structured adsorbents.
In an extensive screening study, Harlick and Tezel [185] con-
cluded that zeolites NaX or NaY were the best adsorbents for cap-
turing CO
2
using a PSA process, in terms of the capacity and
robustness during cyclic adsorption. The PSA/VSA process had
low partial pressures of CO
2
in the feed, and used low regeneration
pressures. Gomes et al. [186] investigated a two-bed PSA process
using experiments and calculations, and a range of adsorbents
was tested (zeolite 5A, NaX, one type of CMS, and gamma alumina).
A breakthrough test indicated that zeolite NaX was the most suit-
able sorbent. Tlili et al. [187] studied the small-scale capture of CO
2
using zeolite 5A, and suggested that VSA may be more useful on a
large scale, while Yoshida et al. [188] showed that trace compo-
nents of a heavy component (CO
2
or Xe) could be successfully en-
riched from air using an enriching reux PSA process with a
parallelized equalization step.
Harlick and Tezel [189] determined the CO
2
adsorption iso-
therms for NaY at different pressures and temperatures (20
200 C, 0.0052.0 bar). The collected data were described well by
the Toth isotherm. From this isotherm, they derived the working
capacities for a few PSA, TSA, and pressure temperature swing
adsorption (PTSA) cycles. For the NaY, they predicted that a PSA
or PTSA process would have the highest capacity when operated
under steep gradient (temperature or pressure) conditions. Phelps
and Ruthven [190] tested the performance of a counter-current ad-
sorber (with silicalite) on an endless belt. Although an endless belt
adsorber would be volume-efcient, its engineering aspects are
complex, and this method is unlikely to nd widespread usage.
Sorbents with mesopores allow faster gas diffusion than micro-
porous sorbents. Mesopores have the potential to reduce the limita-
tions associated with mass transport in micropores during loading
and unloading on the sorbents. Chaffee et al. [25] reported on a
range of amine-modied sorbents with mesopores for applications
in PSA/VSA processes, emphasizing the need to study the effects of
impurities. Of note, it is important in general to study the effects of
impurities on post-combustion capture processesthis is by no
means specic to adsorption-driven processes. In addition, Chaffee
et al. [25] presented data for amine-modied adsorbents that ad-
sorbed more CO
2
at higher temperatures than at low temperatures.
This dependency occurred because the pores became lled with the
reactive aminosilane during synthesis, and access to the pores was
therefore diffusion-limited at low temperatures. Dasgupta et al.
[191] employed PEI-impregnated SBA-15 in a ve-step PSA cycle,
and concluded that the hybrid sorbent was comparable to commer-
cial zeolite adsorbents in terms of productivity (kg-CO
2
/min). The
adsorbent could be regenerated under typical PSA conditions.
Wurzbacher et al. [192] studied temperature-vacuum swing
processes using a sorbent based on a commercial silica gel with
tethered amine groups, revealing a stable performance over 40
adsorption/desorption cycles. Bastin et al. [193] gave the rst dem-
onstration of the separation and removal of CO
2
from binary mix-
tures of CO
2
/N
2
and CO
2
/CH
4
, and from ternary mixtures of CO
2
/N
2
/
CH
4
, using a xed bed with a sorbent based on an interpenetrated
metal organic framework MOF-508b.
4. Structured and supported adsorbents
Conventional swing processes with adsorbents rely on the use
of beds of adsorbents in the form of packed beads or pellets, with
N. Hedin et al. / Applied Energy 104 (2013) 418433 427
the size of the pellets typically being several micrometers [24].
Packed beds have a low operational cost and allow for continuous
operation, but the temperature control is poor and heat waves can
develop, thereby reducing the beds abilities to separate CO
2
from
N
2
. The cycle times for packed beds are commonly on the order of
tenths of minutes, where the mass transfer resistance is not a lim-
iting factor. The beds operate under the uidization threshold,
which makes it difcult to reduce the cycle times. For the rapid cy-
cles needed for the capture of CO
2
from ue gases, the contact time
between adsorbent and adsorbate is likely to be on the order of
seconds. Hence, the timescale of adsorption in the column must
be signicantly reduced to match that of the packed beds, to avoid
the immediate breakthrough of the input gas [194,195]. The time-
scale of adsorption in a packed bed can be reduced by using smal-
ler beads or pellets, thus creating shorter diffusion paths, but the
use of smaller particles will also lead to unmanageable pressure
drops and lowmechanical strength. Structured or supported adsor-
bents offer an alternative to packed beds for swing adsorption pro-
cesses, and could potentially achieve much faster mass transfer
kinetics, smaller pressure drops, improved heat transfer, and a
more uniform temperature distribution. Such rigid and durable
adsorbent structures could also ensure a long lifetime for the
assemblies and a correspondingly high number of operating cycles.
4.1. Structured adsorbents
For all swing adsorption processes, the microscopic sorbent par-
ticles must be assembled into macroscopic shapes, as a ne pow-
der cannot be used directly. For such assemblies, the powder is
structured using different kinds of binders, but binderless methods
are currently being developed. Pavlov et al. [196] reviewed the lit-
erature on the binder-free granulation of zeolites NaX and zeolite
A. These methods could enhance the overall capacity, and possibly
the selectivity, by minimizing the use of binders; however, beads
are less suited for rapid swing cycles, as uidization should be
avoided in PSA/VSA/TSA with beds [197].
Adsorbents can also be structured into more advanced macro-
scopic shapes, including hierarchically porous structures, lami-
nates, and foams that provide enhanced transport and allow the
possibility of using rapid cycle times. Rezaei and Webley [198] pre-
sented a methodology to compare different congurations of struc-
tured adsorbents, indicating that, for example, laminated
structures can be superior to pellets if the sheet is thin enough. Re-
zaei et al. [199,200] showed through a theoretical study that adsor-
bent particles with highly branched and macroporous channels
and an overall macroporosity of 4050% represented a near-opti-
mal structure, when the working capacity for cyclic adsorption
processes was maximized. Rezaei et al. [201] developed a mathe-
matical model to simulate the diffusion and adsorption in a system
of zeolite X lms on cordierite supports, and parameterized this
model to t experimental data. The model indicated that a lm
thickness of 10 lm would be optimal for nonporous supports with
1200 cpsi. For these dimensions, the capacity for CO
2
adsorption
would approach the adsorption capacity of beads, and the CO
2
adsorption kinetics would not be compromised. A simplied math-
ematical model was employed to describe the (dispersed plug)
ow within the monolith, and it was assumed that the uptake of
CO
2
within the zeolite lm and on the porous support occurred
in parallel, due to the presence of open grain boundaries and cracks
in the zeolite lm.
Monoliths could have signicant advantages over packed beds
of adsorbents in, for example, rapid VSA for the post-combustion
capture of CO
2
. Higher massows could be used through the mono-
lith than through a packed bed, if the monolith was appropriately
structured to reduce the pressure drop. The kinetics for the adsorp-
tion of CO
2
have been studied in monolithic carbons by Brandani
et al. [202]. Burchell et al. [203] developed a composite adsorbent
for usage in TSA/ESA systems, where the monolithic body allowed
for rapid heating. The combination of the monolithic body with an
appropriate Ohmic resistance enabled the bed to be heated rapidly
using a current of 5A. Hao et al. [204] synthesized carbon-based
monoliths with pore structures on multiple length scales; these
monoliths were nitrogen-doped, and exhibited a signicant CO
2
uptake capacity and CO
2
-over-N
2
selectivity. Akhtar [205,206]
developed a monolithic form of zeolite NaX, using a slip casting
process followed by a thermal treatment on sacricially templated
zeolite monoliths, and Schumann et al. [207] reduced the amount
of binder used in the granulation of zeolite X via a specialized
treatment. Akhtar [61] showed recently that mechanically strong
zeolite NaKA monoliths had the highest estimated gure of merit
with a K
+
content of 10 at%. The gure of merit combined the
CO
2
-over-N
2
selectivity (in an IAST model) with the time depen-
dent capacity to adsorb CO
2
. Natural extensions of this study
would include structuring such zeolite NaKA monoliths to reduce
the pressure drop, or introducing multifunctionalities to enable
ESA. This could assist in assessing the potential of this technique
for practical use in the post-combustion capture of CO
2
from dry
ue gases.
The adsorption and desorption of CO
2
on surfaces are exother-
mic and endothermic processes, respectively, and cause a thermal
wave to develop within the adsorbent bed [80]. The velocity of this
temperature wave can be many times higher than that of the
adsorbing gas [208210]. The different rates of heat and mass
transfer within the bed can have detrimental effects on the overall
performance of the adsorbent. The heat transfer in structured
adsorbents can be improved compared with packed beds by incor-
porating a thermally conducting support material [204,211].
Increasing the heat transfer is of high importance for the achieve-
ment of rapid temperature swings, and could be helpful in other
congurations, where it could decouple the rates of heat transfer
from the rates of mass transfer; this would make the processes
more robust to variations in the operational parameters. The high
porosity and thin walls of a macroporous structure may also im-
prove the transport of heat through convection, resulting in a more
uniform distribution of heat [198,200].
Small pressure drops can be achieved in structured adsorbents
by using rapid cycling, and by increasing the number of accessible
pathways and widening the channels in the adsorbents [198,200].
However, it should be noted that the volumetric working capacity
of the adsorbent decreases when the size and number of passages
is increased. Introducing transfer paths of random orientation will
increase the average path of transportation for gas molecules,
which is typically quantied by the tortuosity (a geometric factor
that gives the measure of a geometrical constraint independent
of temperature and the nature of the gaseous species [24]). The tor-
tuosity is commonly used in empirical and theoretical models such
as Erguns equation and the KozenyCarman equation, which de-
scribe the pressure drop and mass transport in porous media,
respectively [212,213]. Increasing the tortuosity of a structured
adsorbent will improve the mixing of gases, but will also increase
the pressure drop. Highly ordered structures such as honeycombs
and laminar structures have a tortuosity of 1, which is lower than
the tortuosity values of 23 commonly assigned to packed beds
and irregularly structured adsorbents. In the most fundamental
sense, macropores are needed in adsorbents to facilitate the con-
tact between the adsorbents microparticles and the CO
2
. Struc-
tured adsorbents have certain advantages over packed beds,
since their working capacities are typically higher [198,200].
This example highlights the difculties in quantifying a com-
plex macroporous structure for simulation purposes. Such quanti-
cations are needed to describe the detailed mesoscale
dependencies in packed beds and structured adsorbents for the
428 N. Hedin et al. / Applied Energy 104 (2013) 418433
post-combustion capture of CO
2
. Dantas et al. [214216] simulated
CO
2
-over-N
2
separation in a PSA xed-bed single column, and as-
sessed the separation purity of CO
2
, the recovery of CO
2
, and the
breakthrough curves for different temperatures, feed times, and
purge times, to match experimental data. They used a model based
on the linear driving force (LDF) approximation for the mass
transfer to consider the energy and momentum balances. The
adsorbents studied included zeolite X, activated carbon, and N-
enriched activated carbon. They assumed that the diffusion of N
2
and CO
2
into the adsorbent was controlled by molecular diffusion
in the macropores, following previous studies by Ruthven et al.
[217,218]. The mass transfer in the column was dependent on
parameters that were characteristic of the macroporous structure;
i.e., the particle diameter, the porosity, the axial dispersion coef-
cient, and an assumed tortuosity value of 2.2 for zeolite X and
activated carbon, and 1.8 for N-enriched activated carbon. The
pressure drop and velocity changes were described using Erguns
equation, which considers the overall structure of the xed bed
in terms of the porosity and particle diameter. The performance
of the (macroporous) structure was thus only related to material-
specic and constant parameters, and not directly to the structure
itself. Only the tortuosity could be said to describe the architecture
of the structure, but this parameter was chosen more or less
arbitrarily.
4.2. Supported adsorbents
Supported adsorbents consist of an active adsorbent supported
(typically as a coating) by a matrix that provides the mechanical
strength and the desired macroscopic structure; an example of
such a matrix is a honeycomb structure. The production of sup-
ported adsorbents has strong similarities with that of heteroge-
neous catalysts, where the active catalytic material is distributed
onto the surface of a ceramic support [219,220]. The manufactur-
ing of supported adsorbents typically proceeds via two steps,
where the matrix or support is commonly produced using tradi-
tional ceramic processing techniques that involve extrusion [221]
followed by sintering [222]. The adsorbent can then be applied to
the surfaces of the supports using various techniques, such as
deposition from solgel or hydrothermal solutions, or powder
deposition. Recent examples include the method used by Mosca
et al. [223], who applied NaX coatings to cordierite supports using
a hydrothermal approach originally developed for membrane
lms. They showed that the use of these supported adsorbents in
PSA systems resulted in smaller pressure drops than when regular
adsorbents were used. Rezaei et al. [201] deposited lms of zeolite
X of varying thickness on supports based on cordierite monoliths
with a cell density of 1200 cells per square inch (cpsi).
Carbon supports are also of interest, as they have low density
and relatively high heat conduction. Wu et al. [224] supported
nanosized calcium oxide particles on mesoporous carbon, and
evaluated their capacity for the adsorption of CO
2
at temperatures
of 200500 C. These CaO/C composites, activated at a temperature
of 700 C for 1 h in an N
2
atmosphere, showed the capacity to ad-
sorb up to 7 mmol/g of CO
2
when subjected to an atmosphere of
100% CO
2
for 2 h. Unfortunately, the CaO nanocrystallites grew in
size after repeated adsorptiondesorption cycles. A grain-growth-
related reduced capacity is common for the CaCO
3
/CaO/CO
2
system
used in CLC, which will not be reviewed further here.
5. Conclusions
We have presented an overview of the signicant results for the
capture of CO
2
from ue gases using sorbents. The sorbents were
categorized as microporous and mesoporous, and as chemisor-
bents and physisorbents. Many microporous solids show consider-
able potential as sorbents for CO
2
, especially those with small
pores and a high isosteric heat of physisorption. Certain MOFs also
display advantageous properties, but it is unclear whether these
properties are signicantly better than those of other materials.
It is likely that the enormous ue gas ows in power plants cannot
be compressed, which impacts the choice of adsorbent. Physisor-
bents with small micropores are likely more suitable for the
post-combustion capture of CO
2
than mesoporous sorbents. VSA
or TSA with rapid cycles make physisorbents with large pores less
useful. The study of the kinetic selection of CO
2
over N
2
has begun,
and we expect many more studies of such adsorbents in the future.
Many of the studied sorbents are sensitive to water in the ue
gases, and we expect this to be the subject of further study. VSA
or TSA would most likely be used after the water condensation
step, but the amount of water in the ue gases would still be sig-
nicant. Previous studies suggest that microporous organic, pure
microporous silica, and aluminumphosphate sorbents are less sen-
sitive to water than are traditional zeolites, and most MOFs. We ex-
pect more studies to be performed on microporous polymers to
investigate their suitability as CO
2
sorbents. In general, they are
more water-stable than MOFs, and it might be possible to integrate
them in polymer membranes or MMMs.
Amine-modied mesoporous solids also show considerable po-
tential as chemisorbents for CO
2
. They show high CO
2
-over-N
2
selectivity and are robust towards water. In analogy with liquid
amines, supported amines could be sensitive to impurities such
as SO
2
and COS. The latter has not been considered in detail by
the research community. Studies have indicated that secondary
amines are sensitive towards oxidative degradation, which could
limit the possibilities of using supported TEPA and PEI. We expect
additional studies on the eventual degradation of secondary
amines in representative ue gas mixtures at temperatures rele-
vant for the post-combustion capture of CO
2
. We anticipate ongo-
ing investigations of both mesostructured silica and unordered
mesoporous silica substrates as substrates for amines. For actual
applications in the ue gas capture of CO
2
, the inexpensive and
unordered silica substrates are likely to be most relevant.
We reviewed the literature on structured adsorbents for CO
2
.
Traditionally, adsorption-driven gas separation has been particu-
larly attractive for midsized separation units, because the swing cy-
cles are 10 min long; however, such cycle times are not
compatible with the enormous ows of ue gases in large power
plants. Additional developments in rapid swing cycles for the sepa-
ration of CO
2
from ue gases are forthcoming. Such rapid cycles re-
quire structured adsorbents that are mechanically robust. It should
be noted that structured adsorbents could allow for rapid and more
cost effective adsorption-driven gas separation, but such processes
must be analyzed in combination with the full thermodynamics of
the power plant. Heat integration is expected to be crucial.
We expect more research on the system integration of adsorp-
tionseparation units for the post-combustion capture of CO
2
in
power plants. Such integration must be analyzed while considering
the full thermodynamics, including those of the adsorption-driven
capture of CO
2
. For example, it is essential to study how trace
amounts of various components could interfere with the adsorp-
tion processes and the overall thermodynamics. The parasitic con-
tributions associated with the use of chemisorbents or
physisorbents should be further studied. It seems that the heat of
adsorption should be not too small, nor too large.
The development of adsorbents with high capacity, high selec-
tivity, rapid uptake, easy recycling, and suitable thermal and
mechanical properties is a challenging task. Truly interdisciplinary
studies could enable this delicate optimization to be performed in
the near future.
N. Hedin et al. / Applied Energy 104 (2013) 418433 429
References
[1] Hester RE, Harrision RM, editors. Carbon capture sequestration and
storage. Cambridge, UK: The Royal Society of Chemistry; 2010.
[2] Saundry P, editor. Carbon capture and sequestration program of the US
department of energy. Environmental information coalition, national council
for science and the environment, Washington, DC; 2009.
[3] Philibert C. Technology penetration and capital stock turnover-lessons from
IEA scenario analysis. OECD and IEA information paper: COM/ENV/EPOC/IEA/
SLT. International Energy Agency, Paris; 2007, vol. 4.
[4] Haszeldine RS. Carbon capture and storage: how green can black be? Science
2009;325:164752.
[5] Chen L, Yong SZ, Ghoniem AF. Oxy-fuel combustion of pulverized coal:
Characterization, fundamentals, stabilization and CFD modeling. Prog Energy
Combust Sci 2012;38:156214.
[6] Metz B, Davidson OR, de Conick HC, Loos M, Mayer LA. IPCC special report on
carbon dioxide capture and storage. Cambridge, UK: Cambridge University
Press; 2005.
[7] Aaron D, Tsouris C. Separation of CO
2
from ue gas: a review. Sep Sci Technol
2005;40:32148.
[8] Isaacs EE, Otto FD, Mather AE. Solubility of mixtures of H
2
S and CO
2
in a
monoethanolamine solution at low partial pressures. J Chem Eng Data
1980;25:11820.
[9] Mago BF, West CW. Corrosion inhibitors for alkanolamine gas treating
system. US patent 3,959,170; 1976.
[10] Mamun S, Svendsen HF, Hoff KA, Juliussen O. Selection of new absorbents for
carbon dioxide capture. Energy Convers Manage 2007;48:2518.
[11] Feron PHM. Exploring the potential for improvement of the energy
performance of coal red power plants with post-combustion capture of
carbon dioxide. Int J Greenh. Gas Control 2010;4:15260.
[12] Baxter J, Bian Z, Chen G, Danielson D, Dresselhaus MS, Fedorov AG, et al.
Nanoscale design to enable the revolution in renewable energy. Energy
Environ Sci 2009;2:55988.
[13] Ho MT, Allinson GW, Wiley DE. Reducing the cost of CO
2
capture from ue
gases using pressure swing adsorption. Ind Eng Chem Res 2008;47:
488390.
[14] Pipitone G, Bolland O. Power generation with CO
2
capture: technology for CO
2
purication. Int J Greenh. Gas Control 2009;3:52834.
[15] Li H, Ji X, Yan J. A new modication on RK EOS for gaseous CO
2
and gaseous
mixtures of CO
2
and H
2
O. Int J Energy Res 2006;30:13548.
[16] Li H, Yan J. Impacts of equations of state (EOS) and impurities on the volume
calculation of CO
2
mixtures in the applications of CO
2
capture and storage
(CCS) processes. Appl Energy 2009;86:276070.
[17] Li H, Yan J. Evaluating cubic equations of state for calculation of vaporliquid
equilibrium of CO
2
and CO
2
-mixtures for CO
2
capture and storage processes.
Appl Energy 2009;86:82636.
[18] Li H, Yan J, Yan J, Anheden M. Impurity impacts on the purication process in
oxy-fuel combustion based CO
2
capture and storage system. Appl Energy
2009;86:20213.
[19] Yong Z, Mata V, Rodrigues AE. Adsorption of carbon dioxide at high
temperature a review. Sep Purif Technol 2002;26:195205.
[20] Lee KB, Beaver MG, Caram HS, Sircar S. Reversible chemisorbents for carbon
dioxide and their potential applications. Ind Eng Chem Res 2008;47:804862.
[21] White CM, Strazisar BR, Granite EJ, Hoffman JS, Pennline HW. Separation and
capture of CO
2
from large stationary sources and sequestration in geological
formations Coalbeds and deep saline aquifers. J Air Waste Manage Assoc
2003;53:645715.
[22] Choi S, Drese JH, Jones CW. Adsorbent materials for carbon dioxide capture
from large anthropogenic point sources. ChemSusChem 2009;2:796854.
[23] Hedin N, Chen LJ, Laaksonen A. Sorbents for CO
2
capture from ue gas
aspects from materials and theoretical chemistry. Nanoscale 2010:181941.
[24] Ruthven DM. Principles of adsorption and adsorption processes. New
York: John Wiley & Sons, Inc.; 1984.
[25] Chaffee AL, Knowles GP, Liang Z, Zhang J, Xiao P, Webley PA. CO
2
capture by
adsorption: materials and process development. Int J Greenh Gas Control
2007;1:118.
[26] Derouane EG. Zeolites as solid solvents. J Mol Catal A Chem 1998;134:
2945.
[27] Reid RC, Prausnitz JM, Sherwood TK. The properties of gases and liquids. New
York: McGraw-Hill; 1977.
[28] Yang RT. Gas separation by adsorption processes. London: Imperical College
Press; 1997.
[29] Breck DW. Zeolite molecular sieves. Malabar: Robert E. Krieger Pub. Co.; 1974.
[30] Graham C, Imrie DA, Raab RE. Measurement of the electric quadrupole
moments of CO
2
, CO., N
2
, Cl
2
and BF
3
. Mol Phys 1998;93:4956.
[31] Cheung O, Liu Q, Bacsik Z, Hedin N. Silicoaluminophosphates as CO
2
sorbents.
Microporous Mesoporous Mater 2012;156:906.
[32] Liu Q, Mace A, Bacsik Z, Sun J, Laaksonen A, Hedin N. NaKA sorbents with high
CO
2
-over-N
2
selectivity and high capacity to adsorb CO
2
. Chem Commun
2010;46:45024.
[33] Palomino M, Corma A, Jorda JL, Rey F, Valencia S. Zeolite Rho: a highly
selective adsorbent for CO
2
/CH
4
separation induced by a structural phase
modication. Chem Commun 2012;48:2157.
[34] Liu Q, Cheung NCO, Garcia-Bennett AE, Hedin N. Aluminophosphates for CO
2
separation. ChemSusChem 2011;4:917.
[35] Britt D, Furukawa H, Wang B, Glover TG, Yaghi OM. Highly efcient
separation of carbon dioxide by a metal-organic framework replete with
open metal sites. Proc Natl Acad Sci USA 2009;106:2063740.
[36] Luebke R, Eubank JF, Cairns AJ, Belmabkhout Y, Wojtas L, Eddaoudi M. The
unique rht-MOF platform, ideal for pinpointing the functionalization and CO
2
adsorption relationship. Chem Commun 2012;48:14557.
[37] Gassensmith JJ, Furukawa H, Smaldone RA, Forgan RS, Botros YY, Yaghi OM,
et al. Strong and reversible binding of carbon dioxide in a green metal organic
framework. J Am Chem Soc 2011;133:153125.
[38] Yao Q, Su J, Cheung O, Liu Q, Hedin N, Zou X. Interpenetrated metal-organic
frameworks and their uptake of CO
2
at relatively low pressures. J Mater Chem
2012;22:1034551.
[39] Silvestre-Albero J, Wahby A, Seplveda-Escribano A, Martnez-Escandell M,
Kaneko K, Rodrguez-Reinoso F. Ultrahigh CO
2
adsorption capacity on carbon
molecular sieves at room temperature. Chem Commun 2011;47:68402.
[40] Laybourn A, Dawson R, Clowes R, Iggo JA, Cooper AI, Khimyak YZ, et al.
Branching out with aminals: microporous organic polymers from
difunctional monomers. Polym Chem 2012;3:5337.
[41] Rabbani MG, El-Kaderi HM. Synthesis and characterization of porous
benzimidazole-linked polymers and their performance in small gas storage
and selective uptake. Chem Mater 2012;24:15117.
[42] Kiskan B, Antonietti M, Weber J. Teaching new tricks to an old indicator: pH-
switchable, photoactive microporous polymer networks from
phenolphthalein with tunable CO
2
adsorption power. Macromolecules
2012;45:135661.
[43] Furukawa H, Yaghi OM. Storage of hydrogen, methane, and carbon dioxide in
highly porous covalent organic frameworks for clean energy applications. J
Am Chem Soc 2009;131:887583.
[44] Belmabkhout Y, Sayari A. Effect of pore expansion and amine
functionalization of mesoporous silica on CO
2
adsorption over a wide range
of conditions. Adsorpt 2009;15:31828.
[45] Bacsik Z, Ahlsten N, Ziadi A, Zhao G, Garcia-Bennett AE, Martn-Matute B,
et al. Mechanisms and kinetics for sorption of CO
2
on bicontinuous
mesoporous silica modied with n-propylamine. Langmuir
2011;27:1111828.
[46] Si X, Jiao C, Li F, Zhang J, Wang S, Liu S, et al. High and selective CO
2
uptake, H
2
storage and methanol sensing on the amine-decorated 12-connected MOF
CAU-1. Energy Environ Sci 2011;4:45227.
[47] Lin L-C, Berger AH, Martin RL, Kim J, Swisher JA, Jariwala K, et al. In silico
screening of carbon-capture materials. Nat Mater 2012;11:63341.
[48] Baerlocher C, McCusker LB. Database of zeolite structures. <http://www.iza-
structure.org/databases/>.
[49] Zhang Z, Zhang W, Chen X, Xia Q, Li Z. Adsorption of CO
2
on zeolite 13X and
activated carbon with higher surface area. Sep Sci Technol 2010;45:7109.
[50] da Silva FWM, Maia DAS, Oliveira RS, Moreno-Pirajn JC, Sapag K, Cavalcante
CL, et al. Adsorption microcalorimetry applied to the characterisation of
adsorbents for CO
2
capture. Can J Chem Eng 2012. http://dx.doi.org/10.1002/
cjce.21692.
[51] Zimmermann W, Keller JU. A new calorimeter for simultaneous measurement
of isotherms and heats of adsorption. Thermochim Acta 2003;405:3141.
[52] Dunne JA, Mariwals R, Rao M, Sircar S, Gorte RJ, Myers AL. Calorimetric heats
of adsorption and adsorption isotherms. 1. O
2
, N
2
, Ar, CO
2
, CH
4
, C
2
H
6
and SF
6
on silicalite. Langmuir 1996;12:588895.
[53] Wang Y, LeVan MD. Adsorption equilibrium of binary mixtures of carbon
dioxide and water vapor on zeolites 5A and 13X. J Chem Eng Data
2010;55:318995.
[54] Brandani F, Ruthven DM. The effect of water on the adsorption of CO
2
and
C
3
H
8
on type X zeolites. Ind Eng Chem Res 2004;43:833944.
[55] Dragan G. The individual adsorption of carbon dioxide and sulphur dioxide by
Y zeolites. Rev Chim 2010;61:897902.
[56] Yan J, Yu D, Li H, Sun P, Huang H. NaY zeolites modied by La
3+
and Ba
2+
: the
effect of synthesis details on surface structure and catalytic performance for
lactic acid to acrylic acid. J Rare Earth 2010;28:8036.
[57] Pirngruber GD, Raybaud P, Belmabkhout Y, Cejka J, Zukal A. The role of the
extra-framework cations in the adsorption of CO
2
on faujasite Y. Phys Chem
Chem Phys 2010;12:1353446.
[58] Bendenia S, Marouf-Khelifa K, Batonneau-Gener I, Derriche Z, Khelifa A.
Adsorptive properties of X zeolites modied by transition metal cation
exchange. Adsorpt 2011;17:36170.
[59] Breck DW, Eversole WG, Milton RM, Reed TB, Thomas TL. Crystalline zeolites.
1. The properties of a new synthetic zeolite, type-A. J Am Chem Soc
1956;78:596371.
[60] Yeh YT, Yang RT. Diffusion in zeolites containing mixed cations. AICHE J
1989;35:165966.
[61] Akhtar F, Liu Q, Hedin N, Bergstrom L. Strong and binder free structured
zeolite sorbents with very high CO
2
-over-N
2
selectivities and high capacities
to adsorb CO
2
rapidly. Energy Environ Sci 2012;5:766473.
[62] Montanari T, Finocchio E, Salvatore E, Garuti G, Giordano A, Pistarino C, et al.
CO
2
separation and landll biogas upgrading: a comparison of 4A and 13X
zeolite adsorbents. Energy 2011;36:3149.
[63] Tomadakis MM, Heck HH, Jubran ME, Al-Harthi K. Pressure-swing adsorption
separation of H
2
S from CO
2
with molecular sieves 4A, 5A, and 13X. Sep Sci
Technol 2011;46:42833.
[64] Bulnek R, Frolich K, Fry dova E, C

icmanec P. Microcalorimetric and FTIR study


of the adsorption of carbon dioxide on alkali-metal exchanged FER zeolites.
Top Catal 2010;53:134960.
430 N. Hedin et al. / Applied Energy 104 (2013) 418433
[65] Ridha FN, Webley PA. Entropic effects and isosteric heats of nitrogen and
carbon dioxide adsorption on chabazite zeolites. Microporous Mesoporous
Mater 2010;132:2230.
[66] Ridha FN, Webley PA. Investigation of the possibility of low pressure
encapsulation of carbon dioxide in potassium chabazite (KCHA) and sodium
chabazite (NaCHA) zeolites. J Colloid Interface Sci 2009;337:3327.
[67] Ridha FN, Webley PA. Anomalous Henrys law behavior of nitrogen and
carbon dioxide adsorption on alkali-exchanged chabazite zeolites. Sep Purif
Technol 2009;67:33643.
[68] Hudson MR, Queen WL, Mason JA, Fickel DW, Lobo RF, Brown CM.
Unconventional, highly selective CO
2
adsorption in zeolite SSZ-13. J Am
Chem Soc 2012;134:19703.
[69] Park SW, Yun YH, Kim SD, Yang ST, Ahn WS, Seo G, et al. CO
2
retention ability
on alkali cation exchanged titanium silicate, ETS-10. J Porous Mater
2010;17:58995.
[70] Lively RP, Chance RR, Koros WJ. Enabling low-cost CO
2
capture via heat
integration. Ind Eng Chem Res 2010;49:755062.
[71] Lively RP, Chance RR, Kelley BT, Deckman HW, Drese JH, Jones CW, et al.
Hollow ber adsorbents for CO
2
removal from ue gas. Ind Eng Chem Res
2009;48:731424.
[72] Chung T, Jiang LY, Li Y, Kulprathipanja S. Mixed matrix membranes (MMMs)
comprising organic polymers with dispersed inorganic llers for gas
separation. Prog Polym Sci 2007;32:483507.
[73] Bernardo P, Drioli E, Golemme G. Membrane gas separation: a review/state of
the art. Ind Eng Chem Res 2009;48:463863.
[74] Yang S, Kim J, Ahn W. CO
2
adsorption over ion-exchanged zeolite beta with
alkali and alkaline earth metal ions. Microporous Mesoporous Mater
2010;135:904.
[75] Zukal A, Mayerov J, Kubu M. Adsorption of carbon dioxide on high-silica
zeolites with different framework topology. Top Catal 2010;53:13616.
[76] Wilson ST, Lok BM, Messina CA, Cannan TR, Flanigen EM. Aluminophosphate
molecular-sieves a new class of microporous crystalline inorganic solids. J
Am Chem Soc 1982;104:11467.
[77] Lok BM, Messina CA, Patton RL, Gajek RT, Cannan TR, Flanigen EM.
Silicoaluminophosphate molecular-sieves another new class of
microporous crystalline inorganic solids. J Am Chem Soc 1984;106:60923.
[78] Zhang L, Primera-Pedrozo JN, Hernndez-Maldonado AJ. Thermal
detemplation of Na-SAPO-34: effect on Sr
2+
ion exchange and CO
2
adsorption. J Phys Chem C 2010;114:1475562.
[79] Li S, Fan CQ. High-ux SAPO-34 membrane for CO
2
/N
2
separation. Ind Eng
Chem Res 2010;49:4399404.
[80] Ruthven DM, Shamsuzzaman F, Knaebel KS. Pressure swing adsorption. New
York: John Wiley and Sons, Inc.; 1994.
[81] Ruthven DM, Reyes SC. Adsorptive separation of light olens from parafns.
Microporous Mesoporous Mater 2007;104:5966.
[82] Alcaiz-Monge J, Marco-Lozar JP, Lillo-Rdenas MA. CO
2
separation by carbon
molecular sieve monoliths prepared from nitrated coal tar pitch. Fuel Process
Technol 2011;92:9159.
[83] Alcaiz-Monge J, Marco-Lozar JP, Lozano-Castell D. Monolithic carbon
molecular sieves from activated bituminous coal impregnated with a slurry
of coal tar pitch. Fuel Process Technol 2012;95:6772.
[84] Bikshapathi M, Sharma A, Sharma A, Verma N. Preparation of carbon
molecular sieves from carbon micro and nanobers for sequestration of
CO
2
. Chem Eng Res Des 2011;89:173746.
[85] Wahby A, Ramos-Fernndez JM, Martnez-Escandell M, Sepveda-Escribano
A, Silvestre-Albero J, Rodrguez-Reinoso F. High-surface-area carbon
molecular sieves for selective CO
2
adsorption. ChemSusChem 2010;3:97481.
[86] Bae Y, Snurr RQ. Development and evaluation of porous materials for carbon
dioxide separation and capture. Angew Chem Int Ed 2011;50:1158696.
[87] Banerjee R, Furukawa H, Britt D, Knobler C, OKeeffe M, Yaghi OM. Control of
pore size and functionality in isoreticular zeolitic imidazolate frameworks
and their carbon dioxide selective capture properties. J Am Chem Soc
2009;131:38757.
[88] Bae Y, Spokoyny AM, Farha OK, Snurr RQ, Hupp JT, Mirkin CA. Separation of
gas mixtures using Co(II) carborane-based porous coordination polymers.
Chem Commun 2010;46:347880.
[89] Dietzel PDC, Besikiotis V, Blom R. Application of metal-organic frameworks
with coordinatively unsaturated metal sites in storage and separation of
methane and carbon dioxide. J Mater Chem 2009;19:736270.
[90] Demessence A, DAlessandro DM, Foo ML, Long JR. Strong CO
2
Binding in a
water-stable, triazolate-bridged metal-organic framework functionalized
with ethylenediamine. J Am Chem Soc 2009;131:87846.
[91] Bae Y, Farha OK, Hupp JT, Snurr RQ. Enhancement of CO
2
/N
2
selectivity in a
metal-organic framework by cavity modication. J Mater Chem
2009;19:21314.
[92] Debatin F, Thomas A, Kelling A, Hedin N, Bacsik Z, Senkovska I, et al. In situ
synthesis of an imidazolate-4-amide-5-imidate ligand and formation of a
microporous zinc-organic framework with H
2
-and CO
2
-Storage Ability.
Angew Chem Int Ed 2010;49:125862.
[93] McDonald TM, Lee WR, Mason JA, Wiers BM, Hong CS, Long JR. Capture of
carbon dioxide from air and ue gas in the alkylamine-appended metal-
organic framework mmen-Mg
2
(dobpdc). J Am Chem Soc 2012;134:705665.
[94] Devic T, Horcajada P, Serre C, Salles F, Maurin G, Moulin B, et al.
Functionalization in exible porous solids: effects on the pore opening and
the host-guest interactions. J Am Chem Soc 2010;132:112736.
[95] Devic T, Salles F, Bourrelly S, Moulin B, Maurin G, Horcajada P, et al. Effect of
the organic functionalization of exible MOFs on the adsorption of CO
2
. J
Mater Chem 2012;22:1026673.
[96] Thomas A. Functional materials: fromhard to soft porous frameworks. Angew
Chem Int Ed 2010;49:832844.
[97] Yuan D, Lu W, Zhao D, Zhou H. Highly stable porous polymer networks with
exceptionally high gas-uptake capacities. Adv Mater 2011;23:37235.
[98] Ben T, Ren H, Shengqian M, Cao D, Lan J, Jing X, et al. Targeted synthesis of a
porous aromatic framework with high stability and exceptionally high
surface area. Angew Chem Int Ed 2009;48:945760.
[99] Dawson R, Stckel E, Holst JR, Adams DJ, Cooper AI. Microporous
organic polymers for carbon dioxide capture. Energy Environ Sci 2011;4:
423945.
[100] Laybourn A, Dawson R, Clowes R, Iggo JA, Cooper AI, Khimyak YZ, et al.
Branching out with aminals: microporous organic polymers from
difunctional monomers. Polym Chem 2012;3:5337.
[101] Belmabkhout Y, Sayari A. Adsorption of CO
2
from dry gases on MCM-41 silica
at ambient temperature and high pressure. 2: Adsorption of CO
2
/N
2
, CO
2
/CH
4
and CO
2
/H
2
binary mixtures. Chem Eng Sci 2009;64:372935.
[102] Belmabkhout Y, Serna-Guerrero R, Sayari A. Adsorption of CO
2
from dry gases
on MCM-41 silica at ambient temperature and high pressure. 1: Pure CO
2
adsorption. Chem Eng Sci 2009;64:37218.
[103] Ma Y, Zhao H, Tang S, Hu J, Liu H. Synthesis of micro/mesoporous composites
and their application as CO
2
adsorbents. Acta Phys Chim Sin 2011;27:
68996.
[104] Siriwardane RV, Shen MS, Fisher EP, Poston JA. Adsorption of CO
2
on
molecular sieves and activated carbon. Energy Fuels 2001;15:27984.
[105] Dantas TLP, Amorim SM, Luna FMT, Silva Jr IJ, de Azevedo DCS, Rodrigues AE,
et al. Adsorption of carbon dioxide onto activated carbon and nitrogen-
enriched activated carbon: surface changes, equilibrium, and modeling of
xed-bed adsorption. Sep Sci Technol 2010;45:7384.
[106] An H, Feng B, Su S. CO
2
capture capacities of activated carbon bre-phenolic
resin composites. Carbon 2009;47:2396405.
[107] Su F, Lu C, Kuo S, Zeng W. Adsorption of CO
2
on amine-functionalized Y-type
zeolites. Energy Fuels 2010;24:14418.
[108] Ye Q, Jiang J, Wang C, Liu Y, Pan H, Shi Y. Adsorption of low-concentration
carbon dioxide on amine-modied carbon nanotubes at ambient
temperature. Energy Fuels 2012;26:2497504.
[109] Pacheco DM, Johnson JR, Koros WJ. Aminosilane-functionalized cellulosic
polymer for increased carbon dioxide sorption. Ind Eng Chem Res
2012;51:50314.
[110] Couck S, Gobechiya E, Kirschhock CEA, Serra-Crespo P, Juan-Alcaiz J,
Martinez Joaristi A, et al. Adsorption and separation of light gases on an
amino-functionalized metal organic framework: an adsorption and in situ
XRD study. ChemSusChem 2012;5:74050.
[111] Bollini P, Didas SA, Jones CW. Amine-oxide hybrid materials for acid gas
separations. J Mater Chem 2011;21:1510020.
[112] Vaidhyanathan R, Iremonger SS, Shimizu GKH, Boyd PG, Alavi S, Woo TK.
Direct observation and quantication of CO
2
binding within an amine-
functionalized nanoporous solid. Science 2010;330:6503.
[113] Hoerr CW, Harwood HJ, Ralston AW. J Org Chem 1944;9:20110.
[114] Alauzun J, Besson E, Mehdi A, Reye C, Corriu R. Reversible covalent chemistry
of CO
2
: an opportunity for nano-structured hybrid organic-inorganic
materials. Chem Mater 2008;20:50313.
[115] Masuda K, Ito Y, Horiguchi M, Fujita H. Studies on the solvent dependence of
the carbamic acid formation from x-(1-naphthyl)alkylamines and carbon
dioxide. Tetrahedron 2005;61:21329.
[116] Goeppert A, Meth S, Prakash GKS, Olah GA. Nanostructured silica as a support
for regenerable high-capacity organoamine-based CO
2
sorbents. Energy
Environ Sci 2010;3:194960.
[117] Li W, Bollini P, Didas SA, Choi S, Drese JH, Jones CW. Structural changes of
silica mesocellular foam supported amine-functionalized CO
2
adsorbents
upon exposure to steam. ACS Appl Mater Interfaces 2010;2:336372.
[118] Belmabkhout Y, Sayari A. Isothermal versus non-isothermal adsorption-
desorption cycling of triamine-grafted pore-expanded MCM-41 mesoporous
silica for CO
2
capture from ue gas. Energy Fuels 2010;24:527380.
[119] Franchi RS, Harlick PJE, Sayari A. Applications of pore-expanded mesoporous
silica. 2. Development of a high-capacity, water-tolerant adsorbent for CO
2
.
Ind Eng Chem Res 2005;44:800713.
[120] Harlick PJE, Sayari A. Amine grafted, pore-expanded MCM-41 for acid gas
removal: Effect of grafting temperature, water, and amine type on
performance. In: Cejka J, Zilkova N, Nachtigall P, editors. Molecular sieves:
from basic research to industrial applications, pts B; 2005, p. 9874.
[121] Harlick PJE, Sayari A. Applications of pore-expanded mesoporous silicas. 3.
Triamine silane grafting for enhanced CO
2
adsorption. Ind Eng Chem Res
2006;45:324855.
[122] Harlick PJE, Sayari A. Applications of pore-expanded mesoporous silica. 5.
Triamine grafted material with exceptional CO
2
dynamic and equilibrium
adsorption performance. Ind Eng Chem Res 2007;46:44658.
[123] Serna-Guerrero R, Belmabkhout Y, Sayari A. Triamine-grafted pore-expanded
mesoporous silica for CO
2
capture: effect of moisture and adsorbent
regeneration strategies. Adsorpt 2010:19.
[124] Pinto ML, Mafra L, Guil JM, Pires J, Rocha J. Adsorption and activation of CO
2
by amine-modied nanoporous materials studied by solid-state NMR and
13
CO
2
adsorption. Chem Mater 2011;23:138795.
N. Hedin et al. / Applied Energy 104 (2013) 418433 431
[125] Danon A, Stair PC, Weitz E. FTIR study of CO
2
adsorption on amine-grafted
SBA-15: elucidation of adsorbed species. J Phys Chem C 2011;115:
115409.
[126] Bacsik Z, Atluri R, Garcia-Bennett AE, Hedin N. Temperature-induced uptake
of CO
2
and formation of carbamates in mesocaged silica modied with n-
propylamines. Langmuir 2010;26:1001324.
[127] Aziz B, Hedin N, Bacsik Z. Quantication of chemisorption and physisorption
of carbon dioxide on porous silica modied by propylamines: effect of amine
density. Microporous Mesoporous Mater 2012;159:429.
[128] Hiyoshi N, Yogo K, Yashima T. Adsorption characteristics of carbon dioxide on
organically functionalized SBA-15. Microporous Mesoporous Mater
2005;84:35765.
[129] Knfel C, Martin C, Hornebecq V, Llewellyn PL. Study of carbon dioxide
adsorption on mesoporous aminopropylsilane- functionalized silica and
titania combining microcalorimetry and in situ infrared spectroscopy. J
Phys Chem C 2009;113:2172634.
[130] Xu XC, Song CS, Miller BG, Scaroni AW. Adsorption separation of carbon
dioxide from ue gas of natural gas-red boiler by a novel nanoporous
molecular basket adsorbent. Fuel Process Technol 2005;86:145772.
[131] Aziz B, Zhao G, Hedin N. Carbon dioxide sorbents with propylamine groups-
silica functionalized with a fractional factorial design approach. Langmuir
2011;27:382234.
[132] Young PD, Notestein JM. The role of amine surface density in carbon dioxide
adsorption on functionalized mixed oxide surfaces. ChemSusChem
2011;4:16718.
[133] Knowles GP, Graham JV, Delaney SW, Chaffee AL. Aminopropyl-
functionalized mesoporous silicas as CO
2
adsorbents. Fuel Process Technol
2005;86:143548.
[134] Kim M, Park J-W. Reversible, solid state capture of carbon dioxide by
hydroxylated amidines. Chem Commun 2010;46:25079.
[135] Belmabkhout Y, Serna-Guerrero R, Sayari A. Amine-bearing mesoporous
silica for CO
2
removal from dry and humid air. Chem Eng Sci 2010;65:
36958.
[136] Belmabkhout Y, Serna-Guerrero R, Sayari A. Adsorption of CO
2
-containing gas
mixtures over amine-bearing pore-expanded MCM-41 silica: application for
gas purication. Ind Eng Chem Res 2010;49:35965.
[137] Serna-Guerrero R, Belmabkhout Y, Sayari A. Modeling CO
2
adsorption on
amine-functionalized mesoporous silica: 1. A semi-empirical equilibrium
model. Chem Eng J 2010;161:17381.
[138] Serna-Guerrero R, Sayari A. Modeling CO
2
adsorption on amine-
functionalized mesoporous silica: 2. Kinetics and breakthrough curves.
Chem Eng J 2010;161:18290.
[139] Hicks JC, Drese JH, Fauth DJ, Gray ML, Qi G, Jones CW. Designing adsorbents
for CO
2
capture from ue gas-hyperbranched aminosilicas capable, of
capturing CO
2
reversibly. J Am Chem Soc 2008;130:29023.
[140] Rosenholm JM, Penninkangas A, Linden M. Amino-functionalization of large-
pore mesoscopically ordered silica by a one-step hyperbranching
polymerization of a surface-grown polyethyleneimine. Chem Commun
2006:390911.
[141] Drese JH, Choi S, Lively RP, Koros WJ, Fauth DJ, Gray ML, et al. Synthesis-
structure-property relationships for Hyperbranched aminosilica CO
2
adsorbents. Adv Funct Mater 2009;19:382132.
[142] Choi S, Drese JH, Eisenberger PM, Jones CW. Application of amine-tethered
solid sorbents for direct CO
2
capture from the ambient air. Environ Sci
Technol 2011;45:24207.
[143] Satyapal S, Filburn T, Trela J, Strange J. Performance and properties of a solid
amine sorbent for carbon dioxide removal in space life support applications.
Energy Fuels 2001;15:2505.
[144] Chen C, Son W, You K, Ahn J, Ahn W. Carbon dioxide capture using amine-
impregnated HMS having textural mesoporosity. Chem Eng J
2010;161:4652.
[145] Chen C, Yang S, Ahn W, Ryoo R. Amine-impregnated silica monolith with a
hierarchical pore structure: enhancement of CO
2
capture capacity. Chem
Commun 2009:36279.
[146] Liu Y, Shi J, Chen J, Ye Q, Pan H, Shao Z, et al. Dynamic performance of CO
2
adsorption with tetraethylenepentamine-loaded KIT-6. Microporous
Mesoporous Mater 2010;134:1621.
[147] Wang X, Li H, Liu H, Hou X. As-synthesized mesoporous silica MSU-1
modied with tetraethylenepentamine for CO
2
adsorption. Microporous
Mesoporous Mater 2011;142:5649.
[148] Tanthana J, Chuang SSC. In situ infrared study of the role of PEG in stabilizing
silica-supported amines for CO
2
capture. ChemSusChem 2010;3:95764.
[149] Ebner AD, Gray ML, Chisholm NG, Black QT, Mumford DD, Nicholson MA,
et al. Suitability of a solid amine sorbent for CO
2
capture by pressure swing
adsorption. Ind Eng Chem Res 2011;50:563441.
[150] Jiang B, Kish V, Fauthb Daniel J, Grayb McMahan L, Pennlineb Henry W, Li
Bingyun. Performance of amine-multilayered solid sorbents for CO
2
removal:
effect of fabrication variables. Int J Greenh Gas Control 2011;5:11705.
[151] Wang J, Long D, Zhou H, Chen Q, Liu X, Ling L. Surfactant promoted solid
amine sorbents for CO
2
capture. Energy Environ Sci 2012;5:57429.
[152] Qi G, Wang Y, Estevez L, Duan X, Anako N, Park A, et al. High efciency
nanocomposite sorbents for CO
2
capture based on amine-functionalized
mesoporous capsules. Energy Environ Sci 2011;4:44452.
[153] Bollini P, Choi S, Drese JH, Jones CW. Oxidative degradation of aminosilica
adsorbents relevant to postcombustion CO
2
capture. Energy Fuels
2011;25:241625.
[154] Heldebrant DJ, Koech PK, Rainbolt JE, Zheng F. CO
2
-binding organic liquids.
An integrated acid gas capture system. Energy Procedia 2011;4:21623.
[155] Crittenden B, Thomas WJ. Adsorption technology and
design. Oxford: Butterworth-Heinemann; 1998.
[156] Skarstrom CW. Method and apparatus for fractionating. US Patent 2944627;
1960.
[157] Diagne D, Goto M, Hirose T. Parametric studies on CO
2
separation and
recovery by a dual reux PSA process consisting of both rectifying and
stripping sections. Ind Eng Chem Res 1995;34:30839.
[158] Diagne D, Goto M, Hirose T. New PSA process with intermediate feed inlet
position operated with dual reuxes application to carbon dioxide removal
and enrichment. J Chem Eng Jpn 1994;27:859.
[159] Ebner AD, Ritter JA. Equilibrium theory analysis of dual reux PSA for
separation of a binary mixture. AICHE J 2004;50:241829.
[160] Park JH, Beum HT, Kim JN, Cho SH. Numerical analysis on the power
consumption of the PSA process for recovering CO
2
from ue gas. Ind Eng
Chem Res 2002;41:412231.
[161] Fiandaca G, Fraga ES, Brandani S. A multi-objective genetic algorithm for the
design of pressure swing adsorption. Eng Optimz 2009;9:83354.
[162] Thakur RS, Kaistha N, Rao DP. Process intensication in duplex pressure
swing adsorption. Comput Chem Eng 2011;35:97383.
[163] Leavitt FW. Duplex adsorption process, US Patent 5,085,674; 1992.
[164] Franco A, Diaz AR. The future challenges for clean coal technologies: joining
efciency increase and pollutant emission control. Energy 2009;34:34854.
[165] Li H, Yan J. Performance comparison on the evaporative gas turbine cycles
combined with different CO
2
-capture options. Int J Green Energy
2009;6:51226.
[166] Li H, Flores S, Hu Y, Yan J. Simulation and optimization of evaporative gas
turbine with chemical absorption for carbon dioxide capture. Int J Green
Energy 2009;6:52739.
[167] Lisbona P, Marti

nez A, Lara Y, Romeo LM. Integration of carbonate CO


2
capture cycle and coal-red power plants. A comparative study for different
sorbents. Energy Fuels 2010;24:72836.
[168] Fu C, Grundersen T. Heat integration of an oxy-combustion process for coal-
red power plants with CO
2
capture by pinch analysis. Chem Eng Trans
2011;25:58187.
[169] Alabdulkarem A, Hwang Y, Radermacher R. Energy consumption reduction in
CO
2
capturing and sequestration of an LNG plant through process integration
and waste heat utilization. Intern J Greenh Gas Control 2012;10:21528.
[170] Lucquiaud M, Chalmers H, Gibbins J. Capture-ready supercritical coal-red
power plants and exible post-combustion CO
2
capture. Energy Procedia
2009;1:14118.
[171] Steeneveldt R, Berger B, Torp TA. CO
2
capture and storage closing the
knowing-dowing gap. Trans IChemE A 2006;84:73963.
[172] Escosa JM, Romeo LM. Optimizing CO
2
avoided cost by means of repowering.
Appl Energy 2009;86:23518.
[173] Li XN, Hagaman E, Tsouris C, Lee JW. Removal of carbon dioxide from ue gas
by ammonia carbonation in the gas phase. Energy Fuel 2003;17:6974.
[174] Chen CY, Lee KC, Chou CT. Concentration and recovery of carbon dioxide from
ue gas by vacuum swing adsorption. J Chin Inst Chem Eng 2003;34:13542.
[175] Zhang J, Webley PA, Xiao P. Effect of process parameters on power
requirements of vacuum swing adsorption technology for CO
2
capture from
ue gas. Energy Convers Manage 2008;49:34656.
[176] Zhang J, Webley PA. Cycle development and design for CO
2
capture from ue
gas by vacuum swing adsorption. Environ Sci Technol 2008;42:5639.
[177] Delgado JA, Uguina MA, Sotelo JL, gueda VI, Sanz A, Gmez P. Numerical
analysis of CO
2
concentration and recovery from ue gas by a novel vacuum
swing adsorption cycle. Comput Chem Eng 2011;35:10109.
[178] de Visser E, Hendriks C, Barrio M, Mlnvik MJ, de Koeijer G, Liljemark S, et al.
Dynamis CO
2
quality recommendations. Int J Grenh Gas Control
2008;2:47884.
[179] Merel J, Clausse M, Meunier F. Carbon dioxide capture by indirect thermal
swing adsorption using 13X zeolite. EnvironProg 2006;25:32733.
[180] Grande CA, Ribeiro RPPL, Rodrigues AE. CO
2
capture from NGCC power
stations using Electric Swing Adsorption (ESA). Energy Fuel
2009;23:2797803.
[181] Grande CA, Rodrigues AE. Electric swing adsorption for CO
2
removal from ue
gases. Int J Greenh Gas Control 2008;2:194202.
[182] Moate JR, LeVan MD. Temperature swing adsorption compression: Effects of
nonuniform heating on bed efciency. Appl Therm Eng 2010;30:65863.
[183] Dutcher B, Adidharma H, Radosz M. Carbon lter process for ue-gas carbon
capture on carbonaceous Sorbents: steam-aided vacuum swing adsorption
option. Ind Eng Chem Res 2011;50:9696703.
[184] Lopes FVS, Grande CA, Rodrigues AE. Fast-cycling VPSA for hydrogen
purication. Fuel 2012;93:51023.
[185] Harlick PJE, Tezel FH. An experimental adsorbent screening study for CO
2
removal from N
2
. Microporous Mesoporous Mater 2004;76:719.
[186] Gomes VG, Yee KWK. Pressure swing adsorption for carbon dioxide
sequestration from exhaust gases. Sep Purif Technol 2002;28:16171.
[187] Tlili N, Grevillot G, Vallieres C. Carbon dioxide capture and recovery by means
of TSA and/or VSA. Int J Greenh Gas Control 2009;3:51927.
[188] Yoshida M, Ritter JA, Kodama A, Goto M, Hirose T. Enriching reux and
parallel equalization PSA process for concentrating trace components in air.
Ind Eng Chem Res 2003;42:1795803.
[189] Harlick PJE, Tezel FH. Equilibrium analysis of cyclic adsorption processes: CO
2
working capacities with NaY. Sep Sci Technol 2005;40:256991.
432 N. Hedin et al. / Applied Energy 104 (2013) 418433
[190] Phelps DSC, Ruthven DM. An experimental and theoretical study of the
performance of a continuous counter-current adsorber. Adsorpt
2003;9:26574.
[191] Dasgupta S, Nanoti A, Gupta P, Jena D, Goswami AN, Garg MO. Carbon dioxide
removal with mesoporous adsorbents in a single column pressure swing
adsorber. Sep Sci Technol 2009;44:397383.
[192] Wurzbacher JA, Gebald C, Steinfeld A. Separation of CO
2
from air by
temperature-vacuum swing adsorption using diamine-functionalized silica
gel. Energy Environ Sci 2011;4:358492.
[193] Bastin L, Brcia PS, Hurtado EJ, Silva JAC, Rodrigues AE, Chen B. A microporous
metal-organic framework for separation of CO
2
/N
2
and CO
2
/CH
4
by xed-bed
adsorption. J Phys Chem C 2008;112:157581.
[194] Drage TC, Snape CE, Stevens LA, Wood J, Wang J, Cooper AI, et al. Materials
challenges for the development of solid sorbents for post-combustion carbon
capture. J Mater Chem 2012;22:281523.
[195] Rezaei F, Webley P. Structured adsorbents in gas separation processes. Sep
Purif Technol 2010;70:24356.
[196] Pavlov ML, Travkina OS, Kutepov BI. Grained binder-free zeolites: synthesis
and properties. Catal Ind 2012;4:118.
[197] Shirley AI, LaCava AI. PSA performance of densely packed adsorbent beds.
AIChE J 1995;41:138994.
[198] Rezaei F, Webley P. Optimum structured adsorbents for gas separation
processes. Chem Eng Sci 2009;64:518291.
[199] Rezaei F, Grahn M. Thermal management of structured adsorbents in CO
2
capture processes. Ind Eng Chem Res 2012;51:402534.
[200] Rezaei F, Webley PA. Optimal design of engineered gas adsorbents: pore-
scale level. Chem Eng Sci 2012;69:2708.
[201] Rezaei F, Mosca A, Hedlund J, Webley PA, Grahn M, Mouzon J. The effect of
wall porosity and zeolite lm thickness on the dynamic behavior of
adsorbents in the form of coated monoliths. Sep Purif Technol
2011;81:1919.
[202] Brandani F, Rouse A, Brandani S, Ruthven DM. Adsorption kinetics and
dynamic behavior of a carbon monolith. Adsorpt 2004;10:99109.
[203] Burchell TD, Judkins RR, Rogers MR, Williams AM. A novel process and
material for the separation of carbon dioxide and hydrogen sulde gas
mixtures. Carbon 1997;35:127994.
[204] -P Hao G, -C Li W, Qian D, Wang G-H, Zhang W-P, Zhang T, et al. Structurally
designed synthesis of mechanically stable poly(benzoxazine-co- resol)-based
porous carbon monoliths and their application as high-performance CO
2
capture sorbents. J Am Chem Soc 2011;133:1137888.
[205] Akhtar F, Andersson L, Keshavarzi N, Bergstrm L. Colloidal processing and
CO
2
capture performance of sacricially templated zeolite monoliths. Appl
Energy 2012;97:28996.
[206] Akhtar F, Bergstrm L. Colloidal processing and thermal treatment of
binderless hierarchically porous zeolite 13X monoliths for CO
2
capture. J
Am Ceram Soc 2011;94:928.
[207] Schumann K, Unger B, Brandt A, Schefer F. Investigation on the pore
structure of binderless zeolite 13X shapes. Microporous Mesoporous Mater
2012;154:11923.
[208] Pan CY, Basmadjian D. An analysis of adiabatic sorption of single solutes in
xed beds: pure thermal wave formation and its practical limitations. Chem
Eng Sci 1970;25:165364.
[209] Amundson NR, Aris R, Swanson R. On simple exchange waves in xed beds.
Proc R Soc Lond A 1965;286:12939.
[210] Li G, Xiao P, Xu D, Webley PA. Dual mode roll-up effect in multicomponent
non-isothermal adsorption processes with multilayered bed packing. Chem
Eng Sci 2011;66:182534.
[211] Menard D, Py X, Mazet N. Activated carbon monolith of high thermal
conductivity for adsorption processes improvement. Chem Eng Process
2005;44:102938.
[212] Ergun S. Fluid ow through packed columns. Chem Eng Prog 1952;48:8994.
[213] Bear J. Dynamics of uids in porous media. New York: American Elsevier;
1972.
[214] Dantas TLP, Luna FMT, Silva Jr IJ, Torres AEB, de Azevedo DCS, Rodrigues AE,
et al. Modeling of the xed-bed adsorption of carbon dioxide and a carbon
dioxide nitrogen mixture on zeolite 13X, Brazil. J Chem Eng 2011;28:53344.
[215] Dantas TLP, Luna FMT, Silva IJ, de Azevedo DCS, Grande CA, Rodrigues AE,
et al. Carbon dioxide-nitrogen separation through adsorption on activated
carbon in a xed bed. Chem Eng J 2011;169:119.
[216] Dantas TLP, Luna FMT, Silva IJ, Torres AEB, de Azevedo DCS, Rodrigues AE,
et al. Carbon dioxide-nitrogen separation through pressure swing adsorption.
Chem Eng J 2011;172:698704.
[217] Ruthven DM, Xu Z, Farooq S. Sorption kinetics in PSA systems. Gas Sep Purif
1993;7:7581.
[218] Ruthven DM, Xu Z. Diffusion of oxygen and nitrogen in 5A zeolite crystals and
commercial 5A pellets. Chem Eng Sci 1993;48:330712.
[219] Cybulski A, Moulijn JA. Monoliths in heterogeneous catalysis. Catal Rev Sci
Eng 1994;36:179270.
[220] Richardson JT, Peng Y, Remue D. Properties of ceramic foam catalyst
supports: pressure drop. Appl Catal A Gen 2000;204:1932.
[221] Avila P, Montes M, Mir EE. Monolithic reactors for environmental
applications: a review on preparation technologies. Chem Eng J
2005;109:1136.
[222] Greil P. Advanced engineering ceramics. Adv Mater 2002;14:70916.
[223] Mosca A, Hedlund J, Webley PA, Grahn M, Rezaei F. Structured zeolite NaX
coatings on ceramic cordierite monolith supports for PSA applications.
Microporous Mesoporous Mater 2010;130:3848.
[224] Wu Z, Hao N, Xiao G, Liu L, Webley P, Zhao D. One-pot generation of
mesoporous carbon supported nanocrystalline calcium oxides capable of
efcient CO
2
capture over a wide range of temperatures. Phys Chem Chem
Phys 2011;13:2495503.
N. Hedin et al. / Applied Energy 104 (2013) 418433 433

You might also like