You are on page 1of 6

Journal of Colloid and Interface Science 386 (2012) 285–290

Contents lists available at SciVerse ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Effect of exfoliation temperature on carbon dioxide capture of graphene nanoplates


Long-Yue Meng, Soo-Jin Park ⇑
Korea CCS R&D Center, Korea Institute of Energy Research, 152 Gajeongro, Yuseoung-gu, Daejeon 305-343, South Korea
Department of Chemistry, Inha University, 100 Inharo, Nam-gu, Incheon 402-751, South Korea

a r t i c l e i n f o a b s t r a c t

Article history: Thermally exfoliated graphene nanoplates were found to be a novel high efficiency sorbent for the cap-
Received 3 April 2012 ture of CO2. The exfoliated graphene nanoplates were expanded successfully from graphite oxide by a
Accepted 3 July 2012 low-heat treatment at temperatures ranging from 150 to 400 °C under vacuum conditions. The texture
Available online 16 July 2012
was analyzed by N2 full isotherms and XRD. The CO2 capture characteristics of the graphene nanoplates
at 25 °C and 30 bar were examined using a pressure–composition–temperature apparatus. The inter-
Keywords: layer spacing of the graphene layers and pore structure on the CO2 capture capacities were studied as
CO2 capture
a function of the processing conditions. The prepared graphene nanoplates exhibited high capture capac-
Thermally exfoliation
Graphene nanoplates
ities, up to 248 wt.%, at 25 °C and 30 bar. The improved CO2 capture capacity of the graphene nanoplates
was attributed to the larger inter-layer spacing and higher interior void volume.
Crown Copyright Ó 2012 Published by Elsevier Inc. All rights reserved.

1. Introduction relatively moderate strength, easier desorption, and thermal stabil-


ity [12]. Up to now, previous studies have focused on the use of
Recently, increasing CO2 emission due to industrialization is a activated carbon, activated carbon fibers, carbon molecular sieves,
global environmental problem that has attracted increasing atten- carbon nanotubes, and graphene as adsorbents for CO2 capture
tion [1]. Carbon capture and storage (CCS) (carbon capture and [11,16–19].
sequestration) projects refer to technology attempting to prevent Graphene is a very important carbon material that has attracted
the release of large quantities of CO2 into the atmosphere from fos- considerable attention owing to its high aspect ratio, small radius
sil fuel use in power generation and other industries [2]. As a re- of curvature, high mechanical strength, unique electrical proper-
sult, current or proposed research into CO2 capture from flue gas ties, and chemical stability [20,21]. In recent years, there has been
focused on absorption, adsorption (porous materials), cryogenic increasing interest in graphene-based materials for use as sorbents
distillation, and membrane separation [3–6]. in energy and environment-related applications because the highly
Among those methods, the technology currently considered as ordered pores confined geometry in carbon nanostructures, poros-
mature is CO2 absorption by amine based liquids [3]. This technol- ity, and accessible surface area [18,22–26]. On the other hand, few
ogy, however, is expensive and energy intensive. Therefore, im- studies have examined CO2 adsorption on graphene because of its
proved technologies for CO2 adsorption are necessary to achieve higher specific surface area than other carbon materials.
low energy consumption [7]. CO2 adsorption is typically performed Modifications of graphene oxide layers by targeting hydroxy
at a final polishing step in a hybrid CO2 capture system [8,9]. Pre- and epoxy surface functional groups with both organic and inor-
viously, pressure swing adsorption (PSA) is one of the potential ganic compounds can increase the interlayer spacing of graphene
techniques that can be applicable to the removal of CO2 from high oxide layers [27]. For example, Srinivas et al. reported the synthesis
pressure gas streams. of a range of porous graphene oxide frameworks (specific surface
Up to now, different types of adsorbents have been used in CO2 area = 470 m2/g) through the expansion of graphene oxide sheets
PSA, such as zeolites, activated carbons, silica gel, metal–organic with a range of linear boronic acid pillaring units in a solvothermal
frameworks (MOFs), and porous organic polymers [10–15]. Acti- reaction [18]. Although the chemical modification and chemical
vated carbons generally have higher CO2 adsorption capacity than reduction approach can remove most of the oxygen atoms from
zeolite-like materials at pressures greater than atmospheric pres- the surface of graphene oxide, some remaining oxygen atoms
sure due to surface hydrophobicity, extended surface area, and additional functional groups introduced during chemical mod-
ification or reduction might cause the scattering of electrons,
which decreases the structural instability and thermal stability.
⇑ Corresponding author. Address: Department of Chemistry, Inha University, 100
In addition, these methods are very expensive and energy intensive
Inharo, Nam-gu, Incheon 402-751, South Korea. Fax: +82 32 860 8438.
E-mail address: sjpark@inha.ac.kr (S.-J. Park). for large-scale preparation.

0021-9797/$ - see front matter Crown Copyright Ó 2012 Published by Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcis.2012.07.025
286 L.-Y. Meng, S.-J. Park / Journal of Colloid and Interface Science 386 (2012) 285–290

Therefore, it is essential to develop a more effective method to 2.2. Characterization


prepare graphene sheets with high gas adsorption capacity. To
date, several methods have been used to obtain a high specific sur- Wide-angle X-ray diffraction (XRD, Rigaku Model D/MAX-III B
face area of graphene, including heat treatment at high tempera- with a rotation anode) patterns of the samples were obtained using
tures and vacuum conditions [26–31]. Ruoff et al. reported the Cu Ka radiation (k = 0.15418 nm). The XRD patterns were recorded
KOH activation of microwave exfoliated graphene oxide and ther- over the range, 10–60° 2h. The specific surface area and total pore
mally exfoliated graphene oxide to achieve a specific surface area volume were determined by N2 adsorption at 77 K using a surface
up to 3100 m2/g [30,31]. area analyzer (BEL Japan). Before the measurement, the samples
In this study, thermally exfoliated graphene nanoplates were were degassed at 100 °C for 12 h to obtain a residual pressure of
prepared using a novel high efficiency sorbent for the capture of <106 mm Hg. The N2 adsorbed on the graphene nanoplates was
CO2. The exfoliated graphene nanoplates were expanded success- used to calculate the specific surface area using the BET equation.
fully from graphite oxide by a low-heat treatment over the temper- The total pore volume was estimated to be the liquid volume of
ature range, 150–400 °C, under vacuum conditions. the N2 at a relative pressure of approximately 0.995, and the mi-
cro- and mesopore structures were analyzed using Horvath–Kawa-
zoe (HK) and Barrett–Joyner–Halenda (BJH) equations,
2. Experimental respectively. TGA data were obtained with a NETZSCH TG 209 F3
analyzer at a heating rate of 10 °C/min from 30 to 850 °C under a
2.1. Materials and methods N2 atmosphere. XPS measurements were performed on an ESCA-
LAB220i-XL (VG Scientific, UK) spectrometer with monochroma-
A 1.0 g sample of pristine graphite flakes and 50 ml of acetic tized Al Ka X-ray radiation as the X-ray source for excitation in
acid was mixed by ultrasonication for 2 h and then left to stand order to confirm the nitrogen contents.
for 1 week. The resulting graphite nanoplatelets were washed with A CO2 adsorption test was conducted under ambient conditions
acetone and dried in an oven at 80 °C for 12 h. The obtained graph- of 298 K at both low (1 bar) and moderate pressures (30 bar, BEL,
ite nanoplatelets were exfoliated to graphite oxide using Hum- Japan). In each experiment, approximately 0.1 g of the sample
mer’s method [32–34]. 1.0 g of graphite nanoplatelets and 0.5 g was loaded into a stainless chamber. Before the measurements,
of NaNO3 were added to 23 ml of concentrated H2SO4 at 0 °C. With the samples were degassed at 100 °C for 24 h to obtain a residual
vigorous stirring, 3 g of KMnO4 was added gradually, and the tem- pressure of <106 mm Hg. After the chamber was cooled to room
perature of the mixture was controlled to be less than 20 °C. The temperature, CO2 was introduced to pressure of 30 bar. Ultrahigh
ice bath was then removed, and the temperature of the mixture purity grade (99.9999%) CO2 was used to exclude the effects of
was maintained at 35 °C for 0.5 h. Subsequently, 46 ml H2O was moisture and other impurities. Finally, a volumetric measurement
added slowly to the mixture, which made the mixture boil. After method was used to determine the CO2 adsorption capacity.
15 min, 140 ml of hot water and an aqueous H2O2 solution was
added to the deeply brown mixture with constant stirring. The 3. Results and discussion
resulting suspension was filtered while still hot. The solid mixture
was washed sequentially with a 5% HCl aqueous solution and ace- The structural properties of the pristine graphite flakes, graphite
tone. The resulting solid was dried at 60 °C under vacuum. Finally, oxide, and graphene nanoplates were characterized by XRD, as
the prepared graphite oxide nanoplates were placed in an alumina shown in Fig. 2. For the pristine graphite flake, a sharp and intensive
boat located to obtain a residual pressure of less than 106 mm Hg peak at 26.6° corresponding to the graphitic structure (0 0 2) indi-
and then heated to a certain temperature (150, 200, 250, 300, and cated a highly organized layered structure with an interlayer spac-
400 °C, respectively) with the 1 °C/s of heating rate. The samples ing of 0.335 nm (Fig. 2). After oxidation, the graphitic structure
were designated GN-150, GN-200, GN-300, and GN-400, respec- (0 0 2) of the graphite flake powder disappeared and an additional
tively. Fig. 1 shows this process. It was observed that, upon vac- peak at 11.7° assigned to the (0 0 1) diffraction peak of graphite
uum-assisted thermal treatment, the grayish GO powder became oxide was observed [32–35]. According to previous research, the in-
graphene nanoplates with a huge volume expansion. ter-layers of graphite can be intercalated by a variety of molecules

Fig. 1. Schematic diagram of the preparation of graphene nanoplates.


L.-Y. Meng, S.-J. Park / Journal of Colloid and Interface Science 386 (2012) 285–290 287

1200 Graphite
GO

N2 adsorbed (cm3/g, 77 K )
d=0.335 nm

Intensity (a.u.)
GN-150
GN-200
GN-300
GN-400
800

400
Graphite
d=0.806 nm GO
GN-400 0

10 20 30 40 0.0 0.2 0.4 0.6 0.8 1.0


2 theta (°) P/P0

Fig. 2. XRD patterns of the pristine graphite flake, graphite oxide, and graphene Fig. 4. N2 full isotherms of the graphene nanoplates as the function of temperature.
nanoplates (GN-400).

2
or ions, such as oxygen functional groups, NO 3 and SO4 [34]. The studies reported that the graphite and graphite oxide exhibit neg-
XRD peak of graphite oxide corresponds to a layered structure with ligible specific surface area and porosity for N2 adsorption [36,37].
a basal spacing of 0.806 nm. Therefore, the oxidation process leads In contrast, the prepared graphene nanoplates exhibited reason-
to an increase in d-spacing, which can be attributed to the oxide-in- ably higher accessible N2 adsorption capacity. The curves of the
duced oxygen containing functional groups. graphene nanoplates showed type IV isotherms according to the
In contrast, no apparent peaks were detected in the patterns for IUPAC classification, as well as a hysteresis loop at relative pres-
GN-400, which might have two explanations: (a) very thin graph- sures >0.4, indicating the existence of mesopores. This shows that
ene layers due to the high degree of exfoliation; and (b) disordered the treated graphene nanoplates have a broad pore size distribu-
structure of graphene caused by the attachment of oxygenated tion. A comparison of the curves of all the samples showed that
functional groups. After graphite oxide was exfoliated thermally the heat temperature promotes the development of porosity in
at 400 °C, the peak at approximately 11.7° from the GO shifted to the samples through the exfoliation of graphene nanoplates at dif-
a lower angle. The interlayer spacing of GN-400 was estimated to ferent temperatures.
be 0.360 nm at 23.5°, which indicates a distorted graphite struc- More recently, Lv et al. [28] and Zhang et al. [29] prepared few-
ture and suggests the formation of graphene nanoplates. This con- layered graphene sheets by thermally exfoliating GO under vac-
firms that vacuum-assisted thermal treatment occurred at these uum at a significantly low temperature, which is thought to exert
low temperatures, leading to disruption of the periodic layered an outward force on the expanding graphene layers and help accel-
structure of graphite oxide plates [28,29]. erate the expansion of graphene layers [28,29]. However, the sur-
When the exfoliation temperature was changed, the XRD pat- face area (368 m2/g) of Lv and coworker’s graphene sheets was
terns of the graphene nanoplates obtained showed a typical broad much lower than that of Zhang and coworker’s graphene sheets
peak with the obvious disappearance of the characteristic peaks, (758 m2/g). They reported that the exfoliation temperature can
which might be due to very thin graphene layers because of the be further reduced and the surface area greatly improved under
high degree of exfoliation (Fig. 3). This suggests that the graphene vacuum condition. Herein, we prepared graphene nanoplates un-
nanoplates are exfoliated into a few-layers, resulting in a new lat- der 106 mm Hg at a lower temperature for synthesis CO2
tice structure that is significantly different from the pristine graph- adsorbents.
ite flakes and graphite oxide. The interlayer spacing of the Table 1 lists the textural properties of graphite, graphite oxide,
graphene nanoplates (0.358–0.364 nm) was larger than that of nat- and graphene nanoplates. After heat treatment, the specific surface
ural graphite (0.335 nm). As the exfoliation temperature was in- area and total pore volume of graphene nanoplates increased up to
creased, the basal spacing (0 0 2) increased gradually to GN-300 300 °C, but decreased at 400 °C. The increase in total pore volume
and then decreased at higher temperatures. This is probably due was accompanied by an increase in mesopore volume. The Lang-
to the increasing behavior of the graphite peak reflecting the open- muir surface area and average pore diameter generally decreased
ing of graphite layers after heat treatment. with increasing heat treatment. This suggests that heat treatments
Fig. 4 shows typical adsorption/desorption isotherms of N2 at can add to both the specific surface area and pore volume by
77 K onto the pristine graphite flakes, graphite oxide, and graphene exposing the inner graphite layers. The specific surface of the
nanoplates as the function of the heat temperature. Previous graphene nanoplates calculated from the BET equation was
480 m2/g for GN-300. This result is lower than that of Zhang’s ther-
mally exfoliated grapheme sheets, but that is higher than that re-
ported by Mishra’s (443 m2/g), Srinivas’ (470 m2/g), and Lv’s
0.358 nm
(368 m2/g) [18,25,28]. Recent research works have shown that por-
ous materials with a large population of pores of size 1 nm exhi-
Intensity (a.u.)

bit a better CO2 adsorption performance than those with narrower


micropores or with a mesoporous pore network [14]. And, the pre-
GN-200
0.364 nm vious work demonstrated that the porous carbons with ultrami-
cropores (<1 nm) are able to capture a large amount of CO2 at
GN-300
0.360 nm 1bar [6,17].
GN-400 The pore size distributions were examined for better character-
ization of the changes in the porous structure of the graphene
nanoplates. Fig. 5 shows the micro-(Fig. 5a) and meso-(Fig. 5b)
20 22 24 26 28 30
pore size distribution of the graphene nanoplates prepared at tem-
2 theta (°)
peratures ranging from 200 to 400 °C, which were determined
Fig. 3. XRD patterns of the graphene nanoplates as the function of temperature. using the HK and BJH method, respectively. According to the
288 L.-Y. Meng, S.-J. Park / Journal of Colloid and Interface Science 386 (2012) 285–290

Table 1
Pore structure parameters and surface group distribution of C1s XPS regions for the pristine graphite flake, graphite oxide, and graphene nanoplates from N2 adsorption
isotherms.

Samples SBETa (m2/g) SLangmirb (m2/g) VTc (cm3/g) Dd (nm) C1s (surface group distributions (%))
CAC; CAH CAOAC C@O COOH; OAC@O
Graphite 3 3 0.005 7.3 95.8 2.2 1.1 0.9
GO 14 20 0.054 15.4 78.7 12.8 3.7 4.8
GN-150 151 134 0.632 16.7 77.1 17.2 2.8 2.9
GN-200 437 313 1.716 15.7 79.9 15.9 2.2 2.0
GN-300 480 247 1.730 14.4 81.4 15.5 1.7 1.4
GN-400 324 245 1.032 12.7 83.8 14.1 1.3 0.8
a
Specific surface area: BET equation.
b
Specific surface area: Langmuir equation.
c
Total pore volume: Vads (P/P0 = 0.995)  0.001547.
d
Average pore diameter: 2  SBET/Vads.

(a) 0.4 Graphite


100
GO
GN-200 80
0.3 GN-300

TG (wt%)
GN-400
60
dV/dd0

0.2
40
GO
0.1
20
0.0
0
1 2 3 4 0 200 400 600 800 1000
d0 (nm) Temperature (oC)

Fig. 6. TG curves of a graphite oxide.


(b)
2.5
Graphite
GO
2.0 GN-200 macro-pores around 900 Å by partial gasification and expansion
GN-300 of the interlayer spacing between the graphene planes.
GN-400
1.5 The heating-driven structural changes of graphite oxide were
dV/dd0

estimated by TGA, and the results are shown in Fig. 6. The decom-
1.0 posing temperature range of graphite oxide is from 100 to 500 °C,
which indicates that the oxygen-containing functional groups
0.5 bonded to graphene surfaces are removed in this temperature
range [28,29]. However, such a broad temperature range may drive
0.0 oxygen to flee from the planar graphene nanoplates and is not en-
20 40 60 80 100
ough to fully expand the graphene layers under an atmospheric
d0 (nm) pressure [29]. Therefore, the vacuum-assisted thermal treatment
is needed for fast and fully expansion of graphene nanoplates in
Fig. 5. Micropore (a) and mesopore (b) size distribution of the pristine graphite
flake, graphite oxide, and graphene nanoplates. this work. Usually, the deconvolution of the C1s spectra yields four
peaks with different binding energy values representing carbon in
the non-oxygenated CAC or CAH (284.5 eV), single CAO bonds
literature, the collision and kinetic diameter (in Å) of a CO2 (285.5 eV), double C@O bonds (carbonyl, 287.1 eV), and carboxylic
molecule is 3.9 Å and 3.3 Å, respectively, at room temperature. COOH or OAC@O (288.5 eV) [32]. Table 1 also listed the relative
Therefore, large CO2 capture by physisorption is obtained with atomic percentage of CAC increased to 77.1–83.8% for GO and
adsorbents containing a large volume of micropores with a suitable GN-400, respectively, in comparison with 78.7% of graphite oxide
micropore size (2 nm) [38–41]. due to the removal of oxygen functional groups of graphite oxide
The micropore structures were enhanced predominantly by the by heat treatment. This indicates that the reduction of the heat
heat treatment temperature, and the distributions show that slight treatment deoxygenated sp2 carbon from the epoxides and
development can be observed around the micro-region compared hydroxyls. And, some amounts of oxygen functional groups were
to the pristine graphite flake and graphite oxide. Micropore devel- still retained on the graphene surfaces, and the GN-400 has fewer
opment of prepared graphene nanoplates is mainly on the pore re- oxygen functional groups than others.
gions <12.2 Å in diameter. The graphene nanoplates obtained with In this work, the aim is that improving the layer space of graph-
300 °C exhibited exceptional micropore development (supermi- ene nanoplates for CO2 adsorption. The CO2 adsorption capacities of
cropores: 11.7 Å). As shown in Fig. 5b, the prepared graphene the graphene nanoplates were measured from the CO2 adsorption
nanoplates showed a broad pore size distribution. The sample ob- isotherms at 25 °C (Fig. 7). The CO2 adsorption capacity of graphene
tained at mild temperatures (300 °C) was made up mostly of wider nanoplates at 25 °C and 1 bar increased from 1 wt.% to 6.36 wt.% at
pores distributed in two well-defined pore systems: one of them P/P0 = 1, as shown in Fig. 7a. The adsorbed weights (CO2) of all the
formed by uniform supermicropores (11.7 Å) and the other by graphene nanoplates were in the following order at the low relative
mesopores (20–500 Å). The heat treatment gave a broad meso-/ pressure range (P/P0 6 1): GN-300 > GN-200 > GN-400 > graphite
L.-Y. Meng, S.-J. Park / Journal of Colloid and Interface Science 386 (2012) 285–290 289

(a)

(wt.%, 298 K)
250 Zeolites
Graphite Activated Carbons
8
CO2 adsorbed (wt.%,1bar)
GO 200 GN-300
GN-150 G-1050
6 GN-200
GN-300 150
GN-400

CO 2 adsorbed
4 100

50
2
0
0
0 5 10 15 20 25 30 35
0.0 0.2 0.4 0.6 0.8 1.0 P/P0
P/P0
Fig. 8. CO2 adsorption isotherms of zeolite-13X, activated carbons, and prepared
graphene nanoplates (GN-300) measured at 30 bar and 25 °C.
(b) 400
CO2 adsorbed (wt.%,35 bar)

GN-200
GN-300 isotherms of these three materials. Zeolite-13 X, activated carbons,
300 GN-400
and conventional graphene nanosheets show maximum CO2
adsorption capacities at 35.1 wt.%, 67.9 wt.%, and 170.1 wt.%,
200 respectively. The advantage of the larger mesopore volumes in
the graphene nanoplates became more evident as the pressure
100 was increased, leading to a higher CO2 capture capacity of GN-
300 compared to that of zeolite-13X and activated carbon under
0 high pressure conditions (P/P0 > 1 bar).

0 10 20 30 40
4. Conclusions
P/P0
The CO2 adsorption behavior of exfoliated graphene nanoplates
Fig. 7. CO2 full isotherms of the pristine graphite flake, graphite oxide, and
graphene nanoplates measured at 1 bar (a) and 35 bar (b) and 25 °C. is related directly to the heat treatment temperature. In this study,
exfoliated graphene nanoplates were expanded successfully from
graphite oxide by a low-heat treatment in the temperature range
oxide > graphite. This is because the GN-300 sample had a higher of 150–400 °C under vacuum conditions. The heat treatment had
micropore volume of supermicropores than the other samples a significant effect on the property development of the graphene
and hence a narrower micropore size distribution, as shown in nanoplates studied. The inter-layer spacing of the graphene layers
Fig. 5a. At the middle and high relative pressure region (P/ and pore structure on the CO2 capture capacities were examined as
P0 > 30), GN-300 also showed maximum CO2 adsorption capacity. a function of the processing conditions. The prepared graphene
This was attributed to the possible multilayer adsorption of CO2 nanoplates exhibited high capture capacities up to 248 wt.% under
molecules in the pores at higher pressures. 25 °C and 30 bar. The CO2 capture capacities of the graphene nano-
In the case of CO2 capture, the pore width is a key factor in plates were improved by a larger inter-layer spacing and higher
developing a carbon adsorbent with high capacity. For post-com- interior void volume.
bustion CO2 capture using solid adsorbents, the mechanism of solid
physisorbents materials can be visualized as follows: Acknowledgments

CO2 þ Surface () ðCO2 Þ  ðSurfaceÞ


This work was supported by the Korea CCS R&D Center (KCRC)
grant funded by the Korea government (Ministry of Education, Sci-
where the selective adsorption of CO2 compared to N2 is caused by
ence and Technology) (0031985).
van der Waals attraction between the CO2 molecule and adsorbent
surfaces, as well as by pole-ion and pole-pole interactions between
the quadruple of CO2 and the ionic and polar sites of the adsorbent References
surfaces [38,39]. Accordingly, in this study, the CO2 molecules in the
[1] S.J. Davis, K. Caldeira, H.D. Matthews, Science 329 (2010) 1330–1333.
supermicropores were aligned mostly parallel to the pore surfaces [2] S. Chu, Science 325 (2009) 1599.
of the graphene nanoplates by a strong CO2-surface interaction. In [3] G.T. Rochelle, Science 325 (2009) 1652–1654.
the mesopores, the CO2 molecules are aligned multilayer to the pore [4] S. Choi, J.H. Drese, P.M. Eisenberger, C.W. Jones, Environ. Sci. Technol. 45 (2011)
2420–2427.
surfaces due to the weaker CO2ACO2 interactions [40–42]. [5] Z. Bao, L. Yu, Q. Ren, X. Lu, S. Deng, J. Colloid Interf. Sci. 353 (2011) 549–556.
As the pressure was increased, the physical properties of the [6] L.Y. Meng, S.J. Park, J. Colloid Interf. Sci. 366 (2012) 135–140.
adsorbent became increasingly important for CO2 adsorption [7] P.H. Stauffer, G.N. Keating, R.S. Middleton, H.S. Viswanathan, K.A. Berchtold,
R.P. Singh, R.J. Pawar, A. Mancino, Environ. Sci. Technol. 45 (2011) 8597–8604.
because a high micropore volume means more adsorption sites [8] J. Zhang, P.A. Webley, Environ. Sci. Technol. 42 (2008) 563–569.
available and a large pore volume means that more space available [9] M.T. Ho, G.W. Allinson, D.E. Wiley, Ind. Eng. Chem. Res. 47 (2008) 4883–4890.
for CO2 capture. In addition, the CO2 adsorption capacity of various [10] D.M. D’Alessandro, B. Smit, J.R. Long, Angew. Chem. Int. Ed. 49 (2010) 6058–
6082.
porous solids including commercial zeolite 13X (SBET = 839 m2/g,
[11] P. Nachtigall, L. Grajciar, J. Pérez-Pariente, A.B. Pinar, A. Zukal, J. Čejka, Chem.
V micro = 0.393 cm3/g, and Vtotal = 0.396 cm3/g), commercial Chem. Phys. 14 (2012) 1117–1120.
activated carbon (SBET = 1453 m2/g, Vmicro = 0.050 cm3/g, and [12] M. Sevilla, A.B. Fuertes, J. Colloid Interf. Sci. 366 (2012) 147–154.
Vtotal = 1.382 cm3/g), conventional graphene nanosheets (G-1050, [13] N. Jin, J. Seo, K. Hong, H. Chun, Micropor. Mesopor. Mater. 150 (2012) 147–154.
[14] A.M. Ribeiro, J.C. Santos, A.E. Rodrigures, S. Rifflart, Energy Fuels 26 (2012)
SBET = 701 m2/g, and Vtotal = 0.727 cm3/g) [43], and graphene nano- 1246–1253.
plates were tested for comparison. Fig. 8 shows the CO2 adsorption [15] A. Modak, M. Nandi, J. Mondal, Chem. Commun. 48 (2012) 248–250.
290 L.-Y. Meng, S.-J. Park / Journal of Colloid and Interface Science 386 (2012) 285–290

[16] H.Y. Hsiao, C.M. Huang, M.Y. Hsu, H. Chen, Sep. Purif. Technol. 82 (2011) 19– [31] Y. Zhu, S. Murali, M.D. Stoller, K.J. Ganesh, W. Cai, P.J. Ferreira, A. Pirkle, R.M.
27. Wallace, K.A. Cychosz, M. Thommes, D. Su, E.A. Stach, R.S. Ruoff, Science 332
[17] L.Y. Meng, S.J. Park, J. Colloid Interf. Sci. 352 (2010) 498–503. (2011) 1537–1541.
[18] G. Srinivas, J.W. Burress, J. Ford, T. Yildirim, J. Mater. Chem. 21 (2011) 11323– [32] L.Y. Meng, S.J. Park, J. Solid State Chem. 186 (2012) 99–103.
11329. [33] L.Y. Meng, S.J. Park, Bull. Korean Chem. Soc. 3 (2012) 209–214.
[19] F. Su, C. Lu, H.S. Chen, Langmuir 27 (2011) 8090–8098. [34] W.S. Hummers, R.E. Offeman, J. Am. Chem. Soc. 1 (1958) 1339.
[20] Y. Gogotsi, P. Simon, Science 334 (2011) 917–918. [35] A. Rani, S. Nam, K.A. Oh, M. Park, Carbon Lett. 11 (2010) 90–95.
[21] J. Zhang, J. Jiang, H. Li, X.S. Zhao, Energy Environ. Sci. 4 (2011) 4009– [36] B.J. Kim, S.J. Park, Int. J. Hydrogen Energy 36 (2011) 648–653.
4015. [37] J. Burress, S. Gadipelli, J. Ford, J.M. Simmons, W. Zhou, T. Yidirim, Angew.
[22] Y. Li, P. Zhang, Q. Du, X. Peng, T. Liu, Z. Wang, Y. Xia, W. Zhang, K. Wang, H. Zhu, Chem. Ind. Ed. 49 (2010) 8902–8904.
D. Wu, J. Colloid Interf. Sci. 363 (2011) 348–354. [38] A. Samanta, A. Zhao, Ind. Eng. Chem. Res. 51 (2012) 1438–1463.
[23] J. Schrier, ACS Appl. Mater. Interf. 3 (2011) 4451–4458. [39] J.P. Severinghaus, M.O. Battle, Earth Planetary Sci. Lett. 244 (2006) 474–500.
[24] L. Fan, C. Luo, X. Li, F. Lu, H. Qiu, M. Sun, Hazard. Mater. (2012). [40] E. Albrecht, G. Baum, T. Bellunato, A. Bressan, S.D. Torre, C. D’Ambrosio, M.
[25] A.K. Mishra, AIP. Adv. 1 (2011) 032152–032156. Davenport, M. Dragicevic, S.D. Pinto, P. Fauland, S. Ilie, G. Lenzen, P. Pagano, D.
[26] B.Z. Jang, A. Zhamu, J. Mater. Sci. 43 (2008) 5092–5101. Piedigrossi, F. Tessarotto, O. Ullaland, Nucl. Instrum. Method. Phys. Res. A 510
[27] A.K. Mishra, S. Ramaprabhu, J. Mater. Chem. 22 (2012) 3708–3712. (2003) 262–272.
[28] W. Lv, D.M. Tang, Y.B. He, C.H. You, Z.Q. Shi, X.C. Chen, C.M. Chen, P.X. Hou, C. [41] D.M. Ruthven, Principles of Adsorption and Adsorption Processes, Wiley-
Liu, Q.H. Yang, ACS Nano 24 (2009) 3730–3736. Interscience, New York, 1984.
[29] H.B. Zhang, J.W. Wang, Q. Yan, W.G. Zheng, C. Chen, Z.Z. Yu, J. Mater. Chem. 21 [42] Y. Liu, J. Wilcox, Environ. Sci. Technol. 46 (2012) 1940–1947.
(2011) 5392–5397. [43] M.J. McAllister, J.L. Li, D.H. Adamson, H.C. Schniepp, A.A. Abdala, J. Liu, M.
[30] L.L. Zhang, Xin Zhao, M.D. Stoller, Y. Zhu, H. Ji, S. Murali, Y. Wu, S. Perales, B. Herrera-Alonso, D.L. Milius, R. Car, R.K. Prud’homme, I.A. Aksay, Chem. Mater.
Clevenger, R.S. Ruoff, Nano Lett. 12 (2012) 1806–1812. 19 (2007) 4396–4404.

You might also like