Sei sulla pagina 1di 11

Journal of Industrial and Engineering Chemistry 23 (2015) 111

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


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

Review

A review on solid adsorbents for carbon dioxide capture


Seul-Yi Lee, Soo-Jin Park *
Department of Chemistry, Inha University, 100 Inharo, Incheon, Republic of Korea

A R T I C L E I N F O

Article history:
Received 3 June 2014
Received in revised form 31 August 2014
Accepted 4 September 2014
Available online 16 September 2014
Keywords:
CO2 capture
Dry adsorbent
Carbon materials
Zeolite
Silica
Metal-organic framework
Alkali metal
Metal oxide carbonate

A B S T R A C T

Global warming is considered as one of the great challenges of the twenty-rst century. CO2 capture and
storage (CCS) technology is attracting increasing interest to reduce the ever-increasing amount of CO2
released into the atmosphere and its impact on global climate change. CO2 capture process is a core
technology, and accounts for 7080% of the total cost of CCS technologies. CO2 capture technologies are
categorized as post-combustion, pre-combustion, and oxy-fuel combustion. Among these, postcombustion CO2 capture processes are regarded as being important green and economic technologies. It
is very important to develop new, highly efcient adsorbents to achieve techno-economic systems for
post-combustion CO2 capture. In this review, we therefore summarize dry solid adsorbents, which are
divided into non-carbonaceous (e.g., zeolites, silica, metal-organic frameworks and porous polymers,
alkali metal, and metal oxide carbonates) and carbonaceous materials (e.g., activated carbons, ordered
porous carbons, activated carbon bers, and graphene), with a focus on recent research.
2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Global warming and carbon economy . . . . . . . . . . . . . . . .
CO2 capture and storage (CCS) . . . . . . . . . . . . . . . . . . . . . .
CO2 capture technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Classication of CO2 capture technology . . . . . . . . . . . . . .
Post-combustion capture technology. . . . . . . . . . . . . . . . .
CO2 capture using dry adsorbents . . . . . . . . . . . . . . . . . . .
Non-carbonaceous dry adsorbents . . . . . . . . . . . . . . . . . . . . . . . .
Zeolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Silica materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Metal-organic frameworks (MOFs) and porous polymers.
Alkali-metal-based materials . . . . . . . . . . . . . . . . . . . . . . .
Metal oxide carbonate materials . . . . . . . . . . . . . . . . . . . .
Carbonaceous adsorbents . . . . . . . . . . . . . . . . . . . . . . . . . .
Activated carbons . . . . . . . . . . . . . . . . . . . . . . . .
Ordered porous carbons . . . . . . . . . . . . . . . . . . .
Activated carbon bers (ACFs) . . . . . . . . . . . . . .
Graphene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

* Corresponding author. Tel.: +82 32 876 7234; fax: +82 32 867 5604.
E-mail addresses: sjpark@inha.ac.kr, psjin@krict.re.kr (S.-J. Park).
http://dx.doi.org/10.1016/j.jiec.2014.09.001
1226-086X/ 2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

2
2
2
2
2
3
4
5
5
5
6
7
7
7
7
8
8
9
9
9
9

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111

Introduction
Global warming and carbon economy
Humans are endangered by global warming caused by the
greenhouse effect. The greenhouse effect can be attributed to an
increase in the emissions of greenhouse gases such as carbon
dioxide (CO2), methane (CH4), nitrous oxide (N2O), chlorouorocarbons, and sulfur hexauoride (SF6) since the beginning of
industrialization. The average world temperature has increased by
0.74% in the past 100 years and is expected to increase by another
6.4% by the end of the twenty-rst century [1,2].
Global warming leads to droughts, oods, heat waves, and
destruction of ecosystems, and the economic loss due to climate
change is expected to be 520% of the worlds gross domestic
product. CO2 is a major greenhouse gas and is the main cause of
global warming [3,4].
Major sources of CO2 emissions are thermoelectric power
plants and industrial plants (such as steel mills and reneries),
which account for approximately 45% of global CO2 emissions.
According to a report in 2013 by the International Energy Agency
(IEA), the global atmospheric CO2 concentration has increased
from a preindustrial value of 280 parts per million by volume
(ppmv) to 394 ppmv in 2012. Most of the CO2 emissions into the
atmosphere originate from the combustion of fossil fuels (99% of
global annual CO2 emissions of approximately 32 Gt) [5]. The IEA
and the Organization for Economic Co-operation and Development
(OECD) forecast that CO2 capture and storage (CCS) could take care
of approximately 14%, the expected reduction in volume of CO2
emission potentials [6]. Furthermore, the Intergovernmental Panel
on Climate Change (IPCC) reported that it would be possible to
achieve a more than 50% reduction in CO2 emissions from
2009 levels by 2050. According to their model for estimating
CO2 capture potential, it is estimated that 30 billion tons of CO2 can
be captured and stored within the European Union (EU) by
2050. Globally, 240 billion tons of CO2 could be captured by 2050
[7]. Without CCS technology, the cost of meeting a 50% global
reduction target by 2050 will be 70% higher [8].
Because of these global concerns, strict global regulations of
CO2 emissions to the atmosphere have been imposed. With
enforcement of the Kyoto Protocol, there has been increased
interest in the atmospheric residence time of CO2 and its
contribution to the greenhouse effect, and self-reduction techniques for CO2 generation rates and post-treatment of CO2 have
received signicant attention [9,10]. Various industries need to
deal actively with these regulations in order to survive. CCS has

great potential to be one of the more important green technologies


in the future.
There is an urgent need to develop CO2 reduction technologies,
and we believe that CCS is the main technology that can reduce CO2
emissions from the energy sector. However, with the exception of
developed countries, there is very little concern about, and
investment in, developing CCS. Fortunately, the world is now
paying more attention and aggressive efforts are being made to
commercialize CCS [1113]. A global CCS market has begun to
develop, along with certied emission reductions. The IEA/OECD,
EU, and United States forecast that the CCS market will grow to its
full capacity by 2020, and there is a need to obtain better insights
into the techno-economic possibilities of CO2 capture.
CO2 capture and storage (CCS)
The basic concept of CCS is to capture CO2 emissions without
releasing them into the atmosphere; they include sequestration or
storage, as shown in Fig. 1. CCS is the process of capturing and
compressing CO2 generated by existing large sources of highdensity CO2, transporting and depositing it safely in the ground or
an ocean-bedrock sediment layer, and long-term monitoring [14
16]. Transportation of CO2 refers to transporting captured and
compressed CO2 to a storage site via a pipeline or other means of
transport [17,18]. Storage of CO2 includes post-management: the
observation and prediction of the movement of the bedrock layer
after CO2 storage and evaluation of its effect on the environment,
and evaluation of the bedrock layers characteristics for depositing
captured CO2 (basins, oilelds, etc.) in the ground or below the sea
[19,20]. Technologies for capture, transportation, and storage of
CO2 are summarized in Table 1.
All the above processes are called CCS. In addition, CCS
technologies allow the continuing usage of fossil fuels and
stabilize the density of greenhouse gases. Moreover, CCS may
decrease the total cost of CO2 reduction and offer different ways to
reduce CO2 emissions [2123]. CCS is a comprehensive technology
for direct reduction of CO2 using the above-mentioned processes.

CO2 capture technology


Classication of CO2 capture technology
CO2 capture is a core technology and accounts for 7080% of the
total costs of CCS technologies. It is classied at (i) postcombustion, (ii) pre-combustion, and (iii) oxy-fuel combustion

Fig. 1. Concept and summary of CO2 capture and storage (CCS).

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111

Table 1
Technologies for capture, transportation, and storage of CO2.
Technologies

Classication

Applied areas

CO2 capture technologies

 Post-combustion
 Pre-combustion
Oxy-fuel combustion
CO2 compression and transportation

Direct capture from (thermoelectric) power plants and


industries (steel mills, reneries, etc.)

CO2 transportation technologies


CO2 storage technologies

 Characterization and evaluation of bedrock


layers
 Drilling and injection
 Observation and prediction of movement of
bedrock layers
Evaluation of the effect on the environment
and post management

technologies. Schematic diagrams of the three types CO2 capture


process are presented in Fig. 2.
Post-combustion capture technology involves collecting CO2
from the emission gases of a power plant [2426]. Pre-combustion
capture technology is mainly used in fertilizer and hydrogen
production and can be used in integrated gasication combined
cycle plants [2729]. Although the fuel conversion process (e.g.,
gasication of coal) prior to combustion is more complex and costs
more, the density and pressure of CO2 in the gases allow easy
separation. Oxy-fuel combustion, or oxygen-red combustion, is in
a trial phase in a pilot, plant and uses high-purity oxygen instead of
air [3032]. Thus, the density of CO2 in the gas ow, i.e., the
emission gas, becomes high, resulting in easy separation of CO2;
however, the separation of oxygen from the air consumes more
energy.

Transportation of captured CO2 to a storage site via a


pipeline or other means of transport
Deposit to a bedrock layer (basins, oilelds, etc.) in the
ground and below the sea and its management

Post-combustion capture technology


Among the currently available technologies, post-combustion
capture, a technology for capturing CO2 from post-combustion
emission gases, is the most easily applied technology for existing
sources of emissions. Post-combustion capture uses wet/dry
adsorbents, which are used for gas separation, and separates
and collects CO2 by adsorption/desorption. The classication of
technologies for post-combustion capture of CO2 is shown in Fig. 3.
In general, post-combustion capture technologies include wet
absorption [3335], dry adsorption [3638], membrane-based
technologies [3941], and cryogenics [42,43]. Current research
focuses on dry adsorption systems using dry adsorbents. Wet
absorption is good for treating large emission volumes from
combustion and is very useful for changing the density of CO2;

Fig. 2. Schematic diagrams of three types of CO2 capture process: (a) post-combustion, (b) pre-combustion, and (c) oxy-fuel combustion technologies.

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111

Fig. 3. Classication of application technologies for post-combustion capture of CO2.

however, it requires high energy for absorbent regeneration.


Additional drawbacks are (i) the use of heated absorbents, (ii)
erosion of materials, and (iii) slow solidgas reactions [44,45]. A
dry adsorption process has the following advantages: it requires a
simple device, is easy to operate, and has excellent environmental
effects and energy efciency; however, its performance is poor in
the treatment of large emission volumes and its separation
efciency is also poor [46,47]. A membrane-based process has the
following advantages: a simple device, easy operation, and low
energy consumption; however, it requires a high-cost module, is
not suitable for treating large volumes of emission gases, and is not
very durable [48,49]. Finally, cryogenics is a suitable process
because of its low investment cost and high reliability; however, it
is not suitable because of its energy consumption [50].
Recent research has focused on improving adsorbents and
developing efcient processes for cost-effective capture performance. The capture and separation of CO2 from a large volume of
combustion gases are still expensive and have high energy
consumption; therefore, more efcient adsorbents for relatively
concentrated streams need to be developed.
CO2 capture using dry adsorbents
The process of capturing CO2 using a dry adsorbent involves
selective separation of CO2 based on gassolid interactions [51]. In
general, universal dry adsorbents such as activated carbons and
molecular sieves are used in packed columns [52]. The important
variables in a dry adsorption process are the (i) surface tension and
pore size of the adsorbent and (ii) temperature and partial pressure
during the adsorption process [53]. The process involves a repeated
cycle of adsorption and desorption (regeneration).
The different types of adsorption are as follows: (i) pressureswing adsorption (PSA) [5456]; (ii) temperature-swing adsorption (TSA) [57,58], where two processes are combined, i.e.,
adsorption at a low temperature followed by desorption/regeneration by heating or lowering the pressure; (iii) electric-swing
adsorption (ESA) [59], i.e., adsorption/desorption by varying the
electricity supply, with a low-voltage current passing through the

adsorbent; and (iv) vacuum swing adsorption (VSA) [60]. Furthermore, adsorption-based technologies such as pressure/vacuumswing adsorption (PVSA) have frequently been investigated because
of their low energy requirements and relative simplicity [6163].
It is essential to improve the CO2 capture selectivity and
adsorptive capacity of a dry adsorbent for treating large volumes of
combustion emissions from many different sources. The CO2
adsorption can be improved and stabilized by introducing
functional groups (mainly amine groups) with high afnities for
CO2 onto the surface of the adsorbent material to react with CO2,
which can then be selectively adsorbed using the wide specic
surface area and pore structure of the adsorbent [6469].
To achieve high capture performance and high selectivity, the
incorporation of various amine groups into solid materials used as
CO2 capture sorbents is expected to improve polarization and CO2
capture. Such adsorbents have several advantages such as
potential elimination of corrosion problems and lower energy
costs for regeneration. The interactions between CO2 and amine
functional groups produce ammonium carbamates under anhydrous conditions as follows [70]:

1CO2 2RNH2 ! RNHCOO RNH3

2CO2 2R2 NH ! R2 NCOO R2 NH2


The adsorption process is signicant because it is reversible and
the adsorption efciency can be improved by modifying the
structure of the adsorbent materials. The CO2 adsorption efciency
can therefore be improved by selecting an appropriate adsorbent
material. At present, the main commercially available adsorbents
are activated carbons [71,72], zeolites [73], hollow bers, and
alumina. Each material has a different pore structure, specic
surface area, and surface functional groups, and their application
elds are highly specic. Some typical non-carbonaceous dry
adsorbents for CO2 capture are listed in Table 2 and will be
described in the following section.

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111

Table 2
Comparison between major non-carbonaceous dry adsorbents for CO2 capture.
Adsorbents

Advantages

Disadvantages

Zeolites, silica
materials

 Low production cost


 Large micropores/mesopores
 Medium CO2 adsorption
(at 298 K and 1 bar)

Metal organic
frameworks (MOFs)

 Large specic surface area (over 10,000 m2/g) and regular pore
distributions
 Ease of controlling pore sizes
 Possible improvement in CO2 selectivity according to various
combinations of metal clusters and organic ligands

Alkali-based dry
adsorbents
(K-, Na-, etc.)

 Possible adsorption and desorption at a low temperature, i.e.,


313-343 K (similar to amine-based absorption)
 Possible CO2 collection under wet conditions
Absorption and renewal under 473 K and possible operation at
atmospheric pressure (high economic efciency)

Metal oxides-based
adsorbents
(CaO, MgO, etc.)

 Dry chemical absorbents


 Adsorption/desorption at medium to high temperatures
(>673 K)
 Popular as a pre-combustion absorbent

 Poor performance of CO2 adsorption due to easy


moisture absorption
 Heavy energy consumption during CO2 desorption
(poor economic feasibility)
 Renewal difculties
 Poor performance at the partial pressure of CO2
 Poor economic efciency due to high production cost
 Complicated synthetic process
 Moisture-sensitive (possible structure failure due to
moisture absorption during CO2 capture)
 Unsuitable for use at high temperature. Mostly VSA
process (poor economic feasibility)
 Low adsorption capability (311 wt.%)
 High-temperature reactions
 Decrease in the collection ratio of CO2 because of stable
products (e.g., Na2CO3 and NaHCO3)
 Requires high temperatures during desorption
(high energy consumption)
 Complicated operation
 High consumption of energy due to adsorption/desorption
at medium to high temperatures (>673 K)
 High cost for regeneration
 Demand for continuous addition of absorbents
 Complicated process

Non-carbonaceous dry adsorbents


Zeolites
Zeolites are typical naturally occurring microporous crystalline
silicate framework materials; they can also synthesize in the
laboratory. They have uniform pore sizes of 0.51.2 nm, forming
networks of interconnecting channels or cages for adsorbing/
capturing gas molecules [74]. They have therefore been widely
used for gas separation and purication [7577]. Zeolites have
been extensively investigated for CO2 capture because of their
molecular sieving effect and the strong dipolequadrupole
(electrostatic) interactions between CO2 and alkali-metal cations
in the zeolite frameworks [78]. The cations in zeolites, such as Li,
Na, and Al, inuence the heat of adsorption of CO2; the heat of
adsorption increases with increasing monovalent charge density
(of negative charges) [79,80].
Many groups have reported the synthesis and characterization
of zeolites for CO2 capture, mainly zeolite 13X and zeolite 5A,
which gave CO2 capture performances of 325 wt.% at room
temperature and a CO2 pressure of 100% [8184]; they also gave a
CO2 capture performance of 212 wt.% at room temperature and a
CO2 partial pressure of 15% [8587].
Cavenati et al. [88] investigated zeolite 13X as a solid adsorbent.
Zeolite 13X exhibited a CO2 adsorption capacity of 28.7 wt.% and
CO2/N2 separation capacity of 3.65 at 298 K and 10 bar.
Jadhav et al. modied zeolite 13X via monoethanol amine
(MEA) impregnation to improve the CO2 adsorption capacity. The
CO2 adsorption capacity was better than that of the unmodied
zeolite by a factor of approximately 1.6 at 303 K, and at 393 K, the
efciency improved by a factor of 3.5. The higher capacity of MEAmodied zeolite 13X at 393 K indicated that the chemical
interactions between CO2 and amine groups might play a key
role in CO2 sorption, despite the reduced pore volume and lower
surface area resulting from MEA impregnation [89].
In the case of zeolite 5A, a typical CO2 adsorbent, the CO2
adsorption capacity and CO2/N2 selectivity were 20.8 wt.% and
8.45, respectively, at 298 K and 1 bar. At a higher pressure of
10 bar, the CO2 adsorption capacity of zeolite 5A was 22.3 wt.%
[90].

Recently, binderless zeolite NaX microspheres for CO2 capture


were prepared using chitosan-assisted synthesis. The obtained
microspheres had a uniform particle size of approximately 1.3 mm,
a specic surface area of 931 m2/g, and a moderate crushing
strength of 0.46 MPa. The CO2 adsorption capacities were 22.7 and
31.2 wt.% at 1 and 10 bar, respectively, at 298 K [91].
Most of the research using zeolites for CO2 capture from ue gas
has focused on PSA or VSA processes [9294]. The CO2 capture
performances of zeolites are greatly inuenced by the temperature
and pressure. Zeolites are typically used at pressures above 2 bar
and require very high regeneration temperatures, often above
573 K; therefore, CO2 regeneration results in huge energy loss
[95,96]. In addition, the adsorption performances of zeolites
decrease because they absorb moisture easily; therefore, many
studies have been conducted improve this [9799].
Silica materials
Many studies on the synthesis and modication of silica
materials containing a large amount of mesoporous materials for
efcient CO2 capture have been reported [100111].
An amine-rich nano-silica adsorbent was synthesized using
poly(acrylic acid) as a multi-functional bridge for amine
immobilization on the silica nanoparticle surfaces, followed
by treatment with polyethylenimine (PEI) as the amine source.
The CO2 adsorption capacity of the sample increased with
increasing temperature and reached a maximum CO2 adsorption
capacity of 16.7 wt.% at 313 K and 1 bar. The low adsorption
capacity at low temperatures was caused by kinetic limitations.
When the temperature was increased to 353 K, the adsorption
capacity decreased to only 4.4 wt.%, because the reaction
between CO2 and amines is an exothermic process (van t Hoff
behavior) [112].
Le et al. reported that tetraethylenepentamine (TEA)-functionalized/monodispersed porous silica microspheres were successfully prepared using tetraethoxysilane as the precursor and
dodecylamine as the templating agent and hydrolysis catalyst,
followed by calcining at 873 K, and functionalization with TEA.
They found that for the optimal TEA loading of 34 wt.% on the silica
microspheres, the maximum CO2 adsorption capacity was

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111

18.8 wt.% at 348 K and 1 bar. The sample showed relatively good
thermal stability in CO2 adsorption in the TSA process (desorption
temperature 393 K) [113].
Xu et al. reported that PEI-modied mesoporous MCM-41-type
molecular sieves (MCM-41-PEI) can serve as molecular baskets
for condensed CO2. MCM-41 has a synergetic effect on CO2
adsorption by PEI. The CO2 adsorption capacities of MCM-41-PEI50 (PEI loading 50 wt.%) and MCM-41-PEI-75 (PEI loading 75 wt.%)
were 11.2 and 13.3 wt.%, respectively, at 348 K and 1 bar [114].
Hicks et al. synthesized covalently tethered hyperbranched
aminosilica-based SBA-15 materials capable of adsorbing CO2
reversibly, with very high CO2 capacities of 13.6 wt.% at 298 K.
They suggested that the advantages of this adsorbent are its large
CO2 capacity and multicycle stability. The material was recycled by
thermally desorbing the CO2 from the surface with essentially no
changes in the capacity [65].
Sanz-Perez et al. tested tetraethylenepentamine (TEPA)functionalized SBA-15 (SBA-15-TEPA) using a simulated gas
mixture similar to that from a coal-red thermal power plant.
The CO2 adsorption capacity of SBA-15-TEPA (50) after 10 adsorptiondesorption cycles was 86% of the initial value, which was
11.35% under 100% CO2 at 318 K; it also reached a CO2 adsorption
capacity of 16.2% in a humid stream with 15% CO2 at 318 K and
1 bar. They concluded that the presence of 5% moisture enhanced
the CO2 adsorption capacity [115].
Kjdary et al. suggested the incorporation of Cu, Fe, Ag, and Au
nanoparticles into mercaptopropyl-modied silica for CO2 capture.
The CO2 adsorption capacities of the metal ions and nanoparticle
nanocomposites increased signicantly, by 20 and 100%, respectively. Cu-MP-S showed the maximum capacity, 22.9 wt.% at 323 K
and 1 bar [116].
Mello et al. synthesized amine-modied MCM-41 mesoporous
silica for CO2 capture. They showed that at a low pressures of
0.1 bar, the CO2 adsorption capacity of MCM-41-NH2 was
3.08 wt.%, which is much higher than that of 0.53 wt.% for
MCM-41. This is because of the great afnity of CO2, a Lewis acid,
for basic amine sites at low pressures. However, at high pressures
of 2.1 bar, the difference is not so pronounced; the CO2 adsorption
capacities of MCM-41-NH2 and MCM-41 were 5.06 and 4.4 wt.%,
respectively. From the results, they concluded that the most
reactive amine groups were bound to CO2 at low pressures, but the
available pore volume was lled at high pressures [117].
Besides the incorporation of amine-functional groups or metal
nanoparticles on silica supports, various silica materials with
controlled pore diameters and arrangements have been synthesized and they showed signicantly enhanced CO2 capture
performances [118122].
Metal-organic frameworks (MOFs) and porous polymers
Numerous MOFs have been synthesized from combinations
of inorganic (metal clusters) and organic units (ligands) by
strong bonding (reticular synthesis). The exibility with which
the geometries, sizes, and functionalities of the constituents
can be varied has led to more than 20,000 different MOFs being
reported and studied; their specic surface areas typically
range from 1000 to 10,000 m 2/g [123,124]. They are popular
because of their excellent selectivities and superior adsorption
capacities, particularity for H 2 adsorption [125127], CH 4
adsorption [128,129], and adsorption of some toxic gases
[130,131], and they are used as catalysts for ne chemical
synthesis [132,133].
The structures of MOF adsorbents for CO2 capture have
available spaces at the center because the molecules are
intertwined. Such a structure has a large body volume because
the pores are attached to organic molecules and knots of metallic

ions. MOFs are widely used as storage media for various gases
because the pore diameters can be easily controlled [134138].
McDonald et al. reported that N,N0 -dimethylethylenediamine/
H3[(Cu4Cl)3(BTTri)8(CuBTTti; H3BTTri = 1,3,5-tri(1H-1,2,3-triazol4-yl)benzene] composites adsorbed 15.4 wt.% of CO2 at 298 K and
1 bar and 9.5 wt.% of CO2 at 298 K and 0.15 bar CO2/0.75 bar N2,
with a selectivity of 327, determined using the ideal adsorbed
solution theory (IAST). They emphasized that amines tethered to
solid adsorbents have considerable advantages over aqueous
alkanolamines, because of their quick regeneration using mild
temperature swings [139].
Bao et al. reported that Mg-based MOF-74 had a very high CO2
adsorption capacity of 37 wt.% at 298 K and 1 bar [140] and Caskey
et al. reported that it had a considerable adsorption capacity of
23.6 wt.% at 296 K and 0.1 bar [141]; they investigated the
selectivities for CO2 and CH4, but not for CO2/N2. They suggested
that the exceptionally high performance of Mg-based MOF-74 for
CO2 may be attributed to strong interactions between the oxygen
lone pair orbitals of CO2 with coordinatively unsaturated metal
cations, suggesting that it is a promising adsorbent for CO2 capture
[142].
MOFs show high CO2 storage capacities at 298 K and 42 bar.
Gravimetric CO2 isotherms for various MOFs such as MOF-2,
MOF-505, Cu3(BTC)2, MOF-74, IRMOFs-11, -3, -6, and -1, and
MOF-177 were reported by Yaghis group. They conrmed that
the amine functionalities of the IRMOF-3 pores increased the
afnity for CO2. The highest CO2 adsorption capacity was
147 wt.%, for MOF-177 [143].
Yan et al. reported high enhancement of CO2 uptake by an
HKUST-1 MOF via a simple chemical treatment, to improve the CO2
capture performance. It had a maximum CO2 adsorption capacity
of 51 wt.% at 273 K and 1 bar, which is a signicant increase of 61%
compared with the original MOF sample [144]. Llewellyn et al.
reported chemical treatment of MIL-101 with ethanol and then
NH4F aqueous solution; the CO2 adsorption capacity of the
chemically treated MIL-101 increased from 79.2 to 176 wt.% at
304 K and 50 bar [145].
More recently, conjugated microporous polymers (CMPs) have
attracted much interest because of their porous structures, formed
from organic functionalities; they have applications such as gas
storage, capture, and separation [146148]. In particular, the
incorporation of metal-organic species in CMPs provides new
materials capable of CO2 capture and its simultaneous conversion
for cost-effective reduction of CO2. Xie et al. reported that the CO2
sorption properties of the Co-CMP and Al-CMP were comparatively
high, although these materials have a relatively low specic
surface areas compared with other MOF materials. The CO2
adsorption capacity and specic surface area of Co-CMP were
7.93 wt.% and 965 m2/g, respectively; those of Al-CMP were
7.65 wt.% and 798 m2/g, respectively. The co-catalysts Co/AlCMP displayed exceptionally high catalytic activity in the
conversion of propylene oxide and CO2 to propylene carbonate
at room temperature and atmospheric pressure [149].
MOFs have very high CO2 capture capacity at high pressure.
However, most MOFs have poor CO2 capture performances
compared to other solid physical sorbents at a low partial
pressure of CO2, and maintenance of a high partial pressure is
not economically feasible (process unit costs). An economically
viable MOF needs to be developed for the efcient storage and
separation of large volumes of CO2, up to 1 ton/day. Furthermore, the synthesis of MOF, a metal complex, from an organic
ligand is in many cases very expensive, and the synthetic
process is complicated. MOFs suffer from durability and
mechanical strength problems because of moisture absorption
during CO2 capture. The use of MOFs in power plants is therefore
limited.

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111

Alkali-metal-based materials
The use of alkali-metal (such as Na, K, Al) carbonates as
renewable dry adsorbents for capturing CO2 from ue gas at
operating temperatures below 473 K, relatively moderate conditions, with independent supports has been reported [150152]. In
this process, the alkali-metal carbonates are added to various
inorganic supports such as carbon materials, zirconia, ceramics,
silica, and alumina; CO2 absorption is achieved by reaction with
moisture (carbonation; (Eq. (3))), and the absorbent is renewed by
decarbonation (Eq. (4)) [153,154]:
M2 CO3 H2 O CO2 @ 2MHCO3 M Na; K

(3)

DH = 135 kJ/mol and 141 kJ/mol for M = Na and K, respectively.


2MHCO2 @ M2 CO3 H2 O CO2

(4)

It is well known that CO2 and H2O normally react with a


carbonate sorbent at 333383 K to form alkali-metal bicarbonates
in an adsorption process, as in Eq. (3), and the bicarbonate
regenerates the alkali-metal carbonate at 373473 K, with release
of CO2. The theoretical CO2 adsorption capacities of Na2CO3 and
K2CO3 are 41.5 and 31.8 wt.%, respectively.
Li-based zirconate (Li2ZrO3) and Li-based silicate (Li4SiO4) are
also promising CO2 captors and can be used for direct CO2
separation from ue gas at high temperatures, 700900 K
[155,156]. Li4SiO4, in particular, has great potential as a CO2
captor because of its high CO2 sorption capacity of 36.7 wt.% and its
small volume change during the CO2 adsorptiondesorption cycle
[157,158].
Kato et al. investigated the CO2 sorption capacities of Li2ZrO3
and Li4SiO4 at low CO2 concentrations, i.e., 50 ppv, and concluded
that the CO2 sorption capacity of Li4SiO4 is more than 30 times
greater than that on Li2ZrO3. Moreover, silica materials are cheaper
than zirconia materials [159,160]. More recently, Seggiani
et al. reported that Li4SiO4 with addition of 30 wt.% K2CO3 or
Na2CO3 showed a CO2 sorption capacity of 23 wt.%, at an optimum
sorption temperature of 853 K and low CO2 partial pressure of
0.04 bar, corresponding to a Li4SiO4 conversion of about 80%
oz et al. proposed Li8SiO6 as an alternative dry
[161]. Duran-Mun
adsorbent for CO2 capture, it showed a very high sorption capacity
of about 51.9 wt.% over a wide temperature range, with an
efciency of 71.1% [162].
These alkali-metal-based materials are technically and economically attractive for post-combustion CO2 capture at high
temperatures and low concentrations, because they do not need
further cooling processes; however, problems with the long-term
stabilities and sustained performances of these adsorbents under
real ue gas conditions in post-combustion applications need to be
solved.
Metal oxide carbonate materials
Mineral carbonation technology uses metal oxides such as CaO
and MgO as dry chemical adsorbents; a chemical reaction xes CO2
as an insoluble carbonate [163168]. MgO has been shown to be
good a particularly candidate for CO2 capture adsorbents because
of its low cost, abundance, and low toxicity. However, MgO has a
very low sorption capacity of 0.57 wt.% at moderate temperatures
in dry environments [169].
Han et al. suggested novel MgO-based mesoporous composites
with concrete-like structures, which were synthesized using a
facile coprecipitation method, for trapping CO2 in ue gas in the
high temperature range 423673 K. Microcrystalline MgO present
in the alumina frameworks had CO2 adsorption capacities of

7.7 and 13.1 wt.% in the absence or presence of water vapor,


respectively, at 473 K. Regeneration was performed at 873 K, and
stable cyclic adsorption was achieved [170].
Bhagiyalakshmi et al. synthesized mesoporous MgO using
mesoporous carbon, CMK-3, obtained from mesoporous SBA-15, as
an exotemplate. They conrmed that the pore structure of the
mesoporous MgO sample resembled those of SBA-15 and CMK-3.
The CO2 adsorption capacities of mesoporous MgO were 8 wt.% at
298 K and 10 wt.% at 373 K, whereas that of non-porous MgO was
12 wt.% at 298 K [171].
Liu et al. reported that MgO sorbents doped with alkali-metal
carbonates had enhanced CO2 sorption capacities at both low and
moderate operating temperatures, compared with pure MgO
sorbents. MgO-doped Cs2CO3 had a maximum CO2 sorption
capacity greater than 8.36 wt.% at 573 K [172].
However, the chemical reaction is too slow and requires high
consumption of energy. Moreover, the regeneration process is
performed at 6731073 K, i.e., it also has high energy consumption.
Therefore, although CO2 capture using metal-oxide carbonates has
a high absorption volume, this technology may be inappropriate
for CO2 storage.
Carbonaceous adsorbents
Although carbonaceous materials consist of a single element,
they have many advantages such as high thermal/chemical
stabilities, electrical and heat conductivities, strengths, elasticities,
and bio-afnities [173177]. They are particularly good for
applications for gas adsorption or storage, because they are
lightweight, and have very high specic surface areas and large
pore volumes [178180]. They also have advantages for CO2
capture: (i) carbon materials are not moisture sensitive; (ii) the
cost of carbon materials is reasonable; (iii) the adsorption/
desorption temperatures are below 373 K; (iv) they can be used
at atmospheric pressure; and (v) the energy consumption is low;
all of these factors inuence current studies in this eld.
Activated carbons
CO2 adsorption capacity is strongly sensitive to the textural
properties and surface groups of carbon-based adsorbents [181
183]. The pore size distributions of activated carbons vary from
micropore to macropore; therefore, activated carbons are inappropriate for selective adsorption of a specic gas. Generally,
pristine carbon-based adsorbents have weak afnities for CO2,
with adsorption heats of less than 25 kJ/mol [184]. The typical CO2
adsorption capacity of an activated carbon is 5 wt.% at 298 K and
0.1 bar [185,186].
The pore structures of activated carbons can be easily controlled
by varying the preparation and activation conditions [187
190]. Moreover, the functional groups on the activated carbon
surface can also be easily controlled using various treatments
[191194]. The incorporation of various basic groups on activated
carbon has been actively investigated for enhancing the CO2
afnity by increasing the CO2 adsorption capacity [195,196].
Zhao et al. designed and prepared a novel microporous carbon
material (KNC-A-K) with a high N-doping concentration, above
10.5 wt.%, and extra-framework K+ cations, which gave CO2
capture performances of 7.1 wt.% (0.1 bar) and 17.8 wt.% (1 bar)
at 298 K, and a CO2/N2 selectivity of 48, determined using IAST
method. They emphasized that its outstanding low-pressure CO2
adsorption ability makes KNC-A-K a promising candidate for
selective CO2 capture from ue gas. The K+ ions played a key role in
promoting CO2 adsorption via electrostatic interactions [197].
Lee et al. prepared heat-treated carbons by pyrolysis of
poly(vinylidene uoride) at various heat-treatment temperatures
and evaluated the CO2 adsorption capacities. The CO2 adsorption

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111

capacity increased with increasing heat-treatment temperature up


to 873 K (15.5 wt.% at 298 K and 1 bar) and then decreased at
973 K; this correlated with their micropore volumes. Interestingly,
they emphasized that the micropore volume had the greatest effect
on CO2 adsorption at 1 bar; meanwhile, the pore sizes also strongly
affected the CO2 adsorption behavior at pressures below 0.3 bar. In
addition, it was found that ultra-micropores (<0.65 nm) were
strongly preferred sites for CO2 [198].
Sevilla et al. prepared sustainable porous carbons synthesized
using polysaccharides and biomass by hydrothermal carbonization
and chemical activation. A very high CO2 adsorption capacity of
21.2 wt.% was achieved at 298 K and 1 bar. They observed that the
porous carbon samples obtained using mild activating conditions
had smaller specic surface areas; however, the mildly activated
carbons had the best CO2 adsorption capacities because of the
presence of abundant narrow micropores of size less than 1 nm
[199].
Wang et al. synthesized a novel imine-linked polymer (ILP) with
a high specic surface area of over 700 m2/g and N content of up to
10%. Then, by direct pyrolysis of the ILP at 8731073 K, a series of
N-doped porous carbons were formed, with specic surface areas
of up to 366 m2/g and narrow micropore size distributions. The CO2
capture performances of porous ILP and N-doped porous carbons
prepared by ILP pyrolysis at 1073 K were 4.62 and 8.58 wt.%,
respectively, at 298 K. It was conrmed that the CO2 adsorption
capacity was strongly affected by narrower micropore sizes in the
adsorbents [200].
Balsamo et al. synthesized an activated carbon (F50) using a
mixture of coal tar pitch and furfural (50/50 wt.%) and subsequent
steam activation. The F50 sample had a microporous structure and
the pore size distribution showed that most of the pores were
accessible to CO2 molecules. Moreover, the F50 activated carbon
was basic, which is suitable for CO2 capture. The CO2 adsorption
capacity was 2.68 wt.% at 303 K under typical ue gas conditions
(CO2 0.15 bar); this is approximately double and four times the
values obtained at 323 and 353 K, respectively [201].
Meng et al. reported the preparation of nanoporous carbon
containing N from polypyrrole using NaOH as the activating agent.
The N-doped activated carbon had a high specic surface area
(2169 m2/g) and showed a very high CO2 adsorption capacity of
17.7 wt.% at 298 K and 1 bar. They emphasized that this good CO2
adsorption capacity was attributable to the presence of narrow
micropores, together with a high density of N basic groups [202].
Meng et al. synthesized porous carbons with well-developed
pore structures, using a weak-acid cation-exchange resin (CER), by
the carbonization of mixtures containing different amounts of
magnesium acetate. They conrmed that the MgO template
method generated micro/mesoporous carbons with well-developed pore structures. As the magnesium acetate/CER ratio
increased, the specic surface area increased from 326 to
1276 m2/g; the CO2 adsorption capacities were 16.4 and
104.5 wt.% at 1 and 30 bar, respectively, at 298 K [203].
Jang et al. prepared NiO-loaded activated carbons (NiO-ACs)
using a post-oxidation method involving nickel electroless plating
at 573 K in an air stream. The CO2 adsorption capacity of the NiOAC samples increased with increasing oxidation time. The
maximum CO2 adsorption capacity was 49.9 cm3/g, which is
higher than that of pristine activated carbon, i.e., 41.2 cm3/g at
298 K and 1 bar. They conrmed that the NiO acted as electron
donors on the carbon surface, enhancing the adsorption of CO2,
which is an electron acceptor, because of the acidbase properties
[204].
Ordered porous carbons
Ordered porous carbon materials have attracted much research
interest because of their widespread applications in gas storage,

and as catalysts and supports, electrode materials, etc. [205


210]. Many synthetic methods for ordered porous carbons have
been reported including (i) direct synthesis by organicorganic
self-assembly, involving a combination of carbon precursors and
block copolymers as soft templates and (ii) nanocasting using silica
materials as structural directing hard templates [211214].
Ordered porous carbons are expected to act as CO2 adsorbents
because of their high specic surface areas, large pore volumes,
large adsorption capacities, high chemical stabilities, ease of
tuning of pore sizes or channels, and ease of surface modication
[215217]. In particular, their pore size distributions are quite
narrow and uniform compared with commercially available highsurface-area carbons. They are therefore considered to be very
suitable CO2 adsorbents. However, pristine-ordered porous
carbons give poor CO2 capture performances in terms of adsorption
capacity and selectivity; therefore, surface modication and pore
size control are essential for CO2 capture.
Bin et al. reported that an ordered mesoporous carbon
synthesized using a soft-templating method had a CO2 adsorption
capacity of 13.2 wt.% and selectivity for CO2/N2 of 12.8 at 278 K and
1 bar [218]. N-doped porous carbon fabricated by carbonization of
melamine-formaldehyde resins had a CO2 adsorption capacity of
9.9 wt.% at 298 K and 1 bar [219]. Ma et al. reported that N-doped
porous carbon synthesized using a facile one-pot evaporationinduced self-assembly method had a high specic surface area of
1979 m2/g and a high N content of 4.32 wt.%, and showed a CO2
adsorption capacity of 18.9 wt.% and CO2/N2 selectivity of 30 at
298 K and 1 bar [220].
Mahurin et al. investigated a material produced by grafting
amidoxime functional groups onto a hierarchical porous carbon
framework, synthesized using a soft-templating method, for the
selective capture and removal of CO2 from combustion streams.
The best adsorbent showed a CO2 adsorption capacity of 10.9 wt.%
at 298 K and 1 bar, and CO2/N2 selectivity of 22.4. They highlighted
the importance of balancing porosity and functionality to improve
the selective adsorption of CO2 [221].
Yoo et al. investigated the effect of the carbonization temperature of phenolic resins in the specic surface areas and total pore
volumes of ordered nanoporous carbons (ONCs). ONC carbonized
at 1173 K, with a specic surface area of 775 m2/g and a total pore
volume of 0.764 cm3/g, exhibited the highest CO2 adsorption
capacities (15.8 wt.% at 1 bar and 68.5 wt.% at 30 bar) at 298 K,
because of the higher specic surface area and narrower micropore
size distribution [222].
Sevilla et al. studied the chemical activation of two mesoporous
carbons obtained using hexagonal- (SBA-15) and cubic (KIT-6)ordered mesostructured silica as hard templates, to investigate the
effects of dual porosity of mesoporous carbons in CO2 capture. The
results showed that irrespective of the type of template carbon
used as the precursor or the operating conditions used for
synthesis, the activated mesoporous carbons exhibited similar
CO2 adsorption capacities of around 14.1 wt.% and CO2/N2
selectivities of 6.5, at 298 K and 1 bar. They suggested that the
CO2 adsorption capacity depends on the presence of narrow
micropores (<1.0 nm) rather than on the specic surface area
[186].
Activated carbon bers (ACFs)
ACFs are promising adsorbent materials, based on their
nanostructures, abundant micrometer porosities, and properties
such as high specic surface areas and narrow pore size
distributions [223227]. The brous shape of ACFs make them
easier to handle than granular and powdered adsorbents
[228,229].
Lee et al. modied commercially available ACFs by a chemical
activation method and obtained materials with superior CO2

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111

adsorption capacities. They demonstrated that the ACF pore


structure can be designed to enhance the CO2 adsorption capacity
by controlling the concentration of KOH intercalated into the
graphitic lattice. The highest CO2 adsorption capacity was
25.0 wt.% at 298 K and 1 bar, achieved using a materials heattreated with 3:1 KOH/ACF. It was also found that the CO2
adsorption capacity was strongly affected by the ultra-micropore
size distribution rather than by the specic surface area or
micropore volume [230]. This is because ultra-micropores have a
high adsorption potential, which enhances the adsorption of CO2
molecules [231,232].
Thiruvenkatachari et al. prepared large honeycomb-shaped
carbon ber composite (HMCFC) adsorbents for large-scale CO2
capture tests (the adsorbent mass in one column was 4.486 kg).
The average CO2 adsorption capacity was 11.9 wt.% for a simulated
ue gas consisting of 13% CO2, 5.5% O2, and the balance N2, at 293 K.
They also showed that the CO2 capture efciency of HMCFC
adsorbents is more effective in a combined vacuum and thermal
decomposition process [233].
Graphene
Graphene oxide (GO) is a derivative of graphene and can be
synthesized with various functional groups on the basal planes and
edges [234]. The surface modication of GO with various
functional groups and the synthesis of new types of GO-like
derivatives with lightweight frameworks has been widely
researched for applications such as gas storage and separation,
energy conversion, and sensors [235240].
Zhao et al. demonstrated that ethylenediamine (EDA)-intercalated GO had a CO2 adsorption capacity of 4.65 wt.% at 303 K and
1 bar for CO2/N2 mixed gases [241].
Lee et al., prepared isothermally exfoliated GO using a new CO2
pressure swing method. They suggested that the CO2 pressure
swing method exerted sufcient pressure to overcome the GOinterlayer van der Waals binding energy and expand the GO, thus
promoting GO exfoliation by enabling large pores to develop. The
results showed that the best sample had a specic surface area of
547 m2/g and total pore volume of 2.468 cm3/g, resulting in a CO2
adsorption capacity of 28.2 wt.% at 298 K and 30 bar [242].
Meng et al. reported thermally exfoliated graphene nanoplates
as novel high-efciency sorbents for CO2 capture. The prepared
graphene nanoplates had high capture capacities, 248 wt.%, at
298 K and 30 bar. The improved CO2 capture capacity of the
graphene nanoplates was attributed to the larger inter-layer
spacing and high interior void volume [243].
Aminated GO for CO2 adsorption was synthesized by the
intercalation reaction of GO with amines such as EDA, diethylenetriamine (DETA), and triethylene-tetramine (TETA). The results
showed that the adsorption performance of the sample with 50%
EDA covalently attached to GO was better than those of GO
modied with DETA or TETA. The EDA-modied sample had a CO2
adsorption capacity of 4.65 wt.% at 303 K in 15% CO2/N2 mixed
gases at a ow of 40 mL/min [241].
Conclusions
The performances of currently available adsorbents for CO2
capture technologies need to be improved in terms of working
adsorption capacity, cycle lifetime, and multicycle durability. The
development of new highly efcient adsorbents for CO2 capture is
necessary to obtain systems that are techno-economical. It is
therefore very important to acquire data on adsorption reactors,
regeneration processes, and overall process integration of capture
systems in power plants.
Furthermore, through the development of advanced adsorbents
for CO2 capture, highly techno-economical systems will be

achieved by combining advanced CO2 capture technologies with


related systems such as hydrogen generation in the steam
reforming reaction, water gas conversion, electricity generation
using hydrogen, fuel cells, and water treatments.
Acknowledgements
This work was supported by the Carbon Valley Project of the
Ministry of Knowledge Economy and the Energy Efciency &
Resources of the Korea Institute of Energy Technology Evaluation
and Planning (KETEP) Grant funded by the Ministry of Commerce,
Industry, Trade and Energy, Korea.
References
[1] J.T. Houghton, Y. Ding, D.J. Griggs, M. Noguer, P.J. Linden, X. Dai, K. Maskell, C.A.
Johnson, Climate Change: The Scientic Basis, Intergovernmental Panel on
Climate Change, 2001.
[2] Redrawing the Energy-Climate Map, Organisation for Economic Co-operation
and Development (OECD)/International Energy Agency (IEA), France, 2013.
[3] R.H. Williams, Toward Zero Emissions from Coal in China, China Clean Energy
Forum, China, 2001.
[4] M.Z. Jacobson, Energy Environ. Sci. 2 (2009) 148.
[5] M. Hoeven, CO2 Emissions from Fuel Combustion Highlights, International
Energy Agency, 2013.
[6] J. Garrett, S. McCoy, The 4th IEA International CCS Regulatory Network Meeting,
France, Organisation for Economic Co-operation and Development (OECD)/International Energy Agency (IEA), 2012.
[7] A. Stangeland, A Model for Estimating the CO2 Capture Potential, Bellona Paper,
Norway, 2006.
[8] The EU Technology Platform for Zero Emission Fossil Fuel Power Plants (ZEP), A
Vision for Zero-Emission Power Generation, 2006.
[9] Kyoto Protocol to the United Nations Framework Convention on Climate Change,
United Nations, 1998.
[10] T. Dixon, G. Leamon, P. Zakkour, L. Warren, Energy Proceed. 37 (2013) 7596.
[11] R.C. Pietzcker, T. Longden, W. Chen, S. Fu, E. Kriegler, P. Kyle, G. Luderer, Energy
64 (2014) 95.
[12] Economic Assessment of Carbon Capture and Storage Technologies, Global CCS
Institute, 2011.
[13] M.E. Boot-Handford, J.C. Abanades, E.J. Anthony, M.J. Blunt, S. Brandani, N.M.
Dowell, J.R. Fernandez, M.C. Ferrari, R. Gross, J.P. Hallett, R.S. Haszeldine, P.
Heptonstall, A. Lyngfelt, Z. Makuch, E. Mangano, R.T.J. Porter, M. Pourkashanian,
G.T. Rochelle, N. Shah, J.G. Yao, P.S. Fennell, Energy Environ. Sci. 7 (2014) 130.
[14] D.W. Keith, Science 325 (2009) 1654.
[15] B. Li, Y. Duan, D. Luebke, B. Morreale, Appl. Energy 102 (2013) 1439.
[16] T. Kuramochia, A. Ramrez, W. Turkenburg, A. Faaij, Renew. Sustain. Energy Rev.
19 (2013) 328.
[17] J.Y. Jung, C. Huh, S.G. Kang, Y. Seo, D. Chang, Appl. Energy 111 (2013) 1054.
[18] M.M.J. Knoope, A. Ramrez, A.P.C. Faaij, Int. J. Greenhouse Gas Cont. 16 (2013)
241.
[19] H. Hellevang, P. Aagaard, Appl. Geochem. 39 (2013) 108.
[20] J.D.O. Williams, M. Jin, M. Bentham, G.E. Pickup, S.D. Hannis, E. Mackay, Int. J.
Greenhouse Gas Cont. 18 (2013) 38.
[21] A. Kowalczyk, E. Harnisch, S. Schwede, M. Gerber, R. Span, Appl. Energy 112
(2013) 465.
[22] D. Hussain, D.A. Dzombak, P. Jaramillo, G.V. Lowry, Int. J. Greenhouse Gas Cont.
16 (2013) 129.
[23] P. Jaramillo, W.M. Grifn, S.T. McCoy, Environ. Sci. Technol. 43 (2009) 8027.
[24] M. Bui, I. Gunawan, V. Verheyen, P. Feron, E. Meuleman, S. Adeloju, Computers
Chem. Eng. 61 (2014) 245.
[25] A. Cotton, K. Patchigolla, J.E. Oakey, Fuel 117 (2014) 391.
[26] K. Goto, K. Yogo, T. Higashii, Appl. Energy 111 (2013) 710.
[27] J.M. Sanchez, M. Marono, D. Cillero, L. Montenegro, E. Ruiz, Fuel 114 (2013) 191.
[28] P. Babu, R. Kumar, P. Linga, Environ. Sci. Technol. 47 (2013) 13191.
[29] S.D. Kenarsari, M. Fan, G. Jiang, X. Shen, Y. Lin, X. Hu, Energy Fuels 27 (2013)
6938.
[30] S. Black, J. Szuhanszki, A. Pranzitelli, L. Ma, P.J. Stanger, D.B. Ingham, M. Pourkashanian, Fuel 113 (2013) 780.
[31] A. Skorek-Osikowska, L. Bartela, J. Kotowicz, M. Job, Energy Convers. Manage. 76
(2013) 109.
[32] T. Wall, R. Stanger, Y. Liu, Fuel 108 (2013) 85.
[33] Y.E. Kim, S.J. Moon, Y.I. Yoon, S.K. Jeong, K.T. Park, S.T. Bae, S.C. Nam, Sep. Purif.
Technol. 122 (2014) 112.
[34] (a) Y.S. Sistla, A. Khanna, J. Ind. Eng. Chem. 20 (2014) 2497;
(b) Y.E. Kim, J.H. Park, S.H. Yun, S.C. Nam, S.K. Jeong, Y.I. Yoon, J. Ind. Eng. Chem.
20 (2014) 1486.
[35] A.S. Lee, J.C. Eslick, D.C. Miller, J.R. Kitchin, Int. J. Greenhouse Gas Cont. 18 (2013)
68.
[36] C.W. Zhao, X.P. Chen, C.S. Zhao, Y.K. Liu, Energy Fuels 23 (2009) 1766.
[37] K.W. Kim, Y.H. Son, W.B. Lee, K.S. Lee, Int. J. Greenhouse Gas Cont. 17 (2013) 13.
[38] E.A. Roth, S. Agarwal, R.K. Gupta, Energy Fuels 27 (2013) 4129.

10

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111

[39] H. Ahn, D. Kim, V.M.A. Melgar, J. Kim, M.R. Othman, H.V.P. Nguyen, J. Han, S.P.
Yoon, J. Ind. Eng. Chem., in press. http://dx.doi.org/10.1016/j.jiec.2013.12.069.
[40] S.D. Kenarsari, D. Yang, G. Jiang, S. Zhang, J. Wang, A.G. Russell, Q. Wei, M. Fan,
RSC Adv. 3 (2013) 22739.
[41] X. Zhang, X. He, T. Gundersen, Energy Fuels 27 (2013) 4137.
[42] C.A. Scholes, M.T. Ho, D.E. Wiley, G.W. Stevens, S.E. Kentish, Int. J. Greenhouse
Gas Cont. 17 (2013) 341.
[43] C.F. Song, Y. Kitamura, S.H. Li, Appl. Energy 98 (2012) 491.
[44] A.B. Rao, E.S. Rubin, Environ. Sci. Technol. 36 (2002) 4467.
[45] S. Shen, X. Feng, S. Ren, Energy Fuels 27 (2013) 6010.
[46] M. Radosz, X.D. Hu, K. Krutkramelis, Y.Q. Shen, Ind. Eng. Chem. Res. 47 (2008)
3783.
[47] T. Witoon, Mater. Chem. Phys. 137 (2012) 235.
[48] A.A. Olajire, Energy 35 (2010) 2610.
[49] E. Favre, Chem. Eng. J. 171 (2011) 782.
[50] T.H. Oh, Renew. Sustain. Energy Rev. 14 (2010) 2697.
[51] S. Mamun, H.F. Svendsen, K.A. Hoff, O. Juliussen, Energy Convers. Manage. 48
(2007) 251.
[52] M. Balsamo, F. Rodrguez-Reinoso, F. Montagnaro, A. Lancia, A. Erto, Ind. Eng.
Chem. Res. 52 (2013) 12183.
[53] M. Balsamo, T. Budinova, A. Erto, A. Lancia, B. Petrova, N. Petrov, B. Tsyntsarski,
Separ. Purif. Technol. 116 (2013) 214.
[54] Z. Liu, W.H. Green, Ind. Eng. Chem. Res. 52 (2014) 9665.
[55] J. Schell, N. Casas, D. Marx, M. Mazzotti, Ind. Eng. Chem. Res. 52 (2013) 8311.
[56] G.D. Pirngruber, D. Leinekugel-le-Cocq, Ind. Eng. Chem. Res. 52 (2013) 5985.
[57] J.P. Sculley, W.M. Verdegaal, W. Lu, M. Wriedt, H.C. Zhou, Adv. Mater. 25 (2013)
3957.
[58] G.D. Pirngruber, F. Guillou, A. Gomez, M. Clausse, Int. J. Greenhouse Gas Cont. 14
(2013) 74.
[59] S. Datta, M.P. Henry, Y.J. Lin, A.T. Fracaro, C.S. Millard, S.W. Snyder, Ind. Eng.
Chem. Res. 52 (2013) 15177.
[60] Q. Huang, M. Eic, Separ. Puri. Technol. 103 (2013) 203.
[61] A. Goeppert, M. Czaun, R.B. May, G.K.S. Prakash, G.A. Olah, S.R. Narayanan, J. Am.
Chem. Soc. 133 (2011) 20164.
[62] Y.H. Kim, D.G. Lee, D.K. Moon, S.H. Byeon, H.W. Ahn, C.H. Lee, Kor. J. Chem. Eng.
31 (2014) 132.
[63] C.T. Chou, C.Y. Chen, Sep. Purif. Technol. 39 (2004) 51.
[64] T. Prenzel, M. Wilhelm, K. Rezwan, Chem. Eng. J. 235 (2014) 198.
[65] J.C. Hicks, J.H. Drese, D.J. Fauth, M.L. Gray, G. Qi, C.W. Jones, J. Am. Chem. Soc. 130
(2008) 2902.
[66] S.H. Lee, S.Y. Moon, H.S. Kim, J.S. Bae, E.K. Jeon, H.Y. Ahn, J.W. Park, RSC Adv. 4
(2014) 1543.
[67] M. Auta, N.D.A. Darbis, A.T.M. Din, B.H. Hameed, Chem. Eng. J. 233 (2013) 80.
[68] P. Nugent, Y. Belmabkhout, S.D. Burd, A.J. Cairns, R. Luebke, K. Forrest, T. Pham, S.
Ma, B. Space, L. Wojtas, M. Eddaoudi, M.J. Zaworotko, Nature 495 (2013) 80.
[69] H. Wei, S. Deng, B. Hu, Z. Chen, B. Wang, J. Huang, G. Yu, Chem. Sus. Chem. 5
(2012) 2354.
[70] F.Y. Chang, K.J. Chao, H.H. Cheng, C.S. Tan, Sep. Purif. Technol. 70 (2009) 87.
[71] M.G. Plaza, S. Garcia, F. Rubiera, J.J. Pis, C. Pevida, Chem. Eng. J. 163 (2010) 41.
[72] A.A. Adelodun, Y.H. Lim, Y.M. Jo, J. Ind. Eng. Chem. 20 (2014) 2130.
[73] M.T. Ho, G.W. Allinson, D.E. Wiley, Ind. Eng. Chem. Res. 47 (2008) 4883.
[74] A.W. Chester, E.G. Derouane (Eds.), Zeolite Characterization and Catalysis: A
Tutorial, Springer, New York, 2009.
[75] S. Chaupnik, W. Franus, M. Wysocka, G. Gzyl, Environ. Sci. Pollut. Res. 20 (2013)
7900.
[76] R. Yang, Z. Xu, S. Yang, I. Michos, L.F. Li, A.P. Angelopoulos, J. Dong, J. Membr. Sci.
450 (2014) 12.
[77] S. Choi, J.H. Drese, C.W. Jones, Chem. Sus. Chem. 2 (2009) 796.
[78] J. Zhang, R. Singh, P.A. Webley, Micropor. Mesopor. Mater. 111 (2008) 478.
[79] J.A. Dunne, M. Rao, S. Sircar, R.J. Gorte, A.L. Myers, Langmuir 12 (1996) 5896.
[80] R.M. Barrier, R.M. Gibbons, Trans. Faradays Soc. 61 (1965) 948.
[81] R.V. Siriwardane, M.S. Shen, E.P. Fisher, J.A. Poston, Energy Fuels 15 (2001) 279.
[82] P.J.E. Harlick, F.H. Tezel, Micropor. Mesopor. Mater. 76 (2004) 71.
[83] R. Hernandez-Huesca, L. Diaz, G. Aguilar-Armenta, Sep. Purif. Technol. 15 (1999)
163.
[84] T.D. Pham, R. Xiong, S.I. Sandler, R.F. Lobo, Micropor. Mesopor. Mater. 185 (2014)
157.
[85] P.J.E. Harlick, F.H. Tezel, Sep. Purif. Technol. 33 (2003) 199.
[86] G. Calleja, A. Jamenez, J. Pau, L. Dominguez, P. Pbrez, Gas. Separ. Purif. 8 (1994)
247.
[87] J.A. Dunne, R. Mariwhla, M. Rao, S. Sircar, R.J. Gorte, A.L. Myers, Langmuir 12
(1996) 5888.
[88] S. Cavenati, C.A. Grande, A.E. Rodrigues, J. Chem. Eng. Data 49 (2004) 1095.
[89] P.D. Jadhav, R.V. Chatti, R.B. Biniwale, N.K. Labhsetwar, S. Devotta, S.S. Rayalu,
Energy Fuels 21 (2007) 3555.
[90] D. Saha, Z.B. Bao, F. Jia, S.G. Deng, Environ. Sci. Technol. 44 (2010) 1820.
[91] L. Yu, J. Gong, C. Zeng, L. Zhang, Sep. Purif. Technol. 118 (2013) 188.
[92] D. Ko, R. Siriwardane, L.T. Biegler, Ind. Eng. Chem. Res. 42 (2003) 339.
[93] V.G. Gomes, K.W.K. Yee, Sep. Purif. Technol. 28 (2002) 161.
[94] B.J. Maring, P.A. Webley, Int. J. Greenhouse Gas Cont. 15 (2013) 16.
[95] P.J.E. Harlick, A. Sayari, Ind. Eng. Chem. Res. 45 (2006) 3248.
[96] K.T. Chue, J.N. Kim, Y.U. Yoo, S.H. Cho, R.T. Yang, Ind. Eng. Chem. Res. 34 (1995)
591.
[97] F. Brandani, D.M. Ruthven, Ind. Eng. Chem. Res. 43 (2004) 8339.
[98] G. Li, P. Xiao, P. Webley, J. Zhang, R. Singh, M. Marshall, Adsorption 14 (2008) 415.
[99] J. Merel, M. Clausse, F. Meunier, Ind. Eng. Chem. Res. 47 (2007) 209.

[100] M.U.T. Le, S.Y. Lee, S.J. Park, Int. J. Hydrogen Energy, in press. http://dx.doi.org/
10.1016/j.jiec.2013.11.051.
[101] M.A. Alkhabbaz, R. Khunsupat, C.W. Jones, Fuel 121 (2014) 79.
[102] P.J.E. Harlick, A. Sayari, Ind. Eng. Chem. Res. 46 (2007) 446.
[103] R. Sanz, G. Calleja, A. Arencibia, E.S. Sanz-Perez, Appl. Surf. Sci. 256 (2010) 5323.
[104] R.A. Khatri, S.S.C. Chuang, Y. Soong, M. Gray, Energy Fuels 20 (2006) 1514.
[105] D.S. Mebane, J.D. Kress, C.B. Storlie, D.J. Fauth, M.L. Gray, K. Li, J. Phys. Chem. C
117 (2013) 26617.
[106] (a) V. Zelenaka, D. Halamova, L. Gaberova, E. Bloch, P. Llewellyn, Micropor.
Mesopor. Mater. 116 (2008) 358;
(b) P. Li, B. Ge, S. Zhang, S. Chen, Q. Zhang, Y. Zhao, Langmuir 24 (2008) 6567.
[107] C. Knofel, C. Martin, V. Hornebecq, P.L. Llewellyn, J. Phys. Chem. C 113 (2009)
21726.
[108] Y. Liu, J. Shi, J. Chen, Q. Ye, H. Pan, Z. Shao, Y. Shi, Micropor. Mesopor. Mater. 134
(2010) 16.
[109] M.B. Yue, Y. Chun, Y. Cao, X. Dong, J.H. Zhu, Adv. Funct. Mater. 16 (2006) 1717.
[110] (a) M. Yang, Y. Song, L. Jiang, X. Wang, W. Liu, Y. Zhao, Y. Liu, S. Wang, J. Ind. Eng.
Chem. 20 (2014) 322;
(b) A. Garca-Abun, D. Gomez-Daz, J.M. Navaza, J. Ind. Eng. Chem. 20 (2014)
2272.
[111] C. Chen, S.T. Yang, W.S. Ahn, R. Ryoo, Chem. Commun. 24 (2009) 3627.
[112] Y. Du, Z. Du, W. Zou, H. Li, J. Mi, C. Zhang, J. Colloid. Interf. Sci. 409 (2013) 123.
[113] Y. Le, D. Guo, B. Cheng, J. Yu, J. Colloid. Interf. Sci. 408 (2013) 173.
[114] X. Xu, C. Song, J.M. Andresen, B.G. Miller, A.L. Scaroni, Energy Fuels 16 (2002)
1463.
[115] E.S. Sanz-Perez, M. Olivares-Marn, A. Arencibia, R. Sanz, G. Calleja, M.M. MarotoValer, Int. J. Greenhouse Gas Cont. 17 (2013) 366.
[116] N.H. Khdary, M.A. Ghanem, M.G. Merajuddine, F.M. Bin Manie, J. CO2 Utilization
5 (2014) 17.
[117] M.R. Mello, D. Phanon, G.Q. Silveira, P.L. Llewellyn, C.M. Ronconi, Micropor.
Mesopor. Mater. 143 (2011) 174.
[118] W.J. Son, J.S. Choi, W.S. Ahn, Micropor. Mesopor. Mater. 113 (2008) 31.
[119] C. Chen, S.T. Yang, W.S. Ahn, R. Ryoo, Chem. Commun. 242 (2009) 3627.
[120] S. Dasgupta, A. Nanoti, P. Gupta, D. Jena, A.N. Goswami, M.O. Garg, Sep. Sci.
Technol. 44 (2009) 3973.
[121] X.W. Liu, J.W. Li, L. Zhou, D.S. Huang, Y.P. Zhou, Chem. Phys. Lett. 415 (2005) 198.
[122] V. Zelenaka, M. Badanicova, D. Halamova, J. Cejka, A. Zukal, N. Murafa, G. Goerigk,
Chem. Eng. Sci. 144 (2008) 336.
[123] H. Furukawa, K.E. Cordova, M. OKeeffe, O.M. Yaghi, Science 341 (2013) 6149.
. Yazaydin, R.Q.
[124] H. Furukawa, N. Ko, Y.B. Go, N. Aratani, S.B. Choi, E.W. Choi, A.O
Snurr, M. OKeeffe, J. Kim, O.M. Yaghi, Science 329 (2010) 424.
[125] W. Qin, W. Cao, H. Liu, Z. Lia, Y. Li, RSC Adv. 4 (2014) 2414.
[126] (a) D. Sahu, P. Mishra, S. Edubilli, A. Verma, S. Gumma, J. Chem. Eng. Data 58
(2013) 3096;
(b) Y. Feng, H. Jiang, M. Chen, Y. Wang, Powder Technol. 249 (2013) 38.
[127] S. Dutta, J. Ind. Eng. Chem. 20 (2014) 1148.
[128] (a) Y. He, W. Zhou, T. Yildirim, B. Chen, Energy Environ. Sci. 6 (2013) 2735;
(b) L. Ding, A.O. Yazaydin, Micropor. Mesopor. Mater. 182 (2013) 185.
[129] M. Anbia, S. Sheykhi, J. Ind. Eng. Chem. 19 (2013) 1583.
[130] C. Petita, T.J. Bandosz, J. Mater. Chem. 19 (2009) 6521.
[131] T.G. Glover, G.W. Peterson, B.J. Schindler, D. Britt, O.M. Yaghi, Chem. Eng. Sci. 66
(2011) 163.
[132] Y.Y. Liu, K. Leus, Y. Bogaerts, K. Hemelsoet, E. Bruneel, V. van Speybroeck, P. van
Der Voort, Chem. Cat. Chem. 5 (2013) 3657.
[133] M. Jahan, Z. Liu, K.P. Loh, Adv. Funct. Mater. 23 (2013) 5363.
[134] J.R. Li, Y. Ma, M.C. McCarthy, J. Sculley, J. Yu, H.K. Jeong, P.B. Balbuena, H.C. Zhou,
Coord. Chem. Rev. 255 (2011) 1791.
[135] H. Yang, Z. Xu, M. Fan, R. Gupta, R.B. Slimane, A.E. Bland, I. Wright, J. Environ. Sci.
20 (2008) 14.
[136] K. Sumida, D.L. Rogow, J.A. Mason, T.M. McDonald, E.D. Bloch, Z.R. Herm, T.H.
Bae, J.R. Long, Chem. Rev. 112 (2012) 724.
[137] S. Keskin, T.M. van Heest, D.S. Sholl, Chem. Sus. Chem. 3 (2010) 879.
[138] J.M. Simmons, H. Wu, W. Zhou, T. Yildirim, Energy Environ. Sci. 4 (2011) 2177.
[139] T.M. McDonald, D.M. DAlessandro, R. Krishna, J.R. Long, Chem. Sci. 2 (2011)
2022.
[140] Z. Bao, L. Yu, Q. Ren, X. Lu, S. Deng, J. Colloid Interf. Sci. 353 (2011) 549.
[141] S.R. Caskey, A.G. Wong-Foy, A.J. Matzger, J. Am. Chem. Soc. 130 (2008) 10870.
[142] P.D.C. Dietzel, R.E. Johnson, H. Fjellvag, S. Bordiga, E. Groppo, S. Chavan, R. Blom,
Chem. Commun. 41 (2008) 5125.
[143] A.R. Millward, O.M. Yaghi, J. Am. Chem. Soc. 127 (2005) 17998.
[144] X. Yan, S. Komarneni, Z. Zhang, Z. Yan, Micropor. Mesopor. Mater. 183 (2014) 69.
[145] P.L. Llewellyn, S. Bourrelly, C. Serre, A. Vimont, M. Daturi, L. Hamon, G.D.
Weireld, J.S. Chang, D.Y. Hong, Y.K. Hwang, S.H. Jhung, G. Ferey, Langmuir 24
(2008) 7245.
[146] A. Li, H.X. Sun, D.Z. Tan, W.J. Fan, S.H. Wen, X.J. Qing, G.X. Li, S.Y. Li, W.Q. Deng,
Energy Environ. Sci. 4 (2011) 2062.
[147] L. Chen, Y. Honsho, S. Seki, D.L. Jiang, J. Am. Chem. Soc. 132 (2010) 6742.
[148] J.X. Jiang, C. Wang, A. Laybourn, T. Hasell, R. Clowes, Y.Z. Khimyak, J. Xiao, S.J.
Higgins, D.J. Adams, A.I. Cooper, Angew. Chem. Int. Ed. 50 (2011) 1072.
[149] Y. Xie, T.T. Wang, X.H. Liu, K. Zou, W.Q. Deng, Nature Commun. 4 (2013) 1960.
[150] C. Zhao, X. Chen, E.J. Anthony, X. Jiang, L. Duan, Y. Wu, W. Dong, C. Zhao, Prog.
Energy Combus. Sci. 39 (2013) 515.
[151] Y.H. Duan, B. Zhang, D.C. Sorescu, J.K. Johnson, J. Solid State Chem. 184 (2011)
304.
[152] Y.H. Duan, D.R. Luebke, H.W. Pennline, B.Y. Li, M.J. Janik, J.W. Halley, J. Phys.
Chem. C 116 (2012) 14461.

S.-Y. Lee, S.-J. Park / Journal of Industrial and Engineering Chemistry 23 (2015) 111
[153] A. Samanta, A. Zhao, G.K.H. Shimizu, P. Sarkar, R. Gupta, Ind. Eng. Chem. Res. 51
(2012) 1438.
[154] J.S. Hoffman, H.W. Pennline, Study of Regenerable Sorbents for CO2 Capture, in:
Proceedings of First National Conference on Carbon Sequestration, Washington,
DC, 2001.
[155] M. Kato, S. Yoshikawa, K. Nakagawa, J. Mater. Sci. Lett. 21 (2002) 485.
[156] K. Essaki, M. Kato, H. Uemoto, J. Mater. Sci. 21 (2005) 5017.
[157] M.J. Venegas, E. Fregoso, R. Escamilla, H. Preiffer, Ind. Eng. Chem. Res. 46 (2007)
2407.
[158] R. Pacciani, J. Torres, P. Solsona, C. Coe, R. Quinn, J. Hufton, T. Golden, L.F. Vega,
Environ. Sci. Technol. 45 (2011) 7083.
[159] M. Kato, K. Nakagawa, T. Ohashi, S. Yoshikawa, K. Essaki, Carbon Dioxide Gas
Absorbent Containing Lithium Silicate, US Patent No. US006387845B1 (2002).
[160] M. Kato, K. Nakagawa, K. Essaki, Y. Maezawa, S. Takeda, R. Kogo, Y. Hagiwara, Int.
J. Appl. Ceram. Tech. 2 (2005) 467.
[161] M. Seggiani, M. Puccini, S. Vitolo, Int. J. Greenhouse Gas Cont. 17 (2013) 25.
[162] F. Duran-Munoz, I.C. Romero-Ibarra, H. Pfeiffer, J. Mater. Chem. A 1 (2013) 3919.
[163] (a) P.C. Lin, C.W. Huang, C.T. Hsiao, H. Teng, Environ. Sci. Technol. 42 (2008)
2748;
(b) S.C. Lee, H.J. Chae, S.J. Lee, B.Y. Choi, C.K. Yi, J.B. Lee, C.K. Ryu, J.C. Kim, Environ.
Sci. Technol. 42 (2008) 2736.
[164] (a) A. Alamdari, A. Alamdari, D. Mowla, J. Ind. Eng. Chem., in press. http://
dx.doi.org/10.1016/j.jiec.2013.12.038. (b) M.I. Kim, S.J. Choi, D.W. Kim, D.W.
Park, J. Ind. Eng. Chem., in press. http://dx.doi.org/10.1016/j.jiec.2013.11.051.
[165] H. Gupta, L.S. Fan, Ind. Eng. Chem. Res. 41 (2002) 4035.
[166] M. Sayyah, B.R. Ito, M. Rostam-Abadi, Y. Lu, K.S. Suslick, RSC Adv. 3 (2013) 19872.
[167] F.C. Yu, N. Phalak, Z. Sun, Ind. Eng. Chem. Res. 51 (2012) 2133.
[168] H. Chen, C. Zhao, W. Yu, Appl. Energy 112 (2013) 67.
[169] T. Gaffney, T. Golden, S. Mayorga, J. Brzozowski, F. Taylor, Carbon Dioxide
Pressure Swing Adsorption Process using Modied Alumina Adsorbents, US
Patent No. US005917136A (1999).
[170] K.K. Han, Y. Zhou, Y. Chun, J.H. Zhu, J. Hazard. Mater. 203204 (2012) 341.
[171] M. Bhagiyalakshmi, J.Y. Lee, H.T. Jang, J. Int, Int. J. Greenhouse Gas Cont. 4 (2010)
51.
[172] M. Liu, C. Vogt, A.L. Chaffee, S.L. Chang, J. Phys. Chem. C 117 (2013) 17514.
[173] (a) S.J. Park, K.D. Kim, J. Colloid. Interf. Sci. 212 (1999) 186;
(b) M.K. Seo, S.J. Park, Macromol. Mater. Eng. 289 (2004) 368.
[174] A.M. Pinto, I.C. Goncalves, F.D. Magalhaes, Coll. Surf. B: Biointerf. 111 (2013) 188.
[175] S. Rondeau-Gagne, J.F. Morin, Chem. Soc. Rev. 43 (2014) 85.
[176] P. Bilalis, D. Katsigiannopoulos, A. Avgeropoulos, G. Sakellariou, RSC Adv. 4
(2014) 2911.
[177] L. Zhang, A. Aboagye, A. Kelkar, C. Lai, H. Fong, J. Mater. Sci. 49 (2014) 463.
[178] V.K. Gupta, T.A. Saleh, Environ. Sci. Pollut. Res. 20 (2013) 2828.
[179] S.Y. Lee, S.J. Park, Int. J. Hydrogen Energy 35 (2010) 6757.
[180] H. Seema, K.C. Kemp, N.H. Le, S.W. Park, V. Chandra, J.W. Lee, K.S. Kim, Carbon 66
(2014) 320.
[181] (a) M. Sevilla, P. Valle-Vigon, A.B. Fuertes, Adv. Funct. Mater. 21 (2011) 2781;
(b) M. Sevilla, N. Alam, R. Mokoya, J. Phys. Chem. C 114 (2010) 11314.
[182] B.C. Bai, S. Cho, H.R. Yu, K.B. Yi, K.D. Kim, Y.S. Lee, J. Ind. Eng. Chem. 19 (2013) 776.
[183] J. Yu, M. Guo, F. Muhammad, A. Wang, F. Zhang, Q. Li, G. Zhu, Carbon 69 (2014)
502.
[184] B. Guo, L. Chang, K. Xie, J. Nat. Gas Chem. 15 (2006) 223.
[185] S. Himeno, T. Komatsu, S. Fujita, J. Chem. Eng. Data 50 (2005) 369.
[186] M. Sevilla, A.B. Fuertes, J. Colloid. Interf. Sci. 366 (2012) 147.
[187] J. Wang, S. Kaskerl, J. Mater. Chem. 22 (2012) 23710.
[188] L.Y. Meng, S.J. Park, J. Colloid. Interf. Sci. 352 (2010) 498.
[189] H.M. Yoo, S.Y. Lee, B.J. Kim, S.J. Park, Carbon Lett. 12 (2011) 112.
[190] M. Choi, R. Ryoo, J. Mater. Chem. 17 (2007) 4204.
[191] S.J. Park, Y.S. Jang, J.W. Shim, S.K. Ryu, J. Colloid. Interf. Sci. 260 (2003) 259.
[192] S.J. Park, Y.S. Jang, J.W. Shim, S.K. Ryu, J. Colloid. Interf. Sci. 275 (2004) 342.
[193] K.S. Kim, S.J. Park, Electrochim. Acta 56 (2011) 10130.
[194] M.K. Seo, S.J. Park, Curr. Appl. Phys. 10 (2010) 241.
[195] B.J. Kim, K.S. Cho, S.J. Park, J. Colloid. Interf. Sci. 342 (2010) 575.
[196] L.Y. Meng, S.J. Park, Mater. Chem. Phys. 137 (2012) 91.
[197] Y. Zhao, X. Liu, K.X. Yao, L. Zhao, Y. Han, Chem. Mater. 24 (2012) 4725.

11

[198] S.Y. Lee, S.J. Park, J. Anal. Appl. Pyrol. 106 (2014) 147.
[199] M. Sevilla, A.B. Fuertes, Energy Environ. Sci. 4 (2011) 1765.
[200] J. Wang, I. Senkovska, M. Oschatz, M.R. Lohe, L. Borchardt, A. Heerwig, Q. Liu, S.
Kaskel, ACS Appl. Mater. Interf. 5 (2013) 3160.
[201] M. Balsamo, T. Budinova, A. Erto, A. Lancia, B. Petrova, N. Petrov, B. Tsyntsarski,
Sep. Purif. Technol. 116 (2013) 214.
[202] L.Y. Meng, S.J. Park, Mater. Chem. Phys. 143 (2014) 1158.
[203] L.Y. Meng, S.J. Park, J. Colloid. Interf. Sci. 366 (2012) 125.
[204] D.I. Jang, S.J. Park, Fuel 102 (2012) 439.
[205] T.Y. Ma, L. Liu, Z.Y. Yuan, Chem. Soc. Rev. 42 (2013) 3977.
[206] B. Sakintuna, Y. Yurum, Ind. Eng. Chem. Res. 44 (2005) 2893.
[207] S.Y. Lee, B.J. Kim, S.J. Park, J. Solid State Chem. 199 (2013) 258.
[208] (a) S.Y. Lee, S.J. Park, J. Colloid. Interf. Sci. 384 (2012) 116;
(b) S.Y. Lee, S.J. Park, J. Solid State Chem. 184 (2011) 2655.
[209] (a) S.E. Moradi, J. Ind. Eng. Chem. 20 (2014) 208;
(b) S.P. Dubey, A.D. Dwivedi, C. Lee, Y.N. Kwon, M. Sillanpaa, L.Q. Ma, J. Ind. Eng.
Chem. 20 (2014) 1126;
(c) C.K. How, M.A. Khan, S. Hosseini, T.G. Chuah, T.S.Y. Choong, J. Ind. Eng. Chem.,
in press. http://dx.doi.org/10.1016/j.jiec.2014.01.034.
[210] J.C. Ndamanisha, L. Guo, Anal. Chim. Acta 747 (2012) 19.
[211] J. Lee, J. Kim, T. Hyeon, Adv. Mater. 18 (2006) 2073.
[212] C.M. Ghimbeu, J.M.L. Meins, C. Zlotea, L. Vidal, G. Schrodj, M. Latroche, C. VixGuterl, Carbon 67 (2014) 260.
[213] M. Tian-Yi, L. Lei, Y. Zhong-Yong, Chem. Soc. Rev. 42 (2013) 3977.
[214] S. Karthikeyan, K. Viswanathan, R. Boopathy, P. Maharaja, G.J. Sekaran, Ind. Eng.
Chem., in press. http://dx.doi.org/10.1016/j.jiec.2014.04.036.
[215] J. Wei, D. Zhou, Z. Sun, Y. Deng, Y. Xia, D. Zhao, Adv. Funct. Mater. 23 (2013) 2322.
[216] Z. Tang, Z. Han, G. Yang, B. Zhao, S. Shen, J. Yang, New Carbon Mater. 28 (2013)
55.
[217] C.W. Moon, Y.J. Kim, S.S. Im, S.J. Park, Bull. Korean. Chem. Soc. 35 (2014) 57.
[218] B. Yuan, X. Wu, Y. Chen, J. Huang, H. Luo, S. Deng, Environ. Sci. Technol. 47 (2013)
5474.
[219] C. Pevida, T.C. Drage, C.E. Snape, Carbon 46 (2008) 1464.
[220] X. Ma, M. Cao, C. Hu, J. Mater. Chem. A 1 (2013) 913.
[221] S.M. Mahurin, J. Gorka, K.M. Nelson, R.T. Mayes, S. Dai, Carbon 67 (2014) 457.
[222] H.M. Yoo, S.Y. Lee, S.J. Park, J. Solid State Chem. 197 (2013) 361.
[223] S.J. Park, K.D. Kim, Carbon 39 (2001) 1741.
[224] B.J. Kim, S.J. Park, J. Colloid. Interf. Sci. 315 (2007) 791.
[225] J.S. Im, S.J. Park, Y.S. Lee, Mater. Res. Bull. 44 (2009) 1871.
[226] S.J. Park, M.K. Seo, Y.S. Lee, Carbon 41 (2003) 723.
[227] S.J. Park, B.J. Kim, J. Colloid. Interf. Sci. 291 (2005) 597.
[228] M. Suzuki, Carbon 32 (1994) 577.
[229] S.H. Yoon, S. Lim, Y. Song, Y. Ota, W. Qiao, A. Tanaka, I. Mochida, Carbon 42 (2004)
1723.
[230] S.Y. Lee, S.J. Park, J. Colloid. Interf. Sci. 389 (2013) 230.
[231] H. An, B. Feng, S. Su, Int. J. Greenhouse Gas Cont. 5 (2011) 16.
[232] J. Zhou, W. Li, Z. Zhang, W. Xing, S. Zhuo, RSC Adv. 2 (2012) 161.
[233] R. Thiruvenkatachari, S. Su, X.X. Yu, J.S. Bae, Int. J. Greenhouse Gas Cont. 13
(2013) 191.
[234] C.K. Chua, M. Pumera, Chem. Soc. Rev. 43 (2013) 291.
[235] I.V. Lightcap, P.V. Kamat, Acc. Chem. Res. 46 (2013) 2235.
[236] S.Y. Lee, S.J. Park, J. Nanosci. Nanotechnol. 13 (2013) 443.
[237] S.Y. Lee, S.J. Park, Int. J. Hydrogen Energy 36 (2011) 8381.
[238] K.S. Kim, S.J. Park, Electrochim. Acta 65 (2012) 50.
[239] S.Y. Lee, S.J. Park, Chapter 14: Adsorption Behaviors of Graphene and Graphenerelated Materials, Novel Carbon Adsorbents, Elsevier, New York, 2012.
[240] (a) W.W. Liu, S.P. Chai, A.R. Mohamed, U. Hashim, J. Ind. Eng. Chem. 20 (2014)
1171;
(b) G. Mittal, V. Dhand, K.Y. Rhee, S.J. Park, W.R. Lee, J. Ind. Eng. Chem., in press.
http://dx.doi.org/10.1016/j.jiec.2014.03.022.;
(c) Y. Liu, M.R. Park, H.K. Shin, B. Pant, J. Choi, Y.W. Park, J.Y. Lee, S.J. Park, H.Y.
Kim, J. Ind. Eng. Chem., in press. http://dx.doi.org/10.1016/j.jiec.2014.02.009.
[241] Y. Zhao, H. Ding, Q. Zhong, Appl. Surf. Sci. 258 (2012) 4301.
[242] S.Y. Lee, S.J. Park, Carbon 68 (2014) 112.
[243] L.Y. Meng, S.J. Park, J. Colloid. Interf. Sci. 386 (2012) 285.

Potrebbero piacerti anche