Você está na página 1de 6

Adsorption study of CO

2
, CH
4
, N
2
, and H
2
O on an interwoven copper
carboxylate metalorganic framework (MOF-14)
Jagadeswara R. Karra, Bogna E. Grabicka, You-Gui Huang, Krista S. Walton

Georgia Institute of Technology, School of Chemical and Biomolecular Engineering, 311 Ferst Dr. NW, Atlanta, GA 30332, USA
a r t i c l e i n f o
Article history:
Received 6 July 2012
Accepted 8 October 2012
Available online 22 October 2012
Keywords:
Adsorption separation
Metal-organic frameworks
Catenation
Interwoven
MOF-14
Water adsorption
Flue gas
a b s t r a c t
Metalorganic frameworks (MOFs) are attractive microporous materials for adsorption separations due
to their extraordinary structures and impressive high surface areas. Catenation, or framework interpen-
etration, can signicantly impact the crystal stability and improve the adsorption interactions. This inter-
esting approach was used to obtain {[Cu
3
(BTB)
2
(H
2
O)
3
](DMF)
9
(H
2
O)
2
} (MOF-14) as a microporous
material with a high surface area and large pore volume, which are desirable parameters for adsorption
applications. Here, we report a detailed study of this catenated material with its gas adsorption proper-
ties. The potential for adsorption separations is evaluated by measuring pure-component adsorption iso-
therms for carbon dioxide, methane, and nitrogen. The Ideal Adsorbed Solution Theory (IAST) was used to
evaluate adsorption selectivities of MOF-14 for CO
2
/CH
4
and CO
2
/N
2
equimolar mixtures. In addition,
water adsorption and the impact of exposure on structural degradation are reported. Compared to other
open-metal site MOFs, MOF-14 adsorbs signicantly less water. This interwoven MOF is a promising
competitor to other MOF materials in the gas separation eld due to low interactions with water and high
selectivity for CO
2
over N
2
.
2012 Elsevier Inc. All rights reserved.
1. Introduction
Metalorganic frameworks (MOFs) have attracted signicant
attention in recent years for their potential applications in gas stor-
age, separations, and catalysis [15]. These materials have high
surface areas, high pore volumes, and highly tunable structural
properties. MOFs are synthesized via self-assembly with metal or
metal oxides and organic bridging linkers to form crystalline net-
works. The wide variety of metal vertices and organic ligands al-
lows for the design of materials with specic pore sizes and
volumes, resulting in porous solids with prescribed properties
[68]. MOFs containing unsaturated metal centers, or open-metal
sites, have been used to develop materials with improved adsorp-
tion capacity for gas storage and separations. Recently, special fo-
cus has been placed on carbon emission reduction in coal or
natural gas power plants via carbon dioxide separation from ue
gas mixtures. Selective adsorption and good thermal stability make
these microporous materials a potential candidate for CO
2
adsorp-
tion separations [5,911], but the impact of water vapor on these
features will also be important [1215]. The presence of open-
metal sites has a signicant impact on adsorption behavior since
it strongly favors the direct interaction between metal and adsor-
bate. Well-known MOFs, such as Mg-DOBDC and HKUST-1, pos-
sessing open-metal sites have shown among the highest
adsorption performance for CO
2
but are also known to adsorb up
to 30 mol/kg of water under exposure to 90% relative humidity
(RH) [10,12].
The modular approach of MOFs allows their pore size and shape
to be systematically tuned, not only by the choice of metal-contain-
ing secondary building units and/or bridging linkers, but also by dif-
ferent synthesis strategies. One strategy is the use of framework
interpenetration or interweaving to achieve smaller pore sizes and
stronger frameworks. The resulting MOFs possess high surface areas
and frameworks with smaller pore diameters and higher pore vol-
umes compared to the non-catenated forms, which enhance the
adsorption potential of MOFs for certain adsorbates [1620]. Chen
et al. reported such an interwoven MOF ({[Cu
3
(BTB)
2
(H
2
O)
3
]
(DMF)
9
(H
2
O)
2
or MOF-14, where 4,4
0
,4
00
-benzene-1,3,5-triyl-triben-
zoic acid = BTB) on a periodic minimal surface with extra-large
pores [16], but no adsorption data were reported at that time.
As shown in Fig. 1, MOF-14 possesses a dicopper paddle wheel
secondary building unit and BTB ligand. Each BTB is linked to three
paddle wheel building units, and each paddle wheel building unit
is connected to four BTB units. Such connections give rise to (3,4)-
connected net with the Pt
3
O
4
topology (Fig. 2a). This unique struc-
ture is a pair of identical nets that are interwoven with each other
(Fig. 2b). The rings of one net are penetrated by links of the other,
so that they are truly catenated. Each net is involved in numerous
pp and CH p interactions with those of the adjacent interpen-
etrated net. The pore limiting diameter and largest cavity diameter
0021-9797/$ - see front matter 2012 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcis.2012.10.018

Corresponding author. Fax: +1 404 894 2866.


E-mail address: krista.walton@chbe.gatech.edu (K.S. Walton).
Journal of Colloid and Interface Science 392 (2013) 331336
Contents lists available at SciVerse ScienceDirect
Journal of Colloid and Interface Science
www. el sevi er . com/ l ocat e/ j ci s
were estimated as 5.2 and 16.4 , respectively, using the geometric
approach algorithm [21]. The removal of the bound water mole-
cules gives rise to open-metal sites, which are important in
increasing binding strength between the framework and guest
molecules. However, these preferred interactions are magnied
in the case of water adsorption, and this competition for adsorp-
tion sites can lower the adsorption selectivity for the molecules
of interest. There is also a concern for the MOF structural integrity,
since water may degrade the framework [1215].
In this study, we have investigated the adsorption properties of
MOF-14 and the effect of water exposure on the structure. Pure-
component isotherms were measured for carbon dioxide, methane,
nitrogen, and water vapor. These isotherms were then used in the
Ideal Adsorption Solution Theory (IAST) to calculate mixture adsorp-
tion properties for CO
2
/CH
4
and CO
2
/N
2
. To the best of our knowl-
edge, an examination of the adsorption properties of this MOF has
not been presented yet. MOF-14 is shown to exhibit high CO
2
selec-
tivity over CH
4
and N
2
and is shown to adsorb just one-tenth of the
amount of water compared to Cu-BTC and Mg-MOF-74 [12].
2. Materials synthesis and characterization
All chemicals were used as received from commercial suppliers
without further purication. The reported synthesis procedure was
slightly modied [16]. A mixture of Cu(NO
3
)
2
3H
2
0 (30.2 mg,
0.125 mmol), BTB (54.8 mg, 0.125 mmol), and pyrazine (10 mg,
0.125 mmol) was dissolved in dimethylformamide (DMF) (5 ml)
with the addition of two drops of 1 N NaOH at room temperature.
The slurry was transferred to stainless steel reactors and
hydrothermally treated at 110 C for 96 h. Green cubic crystals
were obtained with the molecular formula {[Cu
3
(BTB)
2
(H
2
O)
3
]
(DMF)
9
(H
2
O)
2
}, conrmed by elemental and thermogravimetric
analyses and compared with the previously reported crystal struc-
ture [16]. The synthesized compound was solvent exchanged with
chloroform for 72 h to remove DMF.
Powder X-ray Diffraction patterns (PXRD) were recorded by
XPert X-ray PANalytical diffractometer with an Xaccelerator mod-
ule using Cu Ka radiation (k = 1.5418 ) at room temperature, with
a step size of 0.02 in 2h. Elemental analysis was performed by Des-
ert Analytics (Arizona, US). Thermogravimetric analyses (TGA)
were carried out in the temperature range of 25700 C on a NETS-
ZCH TG/Mass spectrometry analyzer, under owing helium with a
heating rate of 5 C/min. These results are given in the Supporting
information.
Nitrogen adsorption isotherms were measured at 77 K
(196 C) with a Quadrasorb SI analyzer from Quantachrome
Instruments. Solvent exchanged samples were degassed via
evacuation at 150 C under 10
5
Torr dynamic vacuum overnight
to remove residue from the pore structure. The Brunauer
EmmettTeller (BET) surface area was determined in the range
0.007 < P/P
o
< 0.05. CO
2
, N
2
, and CH
4
adsorption experiments were
carried out at 278 K, 298 K, and 318 K using the Intelligent Gravi-
metric Analyzer (IGA-001 series, Hiden Analytical Ltd.). The sam-
ples were activated overnight at 150 C under vacuum. Each
adsorption/desorption step was allowed to reach equilibrium over
a period of 2040 min, until no further weight gain was detected.
Water adsorption measurements were carried out at 298 K
(25 C) using an Intelligent Gravimetric Analyzer (IGA-003 series,
Hiden Analytical Ltd). The sample was activated prior to the run
overnight (150 C under vacuum). Dry air was used as the carrier
gas, with a portion of the carrier gas being bubbled through a ves-
sel of deionized water. The relative humidity (RH) was controlled
by varying the ratio of saturated air and dry air via two mass ow
controllers, where the total gas ow rate was 200 cc/min for the
duration of the experiment. The maximum humidity level exam-
ined was 90% RH due to condensation occurring inside the appara-
tus at higher humidities. Each adsorption/desorption step was
allowed to approach equilibrium over a period of 224 h.
3. Results and discussion
3.1. Structural characterization
Powder X-ray diffraction (PXRD) analysis of MOF-14 is shown in
Fig. 3. The samples were analyzed step-by-step, starting from the
Fig. 1. Coordination environment of MOF-14.
Fig. 2. Perspective view of MOF-14 showing (a) a single Pt
3
O
4
net; and the pair of interwoven Pt
3
O
4
nets with cavity size of (b) 5.19 ; and (c) 16.36 ; color scheme: brown
copper; red oxygen; yellow and green colors represent different frameworks. (For interpretation of the references to color in this gure legend, the reader is referred to the
web version of this article.)
332 J.R. Karra et al. / Journal of Colloid and Interface Science 392 (2013) 331336
as-synthesized MOF, and then followed by samples that were chlo-
roform exchanged, activated, water exposed, regenerated again
after water exposure, and nally as a resolvated compound. The
activated sample was obtained by drying the chloroform-ex-
changed sample at 150 C under vacuum for overnight, while the
regenerated sample was obtained by subjecting the water vapor
exposed sample to the same heat treatment conditions as activated
sample. The resolvated compound was obtained by immersing the
regenerated sample in DMF for two days. The PXRD pattern is al-
most identical with the simulated pattern for MOF-14, and the col-
lected cell parameters also correspond to the structure previously
reported by Chen et al. [16]. The PXRD pattern of the guest-free
phase and chloroform-exchanged phase presents the same struc-
ture as the as-synthesized compound. Furthermore, the PXRD of
the water-exposed sample was not signicantly different from
the activated sample, but some minimal changes in peak intensity
were observed after repeated thermal activation performed at
150 C.
Thermogravimetric analysis data show that the structure is
thermally stable to almost 300 C. The material releases its guest
molecules in the temperature range 25250 C to form the guest-
free phase (see Fig. S1). Elemental analysis was performed to
examine the purity of the MOF by analyzing the concentration of
carbon, hydrogen, and nitrogen elements and comparing it to a
theoretical MOF-14 sample. The calculated values of carbon,
hydrogen, and nitrogen are very close to the synthesized material
(Table S1), which conrms the good quality the sample. Nitrogen
adsorption measurements performed at 77 K (196 C) on the acti-
vated sample exhibit Type I isotherm behavior (Fig. S2), which is
typical for microporous materials with high surface area. The per-
manent porosity of MOF-14 was conrmed by the calculated BET
surface area and total pore volume of 1398 m
2
/g and 0.57 cm
3
/g,
respectively.
3.2. Pure-component adsorption isotherms at 298 K
Pure-component adsorption isotherms of CO
2
, N
2
, and CH
4
on
MOF-14 were measured at 298 K up to 20 bar. Fig. 4 shows the
adsorption isotherms with loadings following the order CO
2
>
CH
4
> N
2
. Carbon dioxide has the highest quadrupole moment of
the aforementioned gases (4.30 10
26
esu cm
2
), followed by
nitrogen (1.52 10
26
esu cm
2
), and nally nonpolar methane
[22]. However, the results do not exactly reect such a trend. Car-
bon dioxide adsorbs at signicantly higher loadings, but the load-
ings for the other two gases follow an opposite trend. Electrostatic
interactions play an important role in such gas adsorption behav-
ior. Methane has a higher polarizability in comparison with nitro-
gen, which causes higher surface interactions affected by polarity.
These trends have also been observed for Cu-BTC, which possesses
open copper sites similar to MOF-14 [23,24]. The methane adsorp-
tion isotherm is slightly lower than the simulated adsorption data
reported by Gallo et al. [19]. This overestimation is likely associ-
ated with crystal defects or impurities present in the experimental
samples, which are not accounted for in the model. MOF-14 ad-
sorbs CO
2
up to 2.5 mol/kg at 0.1 MPa and 298 K, which is lower
than the loading observed for Cu-BTC (5 mol/kg) [24] under the
same conditions due to the more easily accessible pore space in
Cu-BTC, which is a non-catenated MOF. However, the adsorption
capacities are higher than those of other MOFs such as IRMOF-1
(1.1 mol/g) [25], IRMOF-3 (1.3 mol/kg) [25], open-metal site
MOF Cu-BTB (2 mol/kg) [26], and DMOF (2.2 mol/kg) [24] under
the same conditions. At higher pressures, the desolvated MOF ad-
sorbs CO
2
up to 11 mol/kg at 2 MPa, due to its large pore volume
and surface area.
3.3. Analysis of heat of adsorption
The isosteric heat of adsorption (q
ST
) at zero loading reects the
strength of interaction between the framework and adsorbate mol-
ecule. The isosteric heat for CO
2
was calculated by applying the
ClausiusClapeyron equation (Eq. (1)) to isotherms taken at 278,
298, and 318 K (Fig. S3).
q
ST
RT
2

@ln P
@T

n
1
The isosteric heat as a function of CO
2
loading is shown in Fig. 5.
The heat of adsorption approaches 26 kJ/mol near zero loading and
decreases with increasing CO
2
loading. This indicates that the elec-
trostatic interactions between the quadrupole moment of CO
2
and
the exposed copper sites at low gas concentrations are dominant,
as seen for other open-metal site MOFs [9,17,18,23,27,28], and
the values are similar to those obtained for Cu-BTC [24].
3.4. Water vapor adsorption
Water adsorption measured on MOF-14 at 298 K is shown in
Fig. 6. The amount of water vapor adsorbed increases gradually
with a rise in the relative humidity. The isotherm is irreversible
and exhibits a large hysteresis loop. There is some amount of water
retained in the pores of MOF-14 after desorption, even after the
5 10 15 20 25 30 35 40 45 50
2 (

)
I
n
t
e
n
s
i
t
y
As synthesized
From crystal data
Chloroform Exchanged
Activated
Water Exposed
Resolvated
Regenerated after water exposure
Fig. 3. Powder X-ray diffraction patterns of as-synthesized, chloroform exchanged,
activated, water exposed, regenerated after water exposure, and resolvated samples
of MOF-14. Fig. 4. CO
2
, CH
4
, and N
2
adsorption isotherms on activated MOF-14 at 298 K.
J.R. Karra et al. / Journal of Colloid and Interface Science 392 (2013) 331336 333
sample is exposed to dry air at the nal stage of the experiment.
The water vapor adsorption capacity is 3 mol/kg (5.3 wt.%) at 90%
relative humidity. This is in sharp contrast to Cu-BTC and Mg-
MOF-74, which adsorb 30 and 35 mol/kg of water, respectively, un-
der the same conditions [12]. The amount of water irreversibly
bound in MOF-14 is approximately 1.45 mol/kg. This corresponds
to 0.84 water molecules per formula unit, which is equivalent to
0.28 water molecules per copper center. In comparison, water
loadings in Cu-BTC under the same experimental conditions result
in 0.69 water molecules per copper center [12].
In order to better understand the reason for the lower water va-
por uptake of MOF-14, the water vapor exposed sample was char-
acterized by nitrogen adsorption measurements performed at 77 K
and by PXRD to examine its possible structural degradation. The
PXRD patterns did not show major changes after water exposure
and regeneration. However, the patterns did indicate a decrease
in its peak intensities after water exposure and after regeneration
compared to its pristine activated form. This suggests that some of
the pores may have collapsed upon water exposure. The partial
pore collapse might have been from the weakening of some of
the copperoxygen bonds of the BTB linker upon interaction with
water vapor. However, it is important to note that PXRD analysis
may not show small transformations in material crystallinity.
Moreover, nitrogen adsorption analysis at 77 K reveals a signicant
surface area loss. The BET surface area was drastically reduced
from 1398 to 643 m
2
/g after water exposure. This suggests that
after water exposure in air and thermal regeneration, it is difcult
to completely remove the adsorbed water molecules, which are
strongly bound to copper atoms, without damaging the structure.
Similar behavior has been observed for Cu-BTC [18], which showed
26% surface area loss compared to 54% reduction in surface area for
MOF-14.
3.5. Binary mixture adsorption calculations
Adsorption equilibrium data and adsorption selectivities for
equimolar mixtures of CO
2
/CH
4
and CO
2
/N
2
were calculated using
the Ideal Adsorbed Solution Theory (IAST). IAST is a well-known
method for calculating mixture adsorption isotherms using pure-
component isotherm data. It has been shown to accurately predict
gas mixture adsorption in many zeolites and MOFs [29]. In calcula-
tion of adsorption selectivities for the binary mixtures through the
IAST method, the Toth equation was used as the best t for the
pure-component experimental data. Model parameters are given
in Table S2. The adsorption separation of CO
2
/CH
4
and CO
2
/N
2
mix-
tures is quantied by calculating the adsorption selectivity S
ij
= (x
i
/
x
j
)(y
j
/y
i
), where x
i
and y
i
are the mole fractions of component i in
the adsorbed and bulk gas phases, respectively.
Adsorption isotherms predicted by IAST for equimolar mixtures
of CO
2
/N
2
and CO
2
/CH
4
in MOF-14 as a function of total bulk pres-
sure are shown in Figs. 7 and 8, respectively. The existence of cat-
enation in this interwoven copper MOF affects gas adsorption
phenomena via enhancement of van der Waals interactions with
adsorbent molecules. The smaller pore diameter causes an increase
in adsorption energies between the adsorbate (gas) and adsorbent
wall, which leads to a porous material with more selective adsorp-
tion properties (Figs. 7 and 8). CO
2
is preferentially adsorbed over
N
2
and CH
4
over the entire pressure range considered. CO
2
adsorp-
tion is essentially unaffected by the presence of N
2
in the mixture
compared to the pure-component isotherm as shown in Fig. 7, but
the N
2
loadings are signicantly lower in the mixture. A slight de-
crease in CO
2
loadings is observed at higher pressures in the case of
CO
2
/CH
4
mixtures (Fig. 8). These differences in mixture behavior
compared to pure-component can be attributed to the types of
interactions occurring within the system. CO
2
will experience the
highest electrostatic interactions with the open-metal sites due
to the quadrupole moment, but van der Waals interactions due
to the catenated framework will also be signicant contributors
[22,24]. Neither CH
4
nor N
2
will interact as strongly with the
open-metal sites compared to CO
2
. Thus, at low pressure, where
sorbatesorbent interactions dominate, N
2
or CH
4
will not compete
effectively with CO
2
for adsorption sites. However, since the pore
Fig. 5. Isosteric heat of adsorption for CO
2
on activated MOF-14.
Fig. 6. Water vapor adsorption/desorption isotherm on activated MOF-14 at 298 K.
Fig. 7. Comparison of adsorption isotherms for CO
2
(red) and N
2
(blue) as pure-
components and in 50:50 CO
2
/N
2
mixtures (dotted line) on MOF-14. Mixture
isotherms are plotted with respect to partial pressure of the component. (For
interpretation of the references to color in this gure legend, the reader is referred
to the web version of this article.)
334 J.R. Karra et al. / Journal of Colloid and Interface Science 392 (2013) 331336
limiting diameter of catenated MOF-14 is only 5.1 , the enhanced
adsorption potential for CH
4
, or high van der Waals interactions, is
enough to increase the adsorption competition such that selectiv-
ities for the equimolar mixtures remain modest, and CO
2
loadings
are slightly depressed at high pressures compared to the pure-
component isotherm [24,26].
Adsorption selectivities for both mixtures are shown in Fig. 9 as
a function of total pressure. The selectivity of CO
2
over N
2
is 32 in
the low-pressure region and then decreases to 19 with increasing
total pressure. This trend occurs due to favorable interactions be-
tween CO
2
molecules and the open-metal sites at lower pressures.
As loadings increase, the sites become occupied and van der Waals
interactions will dominate the adsorption behavior, resulting in
greater competition between the mixture species [24]. The selec-
tivity for CO
2
/N
2
separation at 0.1 MPa is 22 and is slightly higher
than values reported for porous materials including zeolite 4A
(18.8) [30], activated carbon Norit R1 (15.3) [31], Cu-BTC (20)
[24], and MOF-508b (range 36) [32] at similar conditions. The
selectivity for CO
2
over CH
4
at 0.1 MPa is approximately 4. The ob-
served result is in the same range as other MOFs such as IRMOF-1
(range 23) [33], Cu-BTC (69) [33], MOF-508b (36) [32]. It ap-
pears that although CO
2
interacts more strongly with open-metal
sites compared to CH
4
, the small pore size from the catenated
framework and available pore volume result in an adsorption po-
tential that is overall favorable for the adsorption of both gases,
which depresses the selectivity [26]. It would be expected that
for CO
2
/CH
4
mixtures with higher CO
2
compositions, selectivities
will be higher than those exhibited here [26].
4. Conclusions
In summary, this study has demonstrated the signicance of
catenated metalorganic frameworks for potential applications in
gas separation. We have reported an adsorption study of CO
2
, N
2
,
and CH
4
and equimolar mixtures of CO
2
/N
2
and CO
2
/CH
4
on
MOF-14. We obtained high carbon dioxide loadings on the material
(2.5 mol/kg at 0.1 MPa bar and 298 K) due to its small pore size,
high pore volume, and high BET specic surface area. Mixture
adsorption results exhibited higher adsorption selectivities for
CO
2
over N
2
in comparison with the CO
2
/CH
4
gas pair. Moreover,
we examined the inuence of water exposure on MOF-14. PXRD
patterns appeared to be unchanged for the water-exposed samples,
but the BET surface area was signicantly reduced (54% loss). Nota-
bly, MOF-14 adsorbs less than one-tenth of the amount of water
adsorbed by other well-known MOFs with open-metal sites such
as Cu-BTC and Mg-MOF-74. Overall, the small pores and open-me-
tal sites of MOF-14 lead to high selectivity of CO
2
over N
2
, but the
water adsorption is relatively small. These features make MOF-14 a
promising porous material for CO
2
adsorption.
Acknowledgments
This material is based upon work supported by the National Sci-
ence Foundation under CBET Grant No. 1009682 and the Depart-
ment of Defense under PECASE Award W911NF-10-1-0079.
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.jcis.2012.10.018.
References
[1] J.R. Li, R.J. Kuppler, H.C. Zhou, Chem. Soc. Rev. 38 (2009) 1477.
[2] J. Lee, O.K. Farha, J. Roberts, K.A. Scheidt, S.T. Nguyen, J.T. Hupp, Chem. Soc. Rev.
38 (2009) 1450.
[3] J.L.C. Roswell, O.M. Yaghi, Micropor. Mesopor. Mater. 73 (2004) 3.
[4] G. Ferey, Chem. Soc. Rev. 37 (2008) 191.
[5] B.L. Chen, C.D. Liang, J. Yang, D.S. Contreras, Y.L. Clancy, E.B. Lobkovsky, O.M.
Yaghi, S. Dai, Angewandte Chemie (International ed.) 45 (2006) 1390.
[6] S. Kitagawa, R. Kitaura, S. Noro, Angewandte Chemie (International ed.) 43
(2004) 2334.
[7] O.M. Yaghi, M. OKeeffe, N.W. Ockwig, H.K. Chae, M. Eddaoudi, J. Kim, Nature
423 (2003) 705.
[8] H. Deng, C.J. Doonan, H. Furukawa, R.B. Ferreira, J. Towne, C.B. Knobler, B.
Wang, O.M. Yaghi, Science 327 (2010) 846.
[9] S.R. Caskey, A.G. Wong-Foy, A.J. Matzger, J. Am. Chem. Soc. 130 (2008) 10870.
[10] J. Liu, Y. Wang, A.I. Benin, P. Jakubczak, R.R. Willis, M.D. LeVan, Langmuir 26
(2010) 14301.
[11] Y.-S. Bae, O.K. Farha, J.T. Hupp, R.Q. Snurr, J. Mater. Chem. 19 (2009) 2131.
[12] P.M. Schoenecker, C.G. Carson, H. Jasuja, C.J.J. Flemming, K.S. Walton, Ind.
Chem. Eng. Res. 51 (2012) 6513.
[13] Z. Liang, M. Marshall, A.L. Chaffee, Micropor. Mesopor. Mater. 132 (3) (2010)
305310.
[14] J.J. Low, A.I. Benin, P. Jakubczak, J.F. Abrahamian, S.A. Faheem, R.R. Willis, J. Am.
Chem. Soc. 131 (43) (2009) 1583415842.
[15] J.A. Greathouse, M.D. Allendorf, J. Am. Chem. Soc. 128 (33) (2006) 10678
10679.
[16] B.L. Chen, M. Eddaoudi, S.T. Hyde, M. OKeeffe, O.M. Yaghi, Science 291 (2001)
1021.
[17] M. Klimakow, P. Klobes, A.F. Thuenamnn, K. Rademann, F. Emmerling, Chem.
Mater. 22 (2010) 5216.
[18] F. Millange, R.E. Osta, M.E. Medina, R.I. Walton, Cryst. Eng. Commun. 13 (2011)
103.
[19] M. Gallo, D. Glossman-Mitnik, J. Phys. Chem. C 113 (2009) 6634.
[20] N. Klein, I. Senkovska, I.A. Baburin, R. Gruenker, U. Stoeck, M. Schlichtenmayer,
B. Streppel, U. Mueller, S. Leoni, M. Hirscher, S. Kaskel, Chem. Eur. J. 17 (2011)
13007.
[21] E. Haldoupis, S. Nair, D.S. Sholl, J. Am. Chem. Soc. 132 (2010) 7528.
Fig. 8. Comparison of adsorption isotherms for CO
2
(red) and CH
4
(blue) as pure-
components and in 50:50 CO
2
/CH
4
mixtures (dotted line) on MOF-14. Mixture
isotherms are plotted with respect to partial pressure of the component. (For
interpretation of the references to color in this gure legend, the reader is referred
to the web version of this article.)
p (MPa)
0.0 0.5 1.0 1.5 2.0
C
O
2

S
e
l
e
c
t
i
v
i
t
y
0
5
10
15
20
25
30
35
40
50% CO
2
50% CH
4
50% CO
2
50% N
2
Fig. 9. Selectivity for CO
2
over N
2
(blue) and CO
2
over CH
4
(red) in equimolar
mixtures of CO
2
/N
2
and CO
2
/CH
4
on MOF-14 at 298 K. (For interpretation of the
references to color in this gure legend, the reader is referred to the web version of
this article.)
J.R. Karra et al. / Journal of Colloid and Interface Science 392 (2013) 331336 335
[22] X. Xu, X. Zhao, L. Sun, X. Liu, J. Natural Gas Chem. 17 (2008) 4.
[23] J.R. Karra, K.S. Walton, Langmuir 24 (2008) 86208626.
[24] J.R. Karra, K.S. Walton, J. Phys. Chem. C 114 (2010) 15735.
[25] A.R. Millward, O.M. Yaghi, Journal of the American Chemical Society 127
(2005) 17998.
[26] B. Mu, F. Li, K.S. Walton, Chem. Commun. 18 (2009) 2493.
[27] P.L. Llewellyn, S. Bourelly, C. Serre, A. Vimont, M. Daturi, L. Hamon, G. De
Weireld, J.S. Chang, D.Y. Hong, Y.K. Hwang, S.H. Jhung, G. Ferey, Langmuir 24
(2010) 7245.
[28] A. Demessence, D.M. DAlessandro, M.L. Foo, J.R. Long, J. Am. Chem. Soc. 131
(2009) 8784.
[29] S. Keskin, J. Liu, R.B. Rankin, J.K. Johnson, D.S. Sholl, Ind. Eng. Chem. Res. 48
(2009) 2355.
[30] E.D. Akten, R. Siriwardane, D.S. Sholl, Energy & Fuels 17 (2003) 977.
[31] F. Dreisbach, R. Staudt, J.U. Keller, Adsorption 5 (1999) 215.
[32] L. Bastin, P.S. Barcia, E.J. Hurtado, J.A.C. Silva, A.E. Rodrigues, B. Chen, J. Phys.
Chem. C 112 (2008) 1575.
[33] Q.Y. Yang, C.L. Zhong, J. Phys. Chem. B. 110 (2006) 17776.
336 J.R. Karra et al. / Journal of Colloid and Interface Science 392 (2013) 331336

Você também pode gostar