You are on page 1of 14

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/45801397

Can Metal-Organic Framework Materials Play a


Useful Role in Large-Scale Carbon Dioxide
Separations?

Article in ChemSusChem August 2010


DOI: 10.1002/cssc.201000114 Source: PubMed

CITATIONS READS

314 114

3 authors, including:

Seda Keskin
Koc University
83 PUBLICATIONS 2,539 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Gas Separation Peformances of Metal Organic Frameworks View project

All content following this page was uploaded by Seda Keskin on 18 September 2014.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
DOI: 10.1002/cssc.201000114

Can MetalOrganic Framework Materials Play a Useful


Role in Large-Scale Carbon Dioxide Separations?
Seda Keskin,[a] Timothy M. van Heest,[b] and David S. Sholl*[b]

Metalorganic frameworks (MOFs) are a fascinating class of surrounding the possibility of using MOFs in large-scale carbon
crystalline nanoporous materials that can be synthesized with dioxide separations. There are reasons to be optimistic that
a diverse range of pore dimensions, topologies, and chemical MOFs may make useful contributions to this important prob-
functionality. As with other well-known nanoporous materials, lem, but there are several critical issues for which only very
such as activated carbon and zeolites, MOFs have potential limited information is available. By identifying these issues, we
uses in a range of chemical separation applications because of provide what we hope is a path forward to definitively answer-
the possibility of selective adsorption and diffusion of mole- ing the question posed in our title.
cules in their pores. We review the current state of knowledge

1. Introduction

Any serious effort to reduce anthropogenic CO2 emissions approximates the concept of rational materials design. Molecu-
must contend with the geopolitical and economic reality that lar modeling can play an important role complementing exper-
fossil fuels will continue to make a dominant contribution to imental studies of MOFs, as described in two recent re-
the worlds energy supply for decades to come. This observa- views.[19, 20] In this article, we focus primarily on what is known
tion implies that scalable, cost-effective technologies for cap- about CO2 separations using MOFs from experiments.
turing CO2 from conventional uses of fossil fuels are of great To set the stage for discussing uses of MOFs for CO2 separa-
importance. In many cases, the key technical challenge is to tions, we first briefly consider how any nanoporous material
achieve the seemingly prosaic task of separating CO2 from could be used in these applications. It is useful to distinguish
other gases. CO2 capture from power plant flue gas, a mixture between two fundamentally different concepts for these sepa-
composed mainly of water-saturated N2 with smaller amounts rations: adsorption-based processes and kinetic separations.
of O2 and other species, is the best-known example. Other gas Adsorption-based separations rely on the fact that gases rever-
separations that could have relevance to mitigating CO2 emis- sibly adsorb in nanopores at densities that far exceed the den-
sions include removing CO2 from sour gas,[1] a process that sities of gases in equilibrium with the porous solid. To give a
could increase supplies of natural gasthe fossil fuel with the representative example, the density of adsorbed CO2 in zeolite
lowest carbon intensityand purification of O2 from air, a key 13X, a widely available adsorbent, in equilibrium with gaseous
element in oxyfiring;[2] an attractive alternative to conventional CO2 at 1 atm pressure and 298 K is 4.7 mmol g 1; this means
combustion. In each of these areas, there is a tremendous in- that at these conditions the density of CO2 in zeolite 13X is
centive to achieve the relevant separations on very large scales 73.7 times larger than the gas density.[21] Adsorbed gases can
with low cost in terms of capital outlay, maintenance, and the be removed either by reducing the gas phase pressure or by
energy requirements of the separation. increasing the temperature (or both). Sophisticated cyclic pro-
In this paper, we review studies considering metalorganic cesses based on varying these quantities are already used
frameworks (MOFs) for CO2-related gas separations. MOFs are a widely in industrial gas separations.[22] Although process design
relatively new class of crystalline nanoporous materials that for these applications is influenced by multiple factors, one key
combine metal organic centers with organic ligands to create metric is the adsorption selectivity for the desired gas relative
extended crystals with permanent porosity. For the purposes
of this Minireview, we include in this class materials that are [a] Dr. S. Keskin
Department of Chemical and Biological Engineering
variously described in the more detailed literature as coordina-
Ko University
tion polymers, ZIFs (zeolite imidazolate frameworks), COFs (co- Rumelifeneri Yolu, 34450, Saryer, Istanbul (Turkey)
valent organic frameworks), and others. Multiple reviews al- [b] T. M. van Heest, Prof. D. S. Sholl
ready describe the rapid developments in the synthesis of School of Chemical and Biomolecular Engineering
these materials in approximately the last decade.[318] Many Georgia Institute of Technology
311 Ferst Dr., Atlanta, GA 30332-0100 (USA)
MOFs have exceptionally high surface areas: examples with
Fax: (+ 1) 404 894 2866
> 4000 m2 g 1are known. Relative to more traditional nanopo- E-mail: david.sholl@chbe.gatech.edu
rous crystals such as zeolites, the synthesis of MOFs is typically Supporting Information for this article is available on the WWW under
more versatile and can be approached in a way that at least http://dx.doi.org/10.1002/cssc.201000114.

ChemSusChem 2010, 3, 879 891  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 879
D. S. Sholl et al.

to an undesired gas. For an arbitrary bulk phase composition, would have a moderate or low heat of adsorption, a large ca-
this quantity is simply Sads = (q1/q2)/(y1/y2) where qi (yi) is the pacity, and high selectivity. In Section 2.1 we examine experi-
mole fraction of species i in the adsorbed (bulk) phase. A ments that probed adsorption of CO2 as a pure component in
second key metric is the heat of adsorption, DHads, which con- MOFs. In Section 2.2, we look at experiments that have studied
trols the temperature required to release an adsorbed gas. the selectivity of CO2 adsorption relative to other gases.
Kinetic separations use differences in adsorption affinities
and differences in diffusivities of adsorbed species in a porous
2.1. Single-component adsorption of CO2 in MOFs
material. Although kinetic separations can be performed with
appropriate operation of a packed-bed adsorber,[22] they are A large number of studies have examined single-component
easiest to appreciate for membrane-based separations. When adsorption of CO2 in MOFs. For example, in our survey of the
permeation through a membrane occurs via a solutiondiffu- available literature, we found > 260 unique isotherm measure-
sion mechanism, the membrane selectivity can often be ap- ments for > 130 MOFs.[11, 17, 27107, 108] A list of these materials and
proximated by Smembrane = (q1/q2)  (D1/D2), where Di is the diffu- the conditions for available isotherm data is tabulated in
sion coefficient of species i in the membrane. It is useful to Table S1 of the Supporting Information. In light of this large
note that Smembrane = Sads  (D1/D2).[23] That is, efforts to select ma- amount of available data, we discuss carbon dioxide adsorp-
terials for kinetic separations have two variables that can con- tion data in terms of two characteristic quantities: the heat of
ceptually be adjusted independently, namely the adsorption adsorption and the adsorbed amount at atmospheric pressure.
and diffusion selectivities, unlike adsorption-based processes,
which have only the former. Impressive examples of this idea
2.1.1. Heat of adsorption
are already known. The pure silica zeolite DDR has Sads ~ 4 for
CO2/CH4 at room temperature, but DDR membranes have se- Although the heat of adsorption is not always reported when
lectivities > 100 because of the large values of DCO2/DCH4.[24] CO2 adsorption isotherms are measured, the heats of adsorp-
We illustrated the concepts of adsorption-based and kinetic tion of CO2 in MOFs are available for a large collection of mate-
separations using data from zeolites. This choice highlights the rials.[17, 31, 34, 37,38, 42, 43, 45, 55,57, 58, 62, 64, 6669, 74, 76, 82,92, 94, 97, 98, 109111] This data
fact that nanoporous materials, most notably activated carbons is summarized in Figure 1. In instances where a loading-depen-
and zeolites, have been used for gas separations for decades. dent heat of adsorption was reported, Figure 1 shows the
These materials are often physically robust and their prices value reported for the lowest loading. In essentially all cases,
vary with quality. For example, the price of natural zeolites
ranges from US$ 0.50 to US$ 4.50 per kilogram.[25, 26] The unde-
niable aesthetic appeal of MOFs, as well as their status as de-
signable nanomaterials, means that the level of research activ-
ity for MOFs in recent years is far higher than for traditional
materials. In taking a pragmatic view towards commercial ap-
plications, however, it is important to inquire with an open
mind whether the properties of MOFs genuinely improve upon
(or seem poised to improve upon) the older materials.
In the remainder of this Minireview, we examine the current
state of knowledge for MOFs applied to CO2 separations. Sec-
tion 2 considers adsorption-based separations, where the ma-
jority of work in this field has focused. Section 3 turns to the
relatively less explored realm of kinetic separations. We are Figure 1. Heat of adsorption for carbon dioxide in MOFs (blue) near ambient
cautiously optimistic that MOFs will make a useful impact on conditions (298 K, 1 atm). Non-MOF materials (red) are included for compari-
future CO2 separation technologies, but it is clear from the cur- son. The heat of liquefaction of bulk CO2 is presented as a horizontal dashed
rent literature that critical issues remain open, that must be ad- line at 17 kJ mol 1. See also Table S2 and S3 in the Supporting Information.

dressed for the field to move towards realistic applications. We


describe these challenges in Section 4.
the data in Figure 1 are from experiments performed close to
room temperature. Although the heat of adsorption and the
2. MOFs for Adsorption-based Separations of
isosteric heat of adsorption are slightly different quantities,[20]
CO2
many studies are unclear as to which quantity they report. For
In this Section, we review experiments that have been used to this reason, we did not attempt to distinguish between these
examine adsorption of CO2 in MOFs under equilibrium condi- two quantities in Figure 1. All of the data in Figure 1 is tabulat-
tions. In assessing the potential of an adsorbent for an adsorp- ed in Table S2. The ten MOFs with the highest reported heats
tion-based separation, information on the heat of adsorption of adsorption are listed in Table 1. For comparison, Figure 1
of the species of interest, the adsorption capacity for the spe- also indicates the heat of adsorption of CO2 in a range of non-
cies of interest, and the selectivity for the species of interest MOF materials, including liquid-phase amines and zeolites. This
relative to other components are needed. An ideal adsorbent data is listed in Table S3.

880 www.chemsuschem.org  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemSusChem 2010, 3, 879 891
Can MOF Materials Play a Useful Role in Large-Scale CO2 Separations?

range of CO2 loadings that is ex-


Table 1. The ten MOFs with the highest reported values for the initial (low loading) heat of adsorption for CO2.
See Tables S2 and S3 in the Supporting Information for a listing of all values shown in Figure 1. pected to be relevant in the ap-
plications of interest.
MOF Heat of adsorption Ref. In many cases, the heat of ad-
[kJ mol 1]
sorption of CO2 in a specific
Cu-BTTri(en) 90 [97] MOF has only been reported
Zn2(tetrakis[4-(carboxyphenyl)oxamethyl] methane)(4,4-bipyridine) 84 [62]
once. There are several materials,
MIL-100 62 [68]
USO-1-Al-A 50 [113] however, where multiple meas-
[Cu(PF6)2(bpetha)2]n 50 [92] urements are available, and
Mg/DOBDC/MOF-74 (Mg)/CPO-27-Mg 47 [57] these results highlight some im-
bio-MOF-11 45 [67]
portant aspects of dealing with
MIL-101 44 [68]
[Pd(m-H-pymo-N1,N3)2(H2O)m]n 44 [43] these materials. The reported
Ni/DOBDC/MOF-74 (Ni)/CPO-27-Ni 41 [57] heats of adsorption for CuBTC
(also known as HKUST-1) range
from 35 kJ mol 1[76] down to
It is clear from Figure 1 that MOFs can exhibit a large range 12 kJ mol . For IRMOF-1, values from 34[34] to 15 kJ mol 1[66]
1 [66]

of heats of adsorption for CO2. The highest value of have been reported and for CPO-27-Mg values of 47[57] to
90 kJ mol 1, for an ethylenediamine-functionalized material,[97] 39 kJ mol 1 have been measured.[69] Some of this variation can
exceeds the heat of adsorption for the best known liquid be attributed to differences in the methods used to extract a
amine, MEA.[112] Materials with heats of adsorption in this range heat of adsorption, or to the loading-dependence of the heat
are unlikely to be satisfactory in practical applications because of adsorption. Much of the variation, however, appears to be
of the large energy requirements associated with desorbing associated with genuine differences between the nominally
CO2. Most MOFs have heats of adsorption between 20 and identical materials that are being measured. The CuBTC results
50 kJ mol 1. This is not a surprising outcome; the data in mentioned above vary considerably in the approach used to
Table S3 shows that this statement is also true for zeolites. It is activate the MOF, that is, to remove residual solvent associat-
useful to note that these heats of adsorption exceed the heat ed with the materials synthesis from the MOF pores and possi-
of liquefaction of CO2, which is 17 kJ mol 1. bly other adsorbed species associated with the history of the
In a recent review of adsorbents for CO2 capture, Choi sample prior to the experimental measurement. One implica-
et al.[113] reviewed adsorption on a wide variety of materials tion of this situation is that the methods that are used for acti-
and showed that temperature and loading have a significant vating MOF samples should be carefully described in experi-
effect on both capacity and heat of adsorption. A subset of mental reports. A second implication is that caution must be
the studies we listed above provides information on the exper- exercised in comparing atomistic models of gas adsorption in
imentally measured heat of adsorption for CO2 on MOFs as a MOFs, which almost always invoke a solvent-free MOF,[19] to ex-
function of pressure[55, 98, 110, 113] or loading.[42, 43, 58, 67, 68, 76, 92, 97, 98, 111] perimental data for new materials where the possible role of
In many cases, the heat of adsorption decreases rapidly with MOF activation has not yet been thoroughly explored.
loading. For example, one amine-functionalized material has
an initial heat of adsorption close to 90 kJ mol 1, but the heat
2.1.2. Adsorption capacity for CO2 in MOFs
of adsorption quickly drops to less than 40 kJ mol 1 as loading
increases.[97] Physically, this occurs because the adsorption sites We now turn to the adsorption capacity of MOFs for CO2. The
with the highest affinity for CO2 are occupied first and the sites high degree of porosity of many MOFs means that they can
that are occupied at higher loadings have weaker affinities. hold very large amounts of CO2 or other gases if they are in
There are a small number of counterexamples to this general equilibrium with gas reservoirs at high pressures. That is, the
phenomenon where the heat of adsorption varies non-monot- absolute saturation capacity for CO2 isotherms in many MOFs
onically[55] or even smoothly increases[58] as a function of load- is large. This fact is not particularly relevant for practical appli-
ing. Snurr et al.[66] used simulations to calculate the initial heat cations of CO2 capture, because the pressure of CO2 that is
of adsorption for HKUST-1 and found a negative dependence available in these applications is modest. It is therefore consid-
of this quantity on temperature. They argued that in this case erably more important to examine the adsorption capacity of
the binding site that dominates the initial adsorption of CO2 MOFs (or other potential porous sorbents) under relatively
changes as a function of temperature because of an interplay mild conditions. To this end, we collected data from the litera-
between enthalpic and entropic effects. Bhatia and Myers have ture that reports the adsorbed amount of pure CO2 in MOFs
argued persuasively that having a strongly loading-dependent at room temperature and 1 atm.[14, 3234, 36, 37, 3943, 4648, 5760,
62, 63, 65, 67, 70, 71, 83, 95, 9799, 104, 106, 108, 115]
heat of adsorption, or equivalently, a strongly heterogeneous This set of data includes experi-
distribution of binding site energies, is not a useful property in ments ranging from 273 to 318.15 K, and pressures ranging
the sense of allowing design of efficient adsorption cycles.[114] from 0.85 to 1 bar. When not explicitly stated in the original
One obvious implication of this idea is that for measurements source, room temperature was taken to mean 298 K and stan-
of the heat of adsorption to be useful in considering practical dard pressure was taken to be 1 atm. This data is listed in
applications, data must be available that describes the full Table S4 in the Supporting Information and is summarized vis-

ChemSusChem 2010, 3, 879 891  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemsuschem.org 881
D. S. Sholl et al.

ually in Figure 2. For ease of comparison, all adsorbed amounts therms from several different studies, and shows variation
are given in mmol g 1. These values can be changed into cm3- from 2.1 to 1.12 mmol g 1at 1 bar and 298 K.[34]
STP g 1 (wt %) by multiplying by 24.4 (4.4). For comparison, The adsorption of CO2 in CuBTC illustrates the potentially
powerful influence of co-adsorbed species on the behavior of
MOFs. In its completely dehydrated form, CuBTC features pairs
of unsaturated Cu atoms. These open metal sites form strong
bonds with water molecules, so a hydrated form of CuBTC can
be created by exposing the material to low partial pressures of
water. The crystal structures of both the hydrated and dehy-
drated crystals have been carefully characterized.[116] In the hy-
drated material, one water molecule is bound to each Cu site.
Hydrated CuBTC shows the second highest CO2 capacity of the
materials summarized in Figure 2, holding > 8 mmol g 1.[59] The
dehydrated form under the same conditions, however, holds
only 45 mmol g 1.[14, 48, 59] It is not accurate to view the results
from the hydrated material as being representative of CO2 ad-
Figure 2. Amount of carbon dioxide adsorbed in MOFs (blue) near ambient sorption in CuBTC from a humid gas of the kind that may be
conditions (298 K, 1 atm). Two non-MOF materials (red) are included for
relevant in practical applications. The partial pressure of water
comparison. The top 10 highest values for MOFs are included in Table 2 and
a full listing of all data is provided in Table S4 of the Supporting Informa- that is necessary to hydrate the Cu sites in this material is
tion. almost vanishingly small from a practical point of view, so the
potential impact of additional physisorbed water that would
Figure 2 also shows data from several traditional porous mate- be present in more realistic scenarios cannot be deduced in a
rials, including zeolite 13X[57] and the activated carbon Norbit simple way from the available data. Nevertheless, it is encour-
RB2,[14] which adsorb 4.7 and 2.5 mmol g 1 CO2 at these condi- aging that at least in this one example, the addition of water
tions, respectively. The MOFs with the ten highest amounts of to the pores of the adsorbent improves the materials proper-
adsorbed CO2 from Figure 2 are listed in Table 2. ties as a CO2 adsorbent.
As with the heat of adsorption of CO2, a wide range of ad- Having compiled data for both the heat of adsorption and
sorption capacities at near-ambient conditions are seen in the ambient adsorption capacity of CO2 in MOFs, it is interest-
MOFs. The values shown in Figure 2 range from 8.48 mmol g 1 ing to explore whether a correlation between these two quan-
(37.3 wt %, 207 cm3-STP g 1) for Mg/DOBDC[108] down to tities exists. To this end, we show in Figure 3 and Table 3 the
0.16 mmol g 1 (0.7 wt %, 3.9 cm3-STP g 1) for ZIF-20.[106] For ma- values of these two quantities for materials where both the ini-
terials with relatively large heats of adsorption, the measured tial heat of adsorption and the amount of CO2 adsorbed under
adsorption capacity is of course closely related to temperature. ambient conditions were reported. Only studies that simulta-
To give one example, changing the temperature from 273 to neously reported both quantities for the same sample were in-
298 K drops the adsorption capacity at standard pressure for cluded in this comparison to eliminate any potential bias due
Bio-MOF-11[67] from 6 to 4 mmol g 1. The small number of to synthesis methods or activation procedures. When heat of
cases where adsorption capacity has been measured for one adsorption values were reported as a range, this range was
material by multiple groups illustrated the importance of syn- plotted in Figure 3 using an error bar. It is evident from
thesis and activation procedures on the overall outcome, as Figure 3 that these two quantities are not strongly correlated;
discussed above for the heat of adsorption. A compilation of for materials with heats of adsorption between 40 and
data for MOF-5 in work by Zhao et al. brings together iso- 50 kJ mol 1, adsorbed amounts vary from 1.68 mmol g 1. Per-
haps the most interesting observation from this analysis is that
the materials with the highest
heat of adsorption show low ad-
Table 2. The ten MOFs with the highest reported amount of CO2 adsorbed near ambient (STP) conditions. A sorption capacities at ambient
full list of the materials shown in Figure 2 is provided in Table S4 of the Supporting Information. conditions. It seems likely that
MOF Conditions Adsorbed amount Ref. this outcome is associated with
[mmol g 1] the rapid saturation of the high-
Mg/DOBDC/MOF-74 (Mg)/CPO-27-Mg] 298 K, 1 bar 8.48 [108] est energy binding sites at low
HKUST-1/Cu-BTC 298 K, 1 bar, 4 % hydrated 8.18 [59] CO2 pressures. This idea is sup-
Co/DOBDC/MOF-74 (Co)/CPO-27-Co 298 K, 1 bar 7.55 [108] ported by the rapid decrease in
Ni/DOBDC/MOF-74 (Ni)/CPO-27-Ni 298 K, 1 bar 7.15 [108] the heat of adsorption for CO2
bio-MOF-11 273 K, 1 bar 6.00 [67]
Zn/DOBDC/MOF-74 (Zn)/CPO-27-Zn 296 K, 1 atm 5.54 [57] as the CO2 loading is increased
CUK-1 298 K, 760 torr 3.48 [63] in these kinds of materi-
YO-MOF 273 K, 1 atm 3.39 [41] als.[42, 67, 68, 97]
SNU-M10 273 K, 1 atm 3.30 [70] A recent study by Yazaydin
H3[(Cu4Cl)3-(BTTri)8] 298 K, 1 bar 3.25 [97]
et al. examined 14 MOF struc-

882 www.chemsuschem.org  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemSusChem 2010, 3, 879 891
Can MOF Materials Play a Useful Role in Large-Scale CO2 Separations?

ative to CH4, CO, and N2, respectively.[117] Zeolite 5A gives


higher CO2 selectivities for dry gases than BPL Carbon, with re-
ported selectivities of 196, 59, and 331 for the same three gas
mixtures.[117] Not all zeolites are this selective; the selectivities
of zeolite 13X[118] and zeolite b[119] for CO2/CH4 are 224 and 28,
respectively. These widely available materials define several
cautionary notes related to examining data for MOFs. For ex-
ample, the strong heats of adsorption in zeolites such as 5A
and 13X make it difficult to regenerate these adsorbents with-
Figure 3. Amount of carbon dioxide adsorbed in MOFs near ambient condi- out significant heating, leading to high process costs.[55] Anoth-
tions (298 K, 1 atm) relative to heat of adsorption for materials where both
er disadvantage of these two zeolites is that they adsorb water
quantities have been measured. The numerical data for each material is
listed in Table 3. very strongly, significantly changing their adsorption properties
relative to those for dry gases.
If both species of an adsorbed
Table 3. Listing of materials and numerical data shown visually in Figure 3. mixture are present at low load-
ings (more precisely, within the
MOF Ambient amount Ambient con- Heat of Ref.
adsorbed ditions 1
adsorption [kJ mol ]
Henrys regime), the adsorption
[mmol g 1] selectivity relative to an equimo-
bio-MOF-11 4.1 298 K, 1 bar 45 [67]
lar bulk mixture is simply the
Co/DOBDC/MOF-74 (Co)/CPO-27-Co 6.95 296 K, 1 atm 37 [57] ratio of the Henrys constants for
Mg/DOBDC/MOF-74 (Mg)/CPO-27-Mg 8.00 296 K, 1 atm 47 [57] each species adsorbed as a pure
Ni/DOBDC/MOF-74 (Ni)/CPO-27-Ni 5.82 296 K, 1 atm 41 [57] component.[120] At non-dilute
H3[(Cu4Cl)3-(BTTri)8] 3.25 298 K, 1 bar 21 [97]
H3[(Cu4Cl)3-(BTTri8]-(en) 1.25 298 K, 1 bar 90 [97]
loadings, however, more infor-
Mg/DOBDC/MOF-74 (Mg)/CPO-27-Mg 6.82 298 K, 1 atm 40 [42] mation is required to describe
Ni/DOBDC/MOF-74 (Ni)/CPO-27-Ni 6.82 298 K, 1 atm 37 [42] multicomponent adsorption.
Ni-STA-12 2.60 298 K, 1 bar 35 [98] One common approach to de-
[Pd(m-H-pymo-N1,N3)2(H2O)m]n 1.64 293 K, 650 torr 41  3 [43]
1 3
[Pd(m-F-pymo-N ,N )2(H2O)m]n 2.49 293 K, 650 torr 44  4 [43]
scribing multicomponent ad-
[Fe(pz)Ni(CN)4] 2.48 298 K, 1 bar 30 [58] sorption in MOFs is to use ob-
Zn2(tetrakis[4-(carboxyphenyl)oxamethyl] meth- 1.14 298 K, 1 bar 84 [62] served single component ad-
ane)(4,4-bipyridine) sorption isotherms to predict
MOF-5/IRMOF-1 1.92 298 K, 1 bar 34 [34]
multi-component adsorption iso-
therms and selectivities using
Ideal Adsorbed Solution Theory
tures and found that heat of adsorption, which was computed (IAST). IAST is a well-developed technique to describe adsorp-
in their studies, was the best indicator of adsorbent perfor- tion equilibria for components in a gaseous mixture using only
mance at 0.1, 0.5, and 1 bar.[108] They also examined adsorbed data for the pure component adsorption without requiring
amounts at 0.5 and 1 bar, and found that available surface direct experimental measurement or simulation of mixture ad-
area and free volume did not correlate with adsorbed amount. sorption.[121] This approximate theory is known to work accu-
Still, they did not examine any functionalized materials on rately in many nanoporous materials except in materials that
their study, and these materials seemed to be the main reason have strong energetic or geometric heterogeneity.[122, 123] In
for the lack of strong correlation between heat of adsorption almost all examples where molecular simulations of mixture
and adsorbed amount in our analysis. adsorption in MOFs have been compared to IAST, the predic-
tions of IAST have been found to be accurate. These compari-
sons include Yang and Zhongs examination of CO2/CH4 mix-
2.2. Adsorption selectivity for CO2 in MOFs
tures in IRMOF-1, the similar calculations of Babarao and co-
All of the data discussed above was for adsorption of pure workers for CO2/CH4 mixture in the same material, and the re-
CO2. To understand whether an adsorbent can be useful in sults of Yang and co-workers for CO2/N2 and CO2/O2 mixtures
chemical separations, it is obviously critical to characterize the in CuBTC.[124126] In all of these examples, the heat of adsorption
materials performance in the presence of gas mixtures. In this for each species is not strongly loading-dependent. It is impor-
section, we describe experimental efforts that have studied tant to note, however, that IAST does not give accurate results
this issue for MOFs. for some cases; Yang and Zhong[124] and Keskin and Sholl[23]
showed that IAST gave poor results for CO2/H2 mixtures in
CuBTC. The discrepancy observed for this mixture was attribut-
2.2.1. Direct measurement of adsorption equilibrium
ed to the differing adsorption sites accessible to the two mo-
Activated carbons and zeolites have been widely used for se- lecular species inside the MOF.
lective separation of CO2. Henrys law selectivities on BPL The applications of IAST we just discussed used single-com-
Carbon at 303 K were reported as 2.5, 7.5, and 11.1 for CO2 rel- ponent isotherms calculated from molecular simulations. These

ChemSusChem 2010, 3, 879 891  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemsuschem.org 883
D. S. Sholl et al.

are the only examples where direct comparisons to multicom- the mixture can be estimated by integration of breakthrough
ponent adsorption are currently possible. We now turn to ap- curves. An average selectivity is then defined as the ratio of
plications of IAST involving experimentally measured single the adsorbed amounts normalized by the initial composition
component isotherms. This approach has been used for CO2/ of the bulk gas mixture.
CH4 adsorption in a number of MOFs. Mu et al.[32] predicted a A number of multicomponent breakthrough experiments
CO2/CH4 selectivity of 12.4 at 1 bar for an equimolar gas mix- with MOFs give only qualitative conclusions. Hayashi et al.[106]
ture for a structure with unsaturated Cu2 + ions, [Cu2(HBTB)2- reported breakthrough for a CO2/CH4 gas mixture using ZIF-20,
(H2O)(EtOH)]H2OEtOH. Lee et al.[99] incorporated an organic azo- showing that this material preferentially adsorbs CO2 from the
lium salt as a strut in a microporous MOF material to form a mixture. Dietzel et al.[42] studied members of the CPO-27-M
metalazolium framework and activated this material at differ- (M = Ni, Mg) series for separation of CO2/CH4 and CO2/N2 mix-
ent temperatures. IAST predictions showed that the MOF tures. For both mixtures the material preferentially adsorbed
which was activated at a higher temperature for a short time is CO2. Yoon et al.[63] explored separation of light gases over a
the most selective for adsorbing CO2 from equimolar CO2/CH4 column packed with CUK-1, showing that CO2 had the longest
mixtures, showing a CO2 selectivity of 20 at 1 bar at room tem- retention time among the gases in a mixture of H2, O2, N2, CH4,
perature. Bae et al.[44] examined the potential of a carborane- and CO2. None of the conclusions from the breakthrough ex-
based MOF for separating mixtures of CO2 and CH4 at room periments we have just mentioned are surprising; essentially
temperature at pressures up to 20 bar. Combining measured all nanoporous materials show preferential adsorption of CO2
single-component isotherms with IAST predicted CO2 selectivi- relative to other light gases.
ties of 731 for bulk gases that are 95 % CH4 and 510 for Some of the multicomponent breakthrough experiments
equimolar bulk gases. The relatively high CO2 selectivity of this with MOFs that have been analyzed quantitatively showed low
MOF was attributed to its open metal sites. Bae et al.[93] also levels of CO2 selectivity relative to other porous materials.
studied adsorption-based separation of CO2/CH4 mixtures in a Bastin et al.[31] examined separation of CO2 from CO2/N2, CO2/
mixed ligand paddlewheel MOF, (Zn2(NDC)2(DPNI), using a CH4 and CO2/CH4/N2 mixtures by MOF-508b. The reported se-
combination of experimental measurements, IAST, and molecu- lectivity of CO2 from CH4 and N2 at 303 K was 3-6, lower than
lar simulation. IAST predicted CO2 selectivities for a bulk mix- selectivity of activated carbon. Finsy et al.[127] studied separa-
ture of 50 % CH4 in the range of 430, higher than carborane- tion of CO2/CH4 mixtures using a fixed-bed packed with MIL-
based MOF, IRMOF-1, CuBTC, zeolites 13X and b. 53(Al). A CO2/CH4 selectivity of ca. 7 was found at pressures up
The experiments listed above focused on co-adsorption of to 5 bar. Adsorption of CO2 in MIL-53(Al) at higher pressures
CO2 from CH4. Bae and co-workers[60] studied separation of CO2 leads to expansion of the framework and an increase in the ad-
from N2 by measuring single-component adsorption isotherms sorbed amount of both gases.[128] As a result, the CO2 selectivi-
and predicting multicomponent adsorption by IAST. They ex- ty as assessed by breakthrough experiments decreased from
amined post-synthesis modification of a MOF[Zn2(1)(DMF)2]n- ca. 7 to ca. 4 at pressures above 5 bar.
(DMF)m, where (1) is 4,4,4,4- benzene-1,2,4,5-tetrayltetraben- A number of experiments showed that the adsorption selec-
zoic acidby replacing coordinated solvent molecules with tivity of several MOFs for CO2 outperforms traditional nanopo-
highly polar ligands to enhance CO2/N2 selectivity. They modi- rous materials such as BPL Carbon. Wang et al.[47] carried out
fied a 3D non-catenated Zn-paddlewheel MOF to form open breakthrough experiments with ZIF-95 and ZIF-100 by expos-
metal sites and to make a py-CF3 modified MOF. Single-compo- ing these materials to streams containing binary mixtures of
nent adsorption isotherms for CO2 and N2 were measured ex- CO2/CH4, CO2/CO, or CO2/N2. In both ZIFs, only CO2 was re-
perimentally in the original MOF and in its modified versions tained in the pores while the other gas passed through with-
to make a comparison. IAST was used to predict the adsorp- out hindrance; supporting the idea of using ZIF-95 and ZIF-
tion selectivities up to 8 bar at room temperature. The py-CF3 100 as selective CO2 reservoirs. They calculated CO2 selectivities
modified MOF showed CO2 selectivities of 1741. of ZIF-100 (ZIF-95) based on the ratio of Henrys constants of 6
(4), 17 (11), and 25 (18) from CH4, CO, and N2, respectively.
Banerjee et al.[71] studied ZIF-68, 69 and 70 for separation
2.2.2. Multicomponent breakthrough experiments
of CO2 from CO. Their preliminary breakthrough experiments
The experiments described above drew conclusions related to showed complete retention of CO2 and passage of CO through
adsorption selectivity based on measurement of single-compo- the pores of ZIF-68, 69, and 70 when they were exposed to
nent adsorption isotherms and application of IAST; none of streams containing a binary mixture of CO2/CO at room tem-
them directly observed multicomponent adsorption. Another perature. The CO2 selectivities from CO were reported based
useful measure of a materials CO2 separation capacity can be on the ratio of Henrys constants as 19.2, 20.9, and 37.8 for
obtained by exposing a bed packed with the material of inter- ZIF-68, 69 and 70, respectively. Banerjee and co-workers[49]
est to mixed gas streams including CO2 and detecting the ap- also studied separation of CO2/CH4, CO2/N2, and CO2/O2 using
pearance or breakthrough of CO2 from the material. Qualita- ZIF-78, 79, 81, and 82. They concluded that ZIF-78 and
tively, a clear difference in breakthrough time between two 82 have higher selectivities than the other ZIFs and BPL
species indicates the bed is selective for the gas that has a carbon for separation of CO2 from CH4, N2, and O2. For exam-
longer retention time. To draw quantitative conclusions from ple, the selectivity for CO2 over N2 (CH4) was estimated based
these experiments, the adsorbed amounts of each species of on the ratio of Henrys constants as ca. 50 (45) and ca. 35 (32)

884 www.chemsuschem.org  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemSusChem 2010, 3, 879 891
Can MOF Materials Play a Useful Role in Large-Scale CO2 Separations?

for ZIF-78 and ZIF-82, respectively. This was attributed to the sulting in a rapid elimination of excess CH4 molecules from the
presence of NO2 and CN groups in ZIF-78 and 82, respec- pores. Their mass balance calculation showed that essentially
tively, which have significant dipole moments that create no CH4 was adsorbed while 0.83 mmol g 1 of CO2 was ad-
dipolequadrupole interactions with CO2. Breakthrough experi- sorbed, indicating that a very high CO2 selectivity could be ob-
ments with CO2/CH4 mixtures for ZIF-78 and ZIF-82 showed tained at 1 bar. Experiments showed the retention time for CO2
that these materials had longer retention times for CO2 than and CH4 is 5.65 min and less than 0.1 at 30 8C, respectively.
BPL Carbon under the same conditions. Another interesting material that exhibits a high CO2 selec-
Britt et al.[69] performed breakthrough experiments using a tivity associated with framework flexibility is SNU-M10. Choi
20 % mixture of CO2 in CH4 to test Mg-MOF-74. The resulting and Suh[70] showed that the adsorption selectivity of SNU-M10
breakthrough curve showed that CO2 is highly preferred over for CO2 from N2 at room temperature is strongly pressure-de-
CH4. The separation capacity was determined by subtracting pendent. Specifically, the CO2/N2 selectivity of SNU-M10 at
the average CO2 concentration prior to breakthrough from the 298 K was reported as 24:1 (v/v) at 0.61 atm and 98:1 (v/v) at
amount of CO2 fed through the bed during that time. This 1.0 atm, based on the ratio of volumetric gas uptakes. This
gave a separation capacity of Mg-MOF-74 for CO2 of 8.9 wt %, phenomenon occurs due to CO2-induced gate opening and
very similar to the result obtained with zeolite NaX under iden- closing in the pores. This work also reported data on reversible
tical conditions. To test the importance of the metal ion in CO2 capture of CO2 in SNU-M10 at 298 K from a gas cycling experi-
adsorption, Britt et al.[69] also performed CO2 breakthrough ment on a thermogravimetric apparatus using a flow of 15 %
measurements on Zn-MOF-74, which differs from Mg-MOF-74 (v/v) CO2 in N2.
by substitution of the metal. Zn-MOF-74 took up just 0.35 wt %
CO2, a reduction of 96 % from Mg-MOF-74. Curiously, the ad-
2.2.3. Open issues for adsorption-based separations with
sorbed amount of CO2 in these two materials measured in
MOFs
other experiments do not differ so strongly (see Table 2)[57] .
The resolution of this apparent disagreement is unclear. Suc- We have summarized the large literature that now exists char-
cessive breakthrough experiments revealed that Mg-MOF-74 acterizing CO2 adsorption in MOFs. We conclude this section
retains a capacity of 7.8 wt %, 87 % of its intrinsic capacity, by briefly commenting on what information is not currently
after room temperature regeneration and further cycling did available. Very little is known about the co-adsorption of CO2
not lead to further reduction of capacity. with water, despite the ubiquitous nature of water vapor in
In addition to the experiments by Finsy et al. with MIL-53(Al) large-scale applications. We feel that this is the most critical
mentioned above, other experiments have examined other var- shortcoming of work to date in this area, and we return to it in
iants of MIL-53. Several studies have reported breakthrough Section 4. In addition to water vapor, realistic gas streams also
curves for variants of MIL-53. This class of materials is known typically contain trace levels of potentially reactive species
to have a bistable crystal structure, with a narrow pore struc- such as SOx and NOx. It appears that almost nothing is current-
ture at low CO2 pressures and a wider pore structure being ly known about the effects of these species on MOFs, particu-
stable at higher CO2 pressures.[55, 110] Hamon et al.[46] examined larly in humid environments. This situation will need to be re-
MIL-53(Cr) for separation of mixtures of CO2 and CH4 by grav- solved before the full potential of MOFs for large scale applica-
imetry and by breakthrough curves. Their gravimetric experi- tions can be fully assessed.
ments suggested that the narrow pore form of MIL-53(Cr) has The large amount of adsorption data that is now available
stronger affinity for CO2 and virtually excludes CH4. The selec- for dry gases in MOFs hopefully means that the research com-
tivity of MIL-53(Cr) measured experimentally is expected to be munity can move towards a nuanced view of future experi-
correlated to the large pore/narrow pore fraction, which de- ments. The fact that a new MOF preferentially adsorbs CO2 rel-
pends on both pressure and initial composition of the gas ative to CH4 or N2 is not intrinsically interesting; this is what es-
feed. Specifically, the exclusion of CH4 from the narrow pore sentially all nanoporous materials do. Similarly, if a new MOFs
structure should lead to very high CO2/CH4 selectivities in adsorption selectivity and/or capacity do not substantially im-
breakthrough experiments carried out under conditions in prove upon well-known and inexpensive adsorbents such as
which the narrow pore form should dominate. This outcome activated carbon, then it can be concluded without exhaustive
was not observed, however; breakthrough experiments analysis that the new material will not find widespread com-
showed CO2 selectivities in the range of 216 for different mercial use.
CO2/CH4 compositions. It seems likely that this was due to a ki-
netic barrier for the large pore to narrow pore transition,
3. MOFs for Kinetic Separations of CO2
which was not completed during the breakthrough experi-
ments. Couck et al.[39] tested an amine-functionalized MIL- We described in Section 1 how chemical separations with
53(Al) in breakthrough experiments using an equimolar CO2/ porous materials can be broadly divided into adsorption-based
CH4 mixture. They concluded that CH4 travels rapidly through methods and kinetic separations. Kinetic separations can be
the column, weakly adsorbing in the pores without causing achieved using packed beds, but the most important applica-
framework contraction, whereas CO2 adsorbs in the open tion of this concept is in membranes. Compared to the efforts
pores and replaces pre-adsorbed CH4 molecules. Beyond a cer- that have been devoted to characterizing adsorption in MOFs,
tain CO2 intrapore concentration, pore contraction occurs, re- exploration of the use of MOFs in membranes is only just be-

ChemSusChem 2010, 3, 879 891  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemsuschem.org 885
D. S. Sholl et al.

ginning. Below, we review what is currently known about membrane was up to 120 % larger than the pure polymer due
MOF-based membranes for CO2 separations in two categories: to the porosity of IRMOF-1 crystals.
Hybrid (composite) membranes with MOFs and pure MOF Adams et al.[137] incorporated CuTPA MOF particles into PVAc
membranes. (polyvinyl acetate) to create 15 % (w/w) CuTPA-PVAc hybrid
membranes. Both CO2 permeability and ideal selectivity were
enhanced with hybrid membranes. The hybrid membrane
showed a slightly higher CO2 permeability (3.26 Barrer) than
3.1. Hybrid membranes with MOFs for separation of CO2
the pure PVAc membrane (2.44 Barrer). The ideal selectivity for
Hybrid membranes are heterogeneous membranes in which CO2 over N2 (CH4) increased from 32.1 (34.9) in polymeric mem-
organic/inorganic fillers are embedded into a polymer matrix brane to 35.4 (40.4) in the hybrid membrane. Diaz et al.[138]
to enhance the selectivity and/or permeability of a pure poly- used pulse field gradient NMR techniques to study transport of
meric membrane. Hybrid membranes have been prepared with CO2 in a pristine poly(1,4-phenylene ether-ether-sulfone) mem-
various fillers such as zeolites and carbon molecular sieves in brane and hybrid membranes including ZIF-8 particles in this
the past to exceed the selectivity/permeability upper bound polymer. They also performed permeation experiments and re-
established for polymeric membranes.[129, 130] The most signifi- ported an increase in CO2 permeation relative to the pure poly-
cant advantage of hybrid membranes is that fabricating these mer.
membranes on large scales can readily be envisioned with rela- Although these initial studies of MOFpolymer hybrid mem-
tively minor adaptation of existing commercial technology. branes have not delivered a device with spectacular properties,
Polymer membranes with active surface area of hundreds of they do hint that this direction will be productive in the future.
thousands of square meters are already routinely used in in- Extensive previous studies of hybrid membranes using zeolites
dustrial applications such as water desalination and reverse os- as fillers have shown that achieving good compatibility be-
mosis,[131, 132] and significant research efforts are already in tween filler particles and the polymer matrix is both crucial
place to study methods of economical implementation of this and challenging to achieve.[129, 130] The results from MOF-poly-
technology for flue gas capture.[132, 133] mer hybrid membranes to date suggest that these compatibili-
Several groups have used MOF particles as fillers in polymer- ty issues may be easier to handle with MOFs than with other
ic membranes and examined the performance of these hybrid kinds of filler particles. For example, Zhang et al.[135] reported
membranes for CO2 separations. The hybrid membranes are that it was relatively easy to make defect free hybrid mem-
typically described in terms of the loading of filler particles in branes with MOFs due to the good compatibility between the
the polymer matrix. Car et al.[134] characterized MOF-based MOF crystals and the polymer matrix.
hybrid membranes using a time lag apparatus. They studied A critical question for the future development of this area is
two different polymers, poly(dimethylsiloxane) (PDMS) and whether MOF-polymer hybrids can be found whose properties
polysulfone (PSf), and two MOFs, CuBTC and Mn(HCOO)2, as greatly improve on the polymer alone. Two of us have recently
filler particles in the polymers for separation of CO2 from N2 introduced the first quantitative models of MOFpolymer
and CH4. Very slight improvements in the ideal selectivity for membranes,[139] and our results suggest that the observations
CO2 over N2 at 10 % MOF loading and CO2 over CH4 up to 30 % from the materials that have been studied to date are not (in
MOF loading of CuBTC in PDMS were observed. For example, retrospect) surprising because the intrinsic kinetic selectivity
the selectivity for CO2 over N2 (CH4) was increased from 8 (3) in for the MOFs used to date is low. This outcome does not
PDMS membrane to 9 (4) in the hybrid membrane. Hybrid appear to be true for all MOFs, however. Detailed models of
membranes with a 5 % loading of Mn(HCCO)2 in PSf showed small pore MOFs have predicted that materials exist that have
essentially the same CO2 selectivity (ca. 20) with the pure PSf very high kinetic selectivities for CO2/CH4 separations based on
membrane. large differences in the diffusion coefficients of the two spe-
Zhang et al.[135] examined membranes composed of the MOF cies.[140] Models that combine this information with experimen-
Cu-BPY-HFS and Matrimid, a commercial polymer, for separa- tal data from readily available polymers suggest that hybrid
tion of CO2/CH4 and H2/CO2 mixtures. A hybrid membrane with MOFpolymer membranes made from these materials will
20 % Cu-BPY-HFS loading had a lower CO2 selectivity (20.5) have both selectivity and permeability for CO2 that strongly ex-
than a pure Matrimid membrane (36.3) for an equimolar CO2/ ceeds the pure polymer membrane[139] .
CH4 mixture. This decrease in CO2 selectivity was attributed to
high CH4 affinity of Cu-BPY-HFS. No significant change was ob-
3.2. Thin-film MOF membranes for separation of CO2
served for selectivity of H2/CO2 mixtures relative to a pure Ma-
trimid membrane. Perez et al.[136] combined IRMOF-1 with Ma- An alternative to using nanoporous materials as fillers in poly-
trimid to study separation of CO2/CH4 and H2/CO2 mixtures. mer membranes is to fabricate membranes in which the active
The results from this hybrid membrane were similar to the Cu- layer is a dense film of the nanoporous material alone. This ap-
BPY-HFS/Matrimid membrane discussed above: the MOF de- proach has been widely studied for zeolite membranes, where
creased the selectivity of the membrane for CH4/CO2 mixtures intergrown films of zeolite crystals are typically fabricated on
and caused no appreciable change in selectivity for an H2/CO2 top of a porous support such as alumina or stainless steel.[141]
mixture. Although the MOF did not lead to an increase in the The scale-up of thin-film membranes of this type to the very
membranes CO2 selectivity, the CO2 permeability of the hybrid large surface areas that will be required for large-scale applica-

886 www.chemsuschem.org  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemSusChem 2010, 3, 879 891
Can MOF Materials Play a Useful Role in Large-Scale CO2 Separations?

tions remains a severe challenge. Nevertheless, these mem- amples with truly outstanding properties have yet been found.
branes continue to receive a great deal of attention because of If work in this area is to progress rapidly, it will be important
their excellent performance in CO2 separations (among other to use knowledge about the diffusion rates of molecules inside
chemical separations). Several zeolite membranes with very MOFs to guide choices of which MOFs are the most likely to
high CO2 selectivity for CO2/CH4 mixtures are known. For exam- be useful. Unfortunately, accurate characterization of molecular
ple, DDR (SAPO-34) membranes have been demonstrated with transport inside nanoporous materials using experimental
a CO2 selectivity of 100 (270) from equimolar CO2/CH4 mixtures techniques is very challenging. As a result, most of what is cur-
at 303 (255) K and 5 (2.22) bar.[142, 143] rently known about molecular diffusion of CO2 in MOFs has
The fabrication of thin-film MOF membranes has only been come from molecular simulation studies.[147150]
explored very recently. The first continuous and well-inter- Recently, several experimental techniques have been used
grown IRMOF-1 membrane was prepared on a porous alumina to assess CO2 diffusion in MOFs. Salles et al.[151] used a combi-
support by in situ solvothermal synthesis by Liu et al.[144] Gas nation of quasi-elastic neutron scattering (QENS) measure-
permeation results showed that diffusion of CO2, CH4, H2, N2, ments and molecular dynamics simulations to examine the
and SF6 through an IRMOF-1 membrane mainly follows a transport diffusivity of CO2 in MIL-53 (Cr). Their experiments
Knudsen diffusion scaling. Bux et al.[145] prepared a continuous and simulations were in fair agreement, and indicated that CO2
intergrown layer of ZIF-8 on a porous titania support and mea- has a one-dimensional diffusion mechanism in both the
sured the volumetric flow rates of single gases: CO2, H2, O2, N2, narrow pore and large pore structures of MIL-53(Cr). Salles and
and CH4. The permeance of H2 was almost five times faster co-workers[152] used the same methods to characterize the self
than that of CO2. Ranjan and Tsapatsis[146] prepared films of Cu- and transport diffusivities of CO2 in the rigid MOF MIL-47(V). In
(hfipbb)(H2hfipbb)0.5 on porous alumina discs and tested this both of these materials, the magnitude of the transport diffu-
membrane for single-gas permeation of CO2, N2, H2, He, n- sivities were found to be 10 8 m2 s 1. Zhao et al.[34] studied the
butane, and iso-butane. Their Cu(hfipbb)(H2hfipbb)0.5 mem- kinetics of adsorption and diffusion of CO2 in IRMOF-1, obtain-
brane exhibited an ideal selectivity of ca. 4 for H2 over CO2 be- ing CO2 diffusion coefficients of 511  10 13 m2 s 1 at 295
tween 25200 8C. The net gas fluxes through this membrane 331 K. The difference between the diffusivities measured in
were low for all of the species tested, and the authors specu- IRMOF-1 and MIL-47 is surprisingly large. For comparison, a
lated that efficient molecular transport through the films may typical diffusivity for a small molecule in a liquid is 10 9 m2 s 1.
be severely reduced by misoriented crystals that block the A critical lesson from the successful examples of highly se-
one-dimensional pore channels in the membrane. lective zeolite membranes for CO2 separations has been that
The experiments listed above examined pure gas permea- an effective membrane shows rapid diffusion of one species
tion through MOF thin films. A limited number of MOF mem- (typically CO2) but very slow diffusion of other gases such as
branes have been tested for separation of gas mixtures. Guo CH4.[24, 142, 143] For the use of MOFs as components in mem-
et al.[28] prepared a Cu(BTC) membrane on a porous copper branes to be successful, the identification of MOFs that have
support using a twin copper source method to study separa- this property is highly desirable. It appears that many, perhaps
tion of H2/CO2 mixtures. They reported a H2 selectivity of 6.84 most, MOFs do not have this property because the pore vol-
from equimolar H2/CO2 mixtures, a value somewhat higher umes that control overall molecular diffusion are significantly
than the ideal selectivity of 4.52. Li et al. studied H2/CO2 per- larger than typical gas molecules.[153] There are examples, how-
meation through ZIF-7 films using membranes fabricated with ever, where MOFs can be expected to have large differences in
seeded secondary growth method on an alumina support. The diffusion coefficients among adsorbed gases. A recent model-
H2/CO2 ideal selectivity and mixture selectivity were 6.7 and ing study that combined molecular simulations with quantum
6.5, respectively. chemistry calculations gave one example of this kind.[140] All as-
The large body of work associated with development of pects of the use of MOFs in membranes for CO2-related separa-
thin-film zeolite membranes makes it clear that the overall per- tions will move forward more quickly if experimental and mod-
formance of these membranes is affected by both the intrinsic eling efforts are focused on materials that have the pore char-
properties of the zeolite crystals that make up the membrane acteristics that allow the possibility of strong diffusion-based
and the microstructure of the intergrown crystals that form selectivity rather than on large pore materials where the diffu-
the membrane.[141] The same set of issues will also be critical in sion coefficients of adsorbed gases can vary only relatively
membranes made from thin films of MOFs. The development weakly from species to species.
of these membranes is currently in its earliest stages, so it is
not yet possible to draw any definitive conclusions about
4. Conclusion and Outlook
whether thin-film MOF membranes will be viable for practical
separations. There are several reasons to be optimistic about the potential
of MOFs for CO2 separations. A huge diversity of MOFs is
known. Avenues towards scalable applications in packed beds
3.3. Characterizing CO2 diffusion in MOFs
or as components in high-surface-area membranes can be rea-
For both kinds of membranes addressed above, the current ex- sonably envisaged. Computational tools for focusing searches
perimental state of the art is that initial demonstrations that through the large space of possible materials to find the small
MOFs can be used in membranes have been made, but no ex- fraction of the best-performing materials are becoming avail-

ChemSusChem 2010, 3, 879 891  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemsuschem.org 887
D. S. Sholl et al.

able. There are, however, critical issues that will severely Addressing this question should include both chemisorption
hamper development of real applications if they are not ad- of water to open metal sites, if they exist, but also the coad-
dressed in a timely way. Below, we describe these issues. sorption of physisorbed water in the MOFs pores. No serious
By far the most critical issue is, to paraphrase Coleridges An- consideration of how a material could be used in a real appli-
cient Mariner, water, water everywhere, but not a drop of data. cation is possible if this question is not answered. In our view,
Real CO2 separations must deal with gas streams containing, progress in this field will stall unless these questions are tack-
and often saturated by, water. It is extraordinarily unlikely that led. Phrased more positively, research teams that successfully
any process that dries flue gas before treating it for CO2 cap- resolve these questions will open up opportunities that may
ture could be economically feasible. permanently elude others who do not address them.
The behavior of MOFs in hydrated conditions water varies The interactions between water and CO2 inside pores of
widely, from materials that irreversibly degrade due to water MOFs will not necessarily prevent these materials from being
under mild conditions[154] to examples that are highly stable in useful in practical applications. One example, CuBTC, has al-
boiling water,[15, 97] and the literature on systematic studies of ready been studied where the hydration of open metal sites in
water adsorption in MOFs is limited. Experimental studies of Li the MOF framework increases the adsorbed amount of CO2 at
and Yang[155] indicated that MOF-177 can adsorb up to 10 wt % mild conditions relative to the dehydrated material (see discus-
H2O at room temperature, but this framework is not stable sion in Section 2). In amine-containing materials, the presence
upon H2O adsorption, and MOF-177 decomposes after expo- of water increases the variety of complexes that can form be-
sure to ambient air in 3 days. Kondo et al.[156] studied water ad- tween CO2 and amines,[113] and there is no reason to suspect
sorption isotherms at 303 K on water-resistant three-dimen- that these interactions will always have a negative effect. The
sional pillared layer MOFs. lack of attention to these issues, however, is a critical limitation
Work specifically focusing on MOF stability has also been of work to date in this area.
sparse. Two notable studies are the high throughput analysis Aside from coadsorption of CO2 and water, it is striking how
of Low et al. and the modeling work of Greathouse and Allen- little data is available regarding the coadsorption of any gas
dorf. Low et al.[157] found that the stability of MOFs in steam at mixtures in MOFs. Our literature survey did not find any meas-
various saturations and temperatures corresponded well with urements that directly characterize multicomponent adsorp-
the dissociation energy of the metalligand bond. Greathouse tion equilibrium involving CO2 in any MOF, although there are
and Allendorf[154] found that ligand displacement at metal sites a sizeable number of packed-bed breakthrough experiments
is a likely mode of dissociation. that give some insight into these issues. There is no doubt
These studies have been beneficial in opening the door to that detailed measurements of multicomponent adsorption
understanding MOF interactions with water, but more system- equilibria are more challenging than pure component iso-
atic studies are needed to gain understanding of the compet- therms. For materials that are judged to be promising for real
ing nature or assisting of water co-adsorption and the reasons applications, it would be fruitful to obtain data of this kind to
for MOF hydrothermal stability and instability. Multiple papers directly test the simple mixing theories such as IAST that are
have suggested that MOFs with open metal sites have favora- typically assumed to be valid in current descriptions. This issue
ble properties for dry gas separations. In all examples we are may be especially important in materials where single-compo-
aware of, however, water binds strongly to these open metal nent isotherms are strongly influenced by framework flexibility
sites at bond strengths that are characteristic of chemisorption, in the MOF.
not physisorption. Unless convincing evidence is given to the No experimental data at all is currently available regarding
contrary for a particular material, it therefore appears reasona- multicomponent diffusion in MOFs. In our view, collecting data
ble to assume that these open metal sites will be fully hydrat- of this kind for MOFs where diffusion of all components of an
ed under essentially any conditions of practical interest for CO2 adsorbed species is expected to be facile would have limited
separations. One interesting report on this interaction comes value. If, in contrast, experiments to gather data of this kind
from the work of Yazadin et al.,[108] where the authors found were performed on MOFs that are expected to show strong
that adsorption of carbon dioxide actually increased at low diffusion-based selectivity for CO2 relative to other adsorbed
water vapor concentrations. Additionally, a number of studies species could potentially play a key role in advancing develop-
have pointed out the flexibility of MIL-53 (Cr) in the presence ments of MOF-based membranes.
of water, which results in a structural switching between a In considering MOFs for adsorption-based separations, infor-
large pore and a narrow pore form. mation on the reproducibility of sample properties and the re-
These observations of interactions with water provoke two sponse of materials to cycling is urgently needed. We have al-
questions that should be answered for any MOF being consid- ready discussed several examples where careful studies by sev-
ered for CO2 separations: eral groups have highlighted the role of solvent removal and
Question 1: Is the MOF stable under long term exposure to activation of MOFs on their adsorption properties. Forward
low temperature steam? progress in the field will be aided by clear descriptions of acti-
If the answer to this question is no, further efforts to devel- vation procedures and open discussion of challenges found in
op the material for CO2 separations will be fruitless. reproducing data either within a single laboratory or when
Question 2: How are the separation characteristics of the making comparisons with literature data. The number of MOFs
material for CO2 affected by the presence of water vapor? for which detailed adsorption measurements have been per-

888 www.chemsuschem.org  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemSusChem 2010, 3, 879 891
Can MOF Materials Play a Useful Role in Large-Scale CO2 Separations?

formed by multiple groups is extremely limited compared to [3] K. Uemura, R. Matsuda, S. Kitagawa, J. Solid State Chem. 2005, 178,
the number for which a single published experiment is avail- 2420.
[4] T. Uemura, N. Yanai, S. Kitagawa, Chem. Soc. Rev. 2009, 38, 1228.
able. In studies of reversible hydrogen storage in solids, efforts [5] G. Frey, C. Serre, Chem. Soc. Rev. 2009, 38, 1380.
to characterize large number of loading/unloading cycles play [6] S. L. James, Chem. Soc. Rev. 2003, 32, 276.
an influential role in demonstrating material stability.[158] Similar [7] U. Mueller, M. Schubert, F. Teich, H. Puetter, K. Schierle-Arndt, J. Pastre,
efforts that demonstrate the stability of MOFs during cycling J. Mater. Chem. 2006, 16, 626.
[8] A. U. Czaja, N. Trukhan, U. Muller, Chem. Soc. Rev. 2009, 38, 1284.
for adsorption-based separations have considerable [9] O. M. Yaghi, H. L. Li, J. Am. Chem. Soc. 1995, 117, 10401.
value.[127, 159] [10] H. Li, M. Eddaoudi, M. OKeeffe, O. M. Yaghi, Nature 1999, 402, 276.
Attentive readers will note that our review jumped back and [11] M. Eddaoudi, H. L. Li, O. M. Yaghi, J. Am. Chem. Soc. 2000, 122, 1391.
forth between experiments aimed at CO2/H2, CO2/CH4, and [12] O. M. Yaghi, M. OKeeffe, N. W. Ockwig, H. K. Chae, M. Eddaoudi, J. Kim,
Nature 2003, 423, 705.
CO2/N2 mixtures. The clear economic incentives associated [13] J. L. C. Rowsell, O. M. Yaghi, Microporous Mesoporous Mater. 2004, 73, 3.
with sour gas processing have motivated much of the work on [14] A. R. Millward, O. M. Yaghi, J. Am. Chem. Soc. 2005, 127, 17998.
CO2/CH4 mixtures, but if the goal of a project is to consider [15] K. S. Park, Z. Ni, A. P. Cote, J. Y. Choi, R. D. Huang, F. J. Uribe-Romo,
flue gas processing, then it is vital to collect data with CO2/N2 H. K. Chae, M. OKeeffe, O. M. Yaghi, Proc. Natl. Acad. Sci. USA 2006,
103, 10186.
mixtures. Much can be inferred about CO2/N2 separations from [16] O. M. Yaghi, Q. W. Li, MRS Bull. 2009, 34, 682.
the CO2/CH4 data. CH4 typically adsorbs more strongly in nano- [17] L. Pan, K. M. Adams, H. E. Hernandez, X. T. Wang, C. Zheng, Y. Hattori,
pores than N2, so adsorption selectivities for CO2/N2 mixtures K. Kaneko, J. Am. Chem. Soc. 2003, 125, 3062.
will typically be higher than for CO2/CH4 mixtures. Neverthe- [18] S. Kitagawa, R. Kitaura, S.-i. Noro, Angew. Chem. 2004, 116, 2388;
Angew. Chem. Int. Ed. 2004, 43, 2334.
less, having data at conditions similar to those of a particular [19] S. Keskin, J. Liu, R. B. Rankin, J. K. Johnson, D. S. Sholl, Ind. Eng. Chem.
application will always be more valuable for making predic- Res. 2009, 48, 2355.
tions for that application than the kind of inference by analogy [20] T. Dren, Y. S. Bae, R. Q. Snurr, Chem. Soc. Rev. 2009, 38, 1237.
just outlined. [21] E. D. Akten, R. Siriwardane, D. S. Sholl, Energy Fuels 2003, 17, 977.
[22] R. T. Yang, Gas Separation by Adsorption Processes, Butterworths,
Throughout this article, we have focused on the question Boston, 1987.
posed in our articles title: can MOFs play a useful role in large [23] S. Keskin, D. S. Sholl, Langmuir 2009, 25, 11786.
scale CO2 related separations? No firm conclusion on this ques- [24] S.-E. Jee, D. S. Sholl, J. Am. Chem. Soc. 2009, 131, 7896.
tion is currently possible. We hope that the next few years will [25] S. Babel, T. A. Kurniawan, J. Hazard. Mater. 2003, 97, 219.
[26] R. L. Virta, C. H. Lindsay, Zeolites, US Geological Survey Minerals Year-
deliver information associated with the open issues we have book, 1999.
raised above that will enable the R&D community to resolve [27] E. Ruiz-Agudo, S. Galli, J. A. R. Navarro, Inorg. Chim. Acta 2007, 360, 84.
this question in a definitive way. [28] H. L. Guo, G. S. Zhu, I. J. Hewitt, S. L. Qiu, J. Am. Chem. Soc. 2009, 131,
We close by briefly suggesting a way in which MOFs may 1646.
[29] C. E. Willans, S. French, L. J. Barbour, J. A. Gertenbach, P. C. Junk, G. O.
enable breakthroughs in the broader area of porous adsorb- Lloyd, J. W. Steed, Dalton Trans. 2009, 6480.
ents for CO2 separations. It is possible that the best adsorbents [30] T. M. Reineke, M. Eddaoudi, M. OKeeffe, O. M. Yaghi, Angew. Chem.
that will emerge for these applications will have high surface 1999, 111, 2712; Angew. Chem. Int. Ed. 1999, 38, 2590.
areas but be intrinsically disordered on molecular length [31] L. Bastin, P. S. Barcia, E. J. Hurtado, J. A. C. Silva, A. E. Rodrigues, B.
Chen, J. Phys. Chem. C 2008, 112, 1575.
scales.[113] Materials of this kind are difficult to optimize, at least [32] B. Mu, F. Li, K. S. Walton, Chem. Commun. 2009, 2493.
in part because the molecular-scale mechanisms that control [33] Z. Y. Guo, J. B. Yu, G. H. Li, Z. J. Si, H. D. Guo, H. J. Zhang, CrystEng-
their behavior are hard to characterize. MOFs have potential to Comm 2009, 11, 2254.
play a useful role in this area because their highly ordered [34] Z. X. Zhao, Z. Li, Y. S. Lin, Ind. Eng. Chem. Res. 2009, 48, 10015.
[35] A. J. Fletcher, E. J. Cussen, T. J. Prior, M. J. Rosseinsky, C. J. Kepert, K. M.
structure makes them amenable to experimental and theoreti- Thomas, J. Am. Chem. Soc. 2001, 123, 10001.
cal understanding. Efforts that seek to use MOFs as prototypes [36] P. D. C. Dietzel, R. E. Johnsen, H. Fjellvag, S. Bordiga, E. Groppo, S.
to gain understanding of the important fundamental issues Chavan, R. Blom, Chem. Commun. 2008, 5125.
that exist for more complex adsorbents could in the future [37] B. Arstad, H. Fjellvag, K. O. Kongshaug, O. Swang, R. Blom, Adsorption
2008, 14, 755.
make significant contributions to this field.
[38] R. Vaidhyanathan, S. S. Iremonger, K. W. Dawson, G. K. H. Shimizu,
Chem. Commun. 2009, 5230.
[39] S. Couck, J. F. M. Denayer, G. V. Baron, T. Remy, J. Gascon, F. Kapteijn, J.
Acknowledgements Am. Chem. Soc. 2009, 131, 6326.
[40] C. Serre, S. Bourrelly, A. Vimont, N. A. Ramsahye, G. Maurin, P. L. Llewel-
lyn, M. Daturi, Y. Filinchuk, O. Leynaud, P. Barnes, G. Ferey, Adv. Mater.
This work was partially supported by the National Science Foun- 2007, 19, 2246.
dation. [41] K. L. Mulfort, O. K. Farha, C. D. Malliakas, M. G. Kanatzidis, J. T. Hupp,
Chem. Eur. J. 2010, 16, 276.
[42] P. D. C. Dietzel, V. Besikiotis, R. Blom, J. Mater. Chem. 2009, 19, 7362.
Keywords: adsorption carbon storage metal-organic [43] J. A. R. Navarro, E. Barea, J. M. Salas, N. Masciocchi, S. Galli, A. Sironi,
frameworks porous materials sustainable chemistry C. O. Ania, J. B. Parra, J. Mater. Chem. 2007, 17, 1939.
[44] Y. S. Bae, O. K. Farha, A. M. Spokoyny, C. A. Mirkin, J. T. Hupp, R. Q.
Snurr, Chem. Commun. 2008, 4135.
[1] E. Drioli, M. Romano, Ind. Eng. Chem. Res. 2001, 40, 1277. [45] Z. J. Liang, M. Marshall, A. L. Chaffee, Energy Fuels 2009, 23, 2785.
[2] T. Wall, Y. H. Liu, C. Spero, L. Elliott, S. Khare, R. Rathnam, F. Zeenathal, [46] L. Hamon, P. L. Llewellyn, T. Devic, A. Ghoufi, G. Clet, V. Guillerm, G. D.
B. Moghtaderi, B. Buhre, C. D. Sheng, R. Gupta, T. Yamada, K. Makino, Pirngruber, G. Maurin, C. Serre, G. Driver, W. van Beek, E. Jolimaitre, A.
J. L. Yu, Chem. Eng. Res. Des. 2009, 87, 1003. Vimont, M. Daturi, G. Ferey, J. Am. Chem. Soc. 2009, 131, 17490.

ChemSusChem 2010, 3, 879 891  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemsuschem.org 889
D. S. Sholl et al.

[47] B. Wang, A. P. Cote, H. Furukawa, M. OKeeffe, O. M. Yaghi, Nature [82] P. Kanoo, R. Matsuda, M. Higuchi, S. Kitagawa, T. K. Maji, Chem. Mater.
2008, 453, 207. 2009, 21, 5860.
[48] P. Chowdhury, C. Bikkina, D. Meister, F. Dreisbach, S. Gumma, Micropo- [83] A. Rossin, A. Ienco, F. Costantino, T. Montini, B. Di Credico, M. Caporali,
rous Mesoporous Mater. 2009, 117, 406. L. Gonsalvi, P. Fornasiero, M. Peruzzini, Cryst. Growth Des. 2008, 8,
[49] R. Banerjee, H. Furukawa, D. Britt, C. Knobler, M. OKeeffe, O. M. Yaghi, 3302.
J. Am. Chem. Soc. 2009, 131, 3875. [84] S. Galli, N. Masciocchi, G. Tagliabue, A. Sironi, J. A. R. Navarro, J. M.
[50] H. Furukawa, J. Kim, N. W. Ockwig, M. OKeeffe, O. M. Yaghi, J. Am. Salas, L. Mendez-Linan, M. Domingo, M. Perez-Mendoza, E. Barea,
Chem. Soc. 2008, 130, 11650. Chem. Eur. J. 2008, 14, 9890.
[51] Y. K. Park, S. B. Choi, H. Kim, K. Kim, B. H. Won, K. Choi, J. S. Choi, W. S. [85] R. Kitaura, K. Seki, G. Akiyama, S. Kitagawa, Angew. Chem. 2003, 115,
Ahn, N. Won, S. Kim, D. H. Jung, S. H. Choi, G. H. Kim, S. S. Cha, Y. H. 444; Angew. Chem. Int. Ed. 2003, 42, 428.
Jhon, J. K. Yang, J. Kim, Angew. Chem. 2007, 119, 8378; Angew. Chem. [86] M. Avila, L. Reguera, J. Rodriguez-Hernandez, J. Balmaseda, E. Reguera,
Int. Ed. 2007, 46, 8230. J. Solid State Chem. 2008, 181, 2899.
[52] J. Liang, G. K. H. Shimizu, Inorg. Chem. 2007, 46, 10449. [87] J. Roque, E. Reguera, J. Balmaseda, J. Rodriguez-Hernandez, L. Reg-
[53] M. Eddaoudi, H. L. Li, T. Reineke, M. Fehr, D. Kelley, T. L. Groy, O. M. uera, L. F. del Castillo, Microporous Mesoporous Mater. 2007, 103, 57.
Yaghi, Top. Catal. 1999, 9, 105. [88] M. P. Suh, Y. E. Cheon, E. Y. Lee, Chem. Eur. J. 2007, 13, 4208.
[54] A. C. Sudik, A. R. Millward, N. W. Ockwig, A. P. Cote, J. Kim, O. M. Yaghi, [89] Z. X. Chen, S. C. Xiang, D. Y. Zhao, B. L. Chen, Cryst. Growth Des. 2009,
J. Am. Chem. Soc. 2005, 127, 7110. 9, 5293.
[55] S. Bourrelly, P. L. Llewellyn, C. Serre, F. Millange, T. Loiseau, G. Ferey, J. [90] J.-R. Li, Y. Tao, Q. Yu, X.-H. Bu, H. Sakamoto, S. Kitagawa, Chem. Eur. J.
Am. Chem. Soc. 2005, 127, 13519. 2008, 14, 2771.
[56] H. Chun, J. Seo, Inorg. Chem. 2009, 48, 9980. [91] Y. E. Cheon, M. P. Suh, Chem. Commun. 2009, 2296.
[57] S. R. Caskey, A. G. Wong-Foy, A. J. Matzger, J. Am. Chem. Soc. 2008, [92] S. Noro, D. Tanaka, H. Sakamoto, S. Shimomura, S. Kitagawa, S. Takeda,
130, 10870. K. Uemura, H. Kita, T. Akutagawa, T. Nakamura, Chem. Mater. 2009, 21,
[58] P. D. Southon, L. Liu, E. A. Fellows, D. J. Price, G. J. Halder, K. W. Chap- 3346.
man, B. Moubaraki, K. S. Murray, J. F. Letard, C. J. Kepert, J. Am. Chem. [93] Y. S. Bae, K. L. Mulfort, H. Frost, P. Ryan, S. Punnathanam, L. J. Broad-
Soc. 2009, 131, 10998. belt, J. T. Hupp, R. Q. Snurr, Langmuir 2008, 24, 8592.
[59] A. O. Yazaydin, A. I. Benin, S. A. Faheem, P. Jakubczak, J. J. Low, R. R. [94] P. S. Barcia, L. Bastin, E. J. Hurtado, J. A. C. Silva, A. E. Rodrigues, B. L.
Willis, R. Q. Snurr, Chem. Mater. 2009, 21, 1425. Chen, Sep. Sci. Technol. 2008, 43, 3494.
[60] Y. S. Bae, O. K. Farha, J. T. Hupp, R. Q. Snurr, J .Mater. Chem. 2009, 19, [95] H. J. Park, M. P. Suh, Chem. Commun. 2010, 46, 610.
2131. [96] H. Furukawa, O. M. Yaghi, J. Am. Chem. Soc. 2009, 131, 8875.
[61] H. Li, M. Eddaoudi, T. L. Groy, O. M. Yaghi, J. Am. Chem. Soc. 1998, 120, [97] A. Demessence, D. M. DAlessandro, M. L. Foo, J. R. Long, J. Am. Chem.
8571. Soc. 2009, 131, 8784.
[62] P. K. Thallapally, J. Tian, M. R. Kishan, C. A. Fernandez, S. J. Dalgarno, [98] S. R. Miller, G. M. Pearce, P. A. Wright, F. Bonino, S. Chavan, S. Bordiga,
P. B. McGrail, J. E. Warren, J. L. Atwood, J. Am. Chem. Soc. 2008, 130, I. Margiolaki, N. Guillou, G. Feerey, S. Bourrelly, P. L. Llewellyn, J. Am.
16842. Chem. Soc. 2008, 130, 15967.
[63] J. W. Yoon, S. H. Jhung, Y. K. Hwang, S. M. Humphrey, P. T. Wood, J. S. [99] J. Y. Lee, J. M. Roberts, O. K. Farha, A. A. Sarjeant, K. A. Scheidt, J. T.
Chang, Adv. Mater. 2007, 19, 1830. Hupp, Inorg. Chem. 2009, 48, 9971.
[64] T. K. Maji, G. Mostafa, R. Matsuda, S. Kitagawa, J. Am. Chem. Soc. 2005, [100] S. Surbl, F. Millange, C. Serre, T. Duren, M. Latroche, S. Bourrelly, P. L.
127, 17152. Llewellyn, G. Ferey, J. Am. Chem. Soc. 2006, 128, 14 889.
[65] J. A. R. Navarro, E. Barea, J. M. Salas, N. Masciocchi, S. Galli, A. Sironi, [101] J. Y. An, R. P. Fiorella, S. J. Geib, N. L. Rosi, J. Am. Chem. Soc. 2009, 131,
C. O. Ania, J. B. Parra, Inorg. Chem. 2006, 45, 2397. 8401.
[66] D. Farrusseng, C. Daniel, C. Gaudillere, U. Ravon, Y. Schuurman, C. Mir- [102] K. S. Walton, A. R. Millward, D. Dubbeldam, H. Frost, J. J. Low, O. M.
odatos, D. Dubbeldam, H. Frost, R. Q. Snurr, Langmuir 2009, 25, 7383. Yaghi, R. Q. Snurr, J. Am. Chem. Soc. 2008, 130, 406.
[67] J. An, S. J. Geib, N. L. Rosi, J. Am. Chem. Soc. 2010, 132, 38. [103] A. A. Lemus-Santana, J. Rodriguez-Hernandez, L. F. del Castillo, M. Bas-
[68] P. L. Llewellyn, S. Bourrelly, C. Serre, A. Vimont, M. Daturi, L. Hamon, G. terrechea, E. Reguera, J. Solid State Chem. 2009, 182, 757.
De Weireld, J. S. Chang, D. Y. Hong, Y. K. Hwang, S. H. Jhung, G. Ferey, [104] E. Neofotistou, C. D. Malliakas, P. N. Trikalitis, Chem. Eur. J. 2009, 15,
Langmuir 2008, 24, 7245. 4523.
[69] D. Britt, H. Furukawa, B. Wang, T. G. Glover, O. M. Yaghi, Proc. Natl. [105] S. Neogi, J. A. R. Navarro, P. K. Bharadwaj, Cryst. Growth Des. 2008, 8,
Acad. Sci. USA 2009, 106, 20637. 1554.
[70] H. S. Choi, M. P. Suh, Angew. Chem. 2009, 121, 6997; Angew. Chem. Int. [106] H. Hayashi, A. P. Cote, H. Furukawa, M. OKeeffe, O. M. Yaghi, Nat.
Ed. 2009, 48, 6865. Mater. 2007, 6, 501.
[71] R. Banerjee, A. Phan, B. Wang, C. Knobler, H. Furukawa, M. OKeeffe, [107] T. Wu, J. Zhang, C. Zhou, L. Wang, X. H. Bu, P. Y. Feng, J. Am. Chem.
O. M. Yaghi, Science 2008, 319, 939. Soc. 2009, 131, 6111.
[72] P. L. Llewellyn, S. Bourrelly, C. Serre, Y. Filinchuk, G. Ferey, Angew. Chem. [108] A. O. Yazaydin, R. Q. Snurr, T. H. Park, K. Koh, J. Liu, M. D. LeVan, A. I.
2006, 118, 7915; Angew. Chem. Int. Ed. 2006, 45, 7751. Benin, P. Jakubczak, M. Lanuza, D. B. Galloway, J. J. Low, R. R. Willis, J.
[73] D. Li, K. Kaneko, Chem. Phys. Lett. 2001, 335, 50. Am. Chem. Soc. 2009, 131, 18 198.
[74] J. T. Culp, M. R. Smith, E. Bittner, B. Bockrath, J. Am. Chem. Soc. 2008, [109] T. Loiseau, L. Lecroq, C. Volkringer, J. Marrot, G. Ferey, M. Haouas, F.
130, 12427. Taulelle, S. Bourrelly, P. L. Llewellyn, M. Latroche, J. Am. Chem. Soc.
[75] J. Seo, H. Chun, Eur. J. Inorg. Chem. 2009, 4946. 2006, 128, 10223.
[76] Q. Min Wang, D. M. Shen, M. Bulow, M. L. Lau, S. G. Deng, F. R. Fitch, [110] N. A. Ramsahye, G. Maurin, S. Bourrelly, P. L. Llewellyn, T. Loiseau, C.
N. O. Lemcoff, J. Semanscin, Microporous Mesoporous Mater. 2002, 55, Serre, G. Ferey, Chem. Commun. 2007, 3261.
217. [111] C. Volkringer, T. Loiseau, M. Haouas, F. Taulelle, D. Popov, M. Burgham-
[77] S. Q. Ma, X. S. Wang, E. S. Manis, C. D. Collier, H. C. Zhou, Inorg. Chem. mer, C. Riekel, C. Zlotea, F. Cuevas, M. Latroche, D. Phanon, C. Knofelv,
2007, 46, 3432. P. L. Llewellyn, G. Ferey, Chem. Mater. 2009, 21, 5783.
[78] D. G. Samsonenko, H. Kim, Y. Y. Sun, G. H. Kim, H. S. Lee, K. Kim, Chem. [112] B. Metz, O. Davidson, H. de Coninck, M. Loos, L. Meyer, IPCC Special
Asian J. 2007, 2, 484. Report on Carbon Dioxide Capture and Storage, Cambridge University
[79] X. F. Wang, Y. B. Zhang, H. Huang, J. P. Zhang, X. M. Chen, Cryst. Growth Press, Cambridge, United Kingdom, 2005.
Des. 2008, 8, 4559. [113] S. Choi, J. H. Drese, C. W. Jones, ChemSusChem 2009, 2, 796.
[80] H. J. Lee, W. Cho, S. Jung, M. Oh, Adv. Mater. 2009, 21, 674. [114] S. K. Bhatia, A. L. Myers, Langmuir 2006, 22, 1688.
[81] H. Deng, C. J. Doonan, H. Furukawa, R. B. Ferreira, J. Towne, C. B. Kno- [115] C. C. Zheng, D. H. Liu, Q. Y. Yang, C. L. Zhong, J. G. Mi, Ind. Eng. Chem.
bler, B. Wang, O. M. Yaghi, Science 2010, 327, 846. Res. 2009, 48, 10479.

890 www.chemsuschem.org  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemSusChem 2010, 3, 879 891
Can MOF Materials Play a Useful Role in Large-Scale CO2 Separations?

[116] S. S. Y. Chui, S. M. F. Lo, J. P. H. Charmant, A. G. Orpen, I. D. Williams, Sci- [141] J. Choi, H. Jeong, M. A. Snyder, J. A. Stoeger, R. I. Masel, M. Tsapatsis,
ence 1999, 283, 1148. Science 2009, 325, 590.
[117] S. Sircar, T. C. Golden, M. B. Rao, Carbon 1996, 34, 1. [142] T. Tomita, K. Nakayama, H. Sakai, Microporous Mesoporous Mater. 2004,
[118] S. Cavenati, C. A. Grande, A. E. Rodrigues, J. Chem. Eng. Data 2004, 49, 68, 71.
1095. [143] S. Li, J. G. Martinek, J. L. Falconer, R. D. Noble, Ind. Eng. Chem. Res.
[119] P. Y. Li, F. H. Tezel, Microporous Mesoporous Mater. 2007, 98, 94. 2005, 44, 3220.
[120] A. L. Myers, in Fundamental of Adsorption (Ed.: A. Liapis), Engineering [144] Y. Y. Liu, Z. F. Ng, E. A. Khan, H. K. Jeong, C. B. Ching, Z. P. Lai, Micropo-
Foundation, New York, 1986. rous Mesoporous Mater. 2009, 118, 296.
[121] A. L. Myers, J. M. Prausnitz, AIChE J. 1965, 11, 121. [145] H. Bux, F. Y. Liang, Y. S. Li, J. Cravillon, M. Wiebcke, J. Caro, J. Am. Chem.
[122] H. Chen, D. S. Sholl, Langmuir 2007, 23, 6431. Soc. 2009, 131, 16000.
[123] M. Murthi, R. Q. Snurr, Langmuir 2004, 20, 2489. [146] R. Ranjan, M. Tsapatsis, Chem. Mater. 2009, 21, 4920.
[124] Q. Y. Yang, C. L. Zhong, J. Phys. Chem. B 2006, 110, 17776. [147] A. I. Skoulidas, D. S. Sholl, J. Phys. Chem. B 2005, 109, 15760.
[125] R. Babarao, Z. Q. Hu, J. W. Jiang, S. Chempath, S. I. Sandler, Langmuir [148] A. I. Skoulidas, J. Am. Chem. Soc. 2004, 126, 1356.
2007, 23, 659. [149] S. Keskin, J. C. Liu, J. K. Johnson, D. S. Sholl, Langmuir 2008, 24, 8254.
[126] Q. Y. Yang, C. Y. Xue, C. L. Zhong, J. F. Chen, AIChE J. 2007, 53, 2832. [150] L. Sarkisov, T. Dren, R. Q. Snurr, Mol. Phys. 2004, 102, 211.
[127] V. Finsy, L. Ma, L. Alaerts, D. E. De Vos, G. V. Baron, J. F. M. Denayer, Mi- [151] F. Salles, H. Jobic, A. Ghoufi, P. L. Llewellyn, C. Serre, S. Bourrelly, G.
croporous Mesoporous Mater. 2009, 120, 221. Ferey, G. Maurin, Angew. Chem. 2009, 121, 8485; Angew. Chem. Int. Ed.
[128] N. A. Ramsahye, G. Maurin, S. Bourrelly, P. L. Llewellyn, T. Devic, C. 2009, 48, 8335.
Serre, T. Loiseau, G. Ferey, Adsorption 2007, 13, 461. [152] F. Salles, H. Jobic, T. Devic, P. L. Llewellyn, C. Serre, G. Ferey, G. Maurin,
[129] L. M. Robeson, J. Membr. Sci. 1991, 62, 165. ACS Nano 2010, 4, 143.
[130] C. M. Zimmerman, A. Singh, W. J. Koros, J. Membr. Sci. 1997, 137, 145. [153] S. Keskin, D. S. Sholl, Ind. Eng. Chem. Res. 2009, 48, 914.
[131] L. F. Greenlee, D. F. Lawler, B. D. Freeman, B. Marrot, P. Moulin, Water [154] J. A. Greathouse, M. D. Allendorf, J. Am. Chem. Soc. 2006, 128, 10678.
Res. 2009, 43, 2317. [155] Y. Li, R. T. Yang, Langmuir 2007, 23, 12937.
[132] T. C. Merkel, H. Lin, X. Wei, R. Baker, J. Membr. Sci. 2010, 359, 126. [156] A. Kondo, T. Daimaru, H. Noguchi, T. Ohba, K. Kaneko, H. Kanob, J. Col-
[133] M. T. Ho, G. W. Allinson, D. E. Wiley, Ind. Eng. Chem. Res. 2008, 47, 1562. loid Interface Sci. 2007, 314, 422.
[134] A. Car, C. Stropnik, K. V. Peinemann, Desalination 2006, 200, 424. [157] J. J. Low, A. I. Benin, P. Jakubczak, J. F. Abrahamian, S. A. Faheem, R. R.
[135] Y. F. Zhang, I. H. Musseman, J. P. Ferraris, K. J. Balkus, J. Membr. Sci. Willis, J. Am. Chem. Soc. 2009, 131, 15834.
2008, 313, 170. [158] J. Lu, Y. J. Choi, Z. Z. Fang, H. Y. Sohn, E. Ronnebro, J. Am. Chem. Soc.
[136] E. V. Perez, K. J. Balkus, J. P. Ferraris, I. H. Musselman, J. Membr. Sci. 2009, 131, 15843.
2009, 328, 165. [159] A. C. McKinlay, B. Xiao, D. S. Wragg, P. S. Wheatley, I. L. Megson, R. E.
[137] R. Adams, C. Carson, J. Ward, R. Tannenbaum, W. Koros, Microporous Morris, J. Am. Chem. Soc. 2008, 130, 10440.
Mesoporous Mater. 2010, 131, 13.
[138] K. Diaz, L. Garrido, M. Lopez-Gonzalez, L. F. del Castillo, E. Riande, Mac-
romolecules 2010, 43, 316.
[139] S. Keskin, D. S. Sholl, J. Phys. Chem. C 2007, 111, 14055.
[140] T. Watanabe, S. Keskin, S. Nair, D. S. Sholl, Phys. Chem. Chem. Phys. Received: April 15, 2010
2009, 11, 11389. Published online on July 9, 2010

ChemSusChem 2010, 3, 879 891  2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemsuschem.org 891

View publication stats

You might also like