ArticlePDF Available

Separation of polar compounds using a flexible metal–organic framework

Authors:

Abstract and Figures

A flexible metal-organic framework constructed from a flexible linker is shown to possess the capability of separating mixtures of polar compounds (propanol isomers) by exploiting the differences in the saturation capacities of the constituents. Transient breakthrough simulations show that these sorption-based separations are in favor of the component with higher saturation capacity.
Content may be subject to copyright.
As featured in:
See R. K. Motkuri, P. K. Thallapally,
R. Krishna et al., Chem. Commun.,
2015, 51, 8421.
Showcasing collaborative research from the Laboratory of
Dr Radha Kishan Motkuri et al. at the Pacifi c Northwest National
Laboratory, Richland, USA and Prof. Rajamani Krishna at the
University of Amsterdam, The Netherlands
Separation of polar compounds using a fl exible metal–organic
framework
Inspired by the need of energy reduction in the area of separation
and purifi cation of organic liquid mixtures, we developed a
exible metal–organic framework that was shown to possess
the separation capability of polar compounds such as alcohols
and ketones, specifi cally propanol isomers. The experimental and
simulations studies revealed that sorption-based separations are
in favor for the component with higher saturation capacity.
Registered charity number: 207890
www.rsc.org/chemcomm
This journal is ©The Royal Society of Chemistry 201 5 Chem. Commun., 2015, 51, 8421--8424 | 8421
Cite this: Chem. Commun., 2015,
51,8421
Separation of polar compounds using a flexible
metal–organic framework
Radha Kishan Motkuri,*
a
Praveen K. Thallapally,*
b
Harsha V. R. Annapureddy,
b
Liem X. Dang,
b
Rajamani Krishna,*
c
Satish K. Nune,
a
Carlos A. Fernandez,
a
Jian Liu
a
and B. Peter McGrail
a
A flexible metal–organic framework constructed from a flexible linker is
shown to possess the capability of separating mixtures of polar com-
pounds (propanol isomers) by exploiting the differences in the saturation
capacities of the constituents. Transient breakthrough simulations show
that these sorption-based separations are in favor of the component
with higher saturation capacity.
Separation and purification of organic liquid isomers are scientifi-
cally important industrial technologies and have received consider-
able attention worldwide.
1
Distillation is clearly the dominating
separation process, accounting for more applications than all the
other chemical-separation processes combined. In fact, distillation
columns consume more than 50% of the total energy used in the
chemical industry worldwide. Even more challenging is separation
of an azeotrope mixture that forms when certain compositions of
liquid isomers are present by weight. Specifically, the separation
of water, alcohols, and ketones is often made difficult because of
azeotrope formation. Separating these mixtures using fractional
distillation or using polymeric membranes is energy-intensive and
is highly complex.
2
Alternatively, these processes sometimes require
the addition of separating agents, called entrainers, that alter
the vapor/liquid equilibrium in a favorable manner to achieve the
desired separation, but the recovery of such entrainers later in the
process not only requires an additional distillation step but also
incurs an increased overall energy penalty. The largest opportunities
for energy reduction in this area are offered by replacing distillation
or membrane-based separations by low-cost adsorption-based
systems. The success of such replacement strategies is crucially
dependent on the development of suitable adsorbents, but there
is very limited information available on adsorption-based separa-
tion of azeotropes using porous media.
Recent developments in porous metal–organic frameworks
(MOFs) have gained much attention because of the outstanding
properties and ability to fine tune the pore apertures and high
stability towards the desired application.
3,4
Such remarkable
properties of MOFs make them an interesting class of materials
for adsorption,
5
and separation applications.
6
Specifically, studies
of the gas separation are extensively reported in the literature;
however, very limited information is available on the separation of
polar molecules, including azeotropic mixtures. For example,
Denayer et al. used highly stable zeolitic imidazole frameworks
(ZIF-8, ZIF-68) to separate butanol from aqueous mixtures in the
presence of organic contaminants like ethanol.
7
Jie Zhang et al.
reported the separation of alcohol and water mixtures using a
charge-polarized MOF that shows selectivity towards polar mole-
cules under an electric field gradient.
8
Similarly, Kitagawa et al.
synthesized a copper-based coordination polymer that selectively
adsorbs methanol and water from bioethanol.
9
Mostofthese
studies focused on purifying bioethanol, but there are few reports
that focus on the separation of mixtures of alcohols and other polar
molecules such as chloroform and acetone.
6g,10
Our experimental
adsorption studies coupled with transient breakthrough simulations
confirm the separation of propanol isomers and various azeotropes.
To our knowledge, this is the first report on the separation of
mixtures of propanol isomers and other binary mixtures containing
alcohols, chloroform, and acetone using flexible MOFs.
TetZB, the flexible porous framework used in this communica-
tion, was synthesized using a flexible tetrahedral organic linker,
tetrakis[4-(carboxyphenyl)-oxamethyl]methane 1(Scheme S1, ESI),
and then was used effectively for the sorption and separation of
polar solvents. The synthesis method and associated sorption
properties of TetZB were reported by us previously.
11
For this study,
we chose adsorption experiments of polar solvents such as C1–C3
alcohols, water, acetone, chloroform, and benzene, respectively.
a
Energy and Environment Directorate, Pacific Northwest National Laboratory (PNNL),
Richland, WA 99352, USA. E-mail: radhakishan.motkuri@pnnl.gov
b
Fundamental and Computational Sciences Directorate, PNNL, Richland, WA 99352,
USA. E-mail: praveen.thallapally@pnnl.gov
c
Van ’t Hoff Institute for Molecular Sciences, University of Amsterdam,
The Netherlands. E-mail: r.krishna@contact.uva.nl
Electronic supplementary information (ESI) available: (a) Material synthesis,
characterization; (b) pure component isotherms and dual-Langmuir–Freundlich
models; (c) adsorption energy calculations; (d) GCMC simulation studies; (e) IAST
calculations; (f) transient breakthrough simulation methodology; and (g) video
animations for transient breakthroughs of several binary mixture separations. See
DOI: 10.1039/c5cc00113g
Received 6th January 2015,
Accepted 19th February 2015
DOI: 10.1039/c5cc00113g
www.rsc.org/chemcomm
ChemComm
COMMUNICATION
Published on 27 February 2015. Downloaded by PNNL Technical Library on 10/10/2015 00:32:12.
View Article Online
View Journal
| View Issue
8422 |Chem. Commun., 2015, 51, 8421--8424 This journal is ©The Royal Society of Chemistry 201 5
Experimentally measured vapor sorption capacities were obtained
using an Intelligent Gravimetric Analyzer (IGA) from Hiden Instru-
ments. The TetZB sample was activated at 473 K under dynamic
vacuum before sorption studies. To evaluate the separation efficiency
of TetZB, we initially considered 1-propanol/2-propanol isomers for
the sorption studies. After sample activation, the MOF sample was
exposed to 1-propanol vapors and the sorption behavior was plotted
against pressure. The adsorption curve shows a sudden increase in
uptake at a relative pressure (P/P
0
)of0.15reachingthefirstsatura-
tion capacity of 7.8 wt%. At a relative pressure (P/P
0
) of 0.27, TetZB
shows another step adsorption reaching the second saturation
capacity of 26 wt% or 4.55 mmol g
1
(Fig. 1). Such two-step
adsorption was observed in TetZB with other gases/vapors and was
shown to expand and contract the framework upon guest removal
and re-adsorption of the same or different guest molecule. The
flexibility arises from the twisting of benzoate moieties around the
central quaternary carbon atom through ethereal links of the tetra-
hedral building block, which result from diverse ligand geometries
such as tetrahedral, irregular, or near-flattened. Such building block
flexibility has been observed both by us and other researchers. The
desorption curve does not follow the adsorption, rather, it shows a
sudden decrease in the sorption capacity at a P/P
0
ratio of 0.03
(Fig. S1, ESI). Similarly, when a freshly activated MOF sample was
exposed to 2-propanol vapors, the first uptake isotherm reached its
first plateau at P/P
0
= 0.3; which was followed by a step adsorption
with approximately 2.5 times higher capacity (25 wt%, 4.1 mmol g
1
at P/P
0
= 0.8), and then the saturation point was reached. Another
significant difference between these two sorption isotherms is the
rates at which they sorb onto the TetZB framework. Sorption profiles
indicate that both propanol isomers can enter the pores of TetZB,
but 1-propanol with its kinetic diameter of 4.7 Å has slightly higher
uptake when compared to 2-propanol with the same kinetic dia-
meter. This may be attributed to the flexibility of 1-propanol, which
is a linear chain that can enter the pore more easily than a branched
isomer. The density functional theory (DFT) estimated dipole
moment value of 2-propanol is slightly higher (1.56D) than that of
1-propanol (1.49D), which shows that the 2-propanol molecule is
more likely to be polarized by the TetZB pore structure, thus having a
sharper uptake at relatively low pressure compared to 1-propanol.
12
To gain further insights into sorption behavior, we performed
grand canonical Monte Carlo (GCMC) simulations using the MuSic
program where the simulation box consisted of one unit cell of
MOF and the periodic boundary conditions were used in all three
dimensions.
13
Because the host framework considered has a rigid
structure, the breathing phenomenon was not observed, but the
overall solvent uptake matched the experimental results at 25 1C
(Fig. S6, ESI). In agreement with experimental results, the simula-
tions of the 2-propanol sorption curve appear to be steeper at lower
loadings when compared to 1-propanol. To understand this
behavior, we computed the interaction energies between the
TetZB framework and propanol isomers as a function of loading.
The simulated results clearly showed more negative interaction
energies for 2-propanol when compared to 1-propanol at lower
loadings (Fig. S6, ESI), but the overall uptake is slightly higher
for 1-propanol. These intriguing experimental and simulation
results suggest that vapor sorption experiments of 1-propanol
and 2-propanol using the flexible TetZB have potential for
separating propanol isomers, which motivated us to undertake
further IAST breakthrough simulations (Fig. 2).
We then focused our attention on lower chain alcohols such
as methanol and ethanol. The sorption isotherm of methanol
shows a sudden increase at a P/P
0
ratio of 0.18 and then reaches
Fig. 1 Adsorption isotherms of alcohol adsorbents and water in TetZB at
298 K.
Fig. 2 Adsorption and desorption isotherms of 1-/2-propanol in TetZB
(top) and the corresponding transient breakthrough simulation character-
istics of an adsorber packed with TetZB for separation of 50/50 mixtures of
1-propanol from 2-propanol (bottom).
Communication ChemComm
Published on 27 February 2015. Downloaded by PNNL Technical Library on 10/10/2015 00:32:12.
View Article Online
This journal is ©The Royal Society of Chemistry 201 5 Chem. Commun., 2015, 51, 8421--8424 | 8423
27 wt% at a P/P
0
ratio of 0.8. For ethanol after the first uptake
at low relative pressure, the isotherm reaches its first plateau
(7.8 wt% at a P/P
0
ratio of 0.12) followed by a step adsorption
with approximately four times higher capacity of ethanol
(25.7 wt%, 6 mmol g
1
at P/P
0
= 0.25). Because of the hydro-
phobic nature of the TetZB framework, water sorption studies
show a low uptake until the P/P
0
ratio reaches 0.7, and
the isotherm does not reach saturation even at a P/P
0
ratio of
0.95. The saturated loadings of alcohols decrease as the size
increases from methanol to propanol because adsorption near
saturation is mainly directed by the entropic (size) effect as
fewer propanol molecules can be adsorbed compared to methanol
(Fig. S2, ESI). Furthermore, interestingly when TetZB is exposed to
acetone vapors, the first plateau is reached at a very low relative
pressure (P/P
0
= 0.05). Chloroform and benzene show a type-I
isotherm that exhibits significant uptakes at low vapor pressures.
The distinct behaviors of the solvent molecules with the host
framework definitely reveal potential for application in separation
technologies that should be studied.
To investigate the separation potential of TetZB, the experi-
mentally measured loadings of 1- and 2-propanols, methanol,
ethanol, acetone, benzene, chloroform, and water were fitted
with the dual-site Langmuir–Freundlich model, and the fits are
excellent over the entire range of pressures. The details of
simulation methodology and the breakthrough simulations
using IAST calculations are outlined in the ESI.The transient
breakthrough simulations suggest that TetZB has the potential to
separate mixtures of alcohols by differentiating on the basis of chain
length and conformation as can be observed for 1-propanol/
2-propanol mixtures (Fig. S7–S9, ESI). The separation of
1-propanol from 2-propanol is governed by molecular packing
effects that favor the adsorption of the linear alcohol when
operating under conditions corresponding to pore saturation.
The better packing efficiency of 1-propanol is reflected in its
higher saturation capacity compared to 2-propanol. It is impor-
tant to note that this separation is not dictated by differences in
binding energies that are higher for 2-propanol (Fig. S6b, ESI).
For other mixtures of 1-alcohols, in the Henry regime, at
pressures below 1 kPa, selectivity favors alcohols with longer
chain lengths; however, at pressures above 10 kPa, selectivity
favors alcohols with shorter chain lengths. This is because of
the higher saturation capacity of the shorter chain alcohols.
The IAST calculations also imply that sharp separations of
alcohol mixtures are possible using TetZB provided the operating
pressures are greater than 10 kPa. This is confirmed in the
transient breakthrough simulations presented for 50/50 mixtures
of methanol/ethanol, ethanol/1-propanol, and ethanol/2-propanol
at a total pressure of 100 kPa (Fig. 3, Fig. S9–S12, ESI). It is
interesting to compare the separations of TetZB with those
obtained with ZIF-8 and CHA zeolites. The shorter chain alcohol
is eluted later than the longer chain alcohol, which is in agreement
with the corresponding results for other microporous materials
such as SAPO-34, and ZIF-8 reported previously.
7b,14
Comparisons
of ethanol/1-propanol adsorption selectivity, and uptake capacity of
ethanol for equimolar ethanol/1-propanol mixtures in TetZB, ZIF-8,
and CHA zeolites are shown in Fig. S10 (ESI). We note that TetZB
has both higher selectivity and higher uptake capacity, making it
more suitable for separation of mixtures of 1-alcohols (Fig. S11–S16,
ESI). Fig. 3f shows the separations of water–ethanol mixtures of
azeotropic composition using TetZB. The separation is selective to
water that has the higher saturation capacity; similar water-selective
separations, achieved as a result of molecular packing effects, have
been reported for CuBTC.
15
The methodology adopted for the
breakthrough simulations is provided in the ESI.Also available
in ESIare seven video animations of the breakthroughs.
The experimental and modeling sorption analysis shows
that the hydrophobic –CH
2
and aryl groups of tectonic acid
and phenyl groups of the 4,40-bipyridine molecules are exposed
inside the pore, thereby creating a hydrophobic environment.
11
To illustrate such an environment, we painted all the hydrophobic
groups in green and the hydrophilic groups in red where it is clearly
evident that hydrophobic groups dominate the surface of the pore
(Scheme S1, Fig. S6, ESI). The metal atoms and the carboxylate
groups that are more hydrophilic are buried deep inside and are
not easily accessible to the guest molecules. Polar alcohol, such as
methanol and ethanol, molecules consisting of hydrophobic and
Fig. 3 Transient breakthrough simulation characteristics of an adsorber
packed with TetZB for separation of ethanol from various solvent mixtures
of (a) methanol, (b) 1-propanol, (c) 2-propanol, (d) chloroform, (e) benzene,
(f) water at 298 K. The total pressure is 100 kPa.
ChemComm Communication
Published on 27 February 2015. Downloaded by PNNL Technical Library on 10/10/2015 00:32:12.
View Article Online
8424 |Chem. Commun., 2015, 51, 8421--8424 This journal is ©The Royal Society of Chemistry 2 015
hydrophilic groups interact favorably with pore components,
leading to higher uptake rates with high adsorption energies.
To our knowledge, this is the first report of the separation of
propanol isomers, mixtures of 1-alcohols with acetone, and
chloroform ketones using MOFs.
In conclusion, we reported that hydrophobic TetZB, a flexible
metal organic framework generated from a flexible tetrahedral
building block, shows remarkable affinity and separation capability
of alcohols and ketones, specifically separation of propanol iso-
mers. If the operating conditions are chosen such that pore
saturation is achieved, separation using TetZB strongly favors the
component with the higher saturation capacity. For mixtures of
alcohols, the separation is selective for the alcohols with the shorter
chain length. For separation of water–alcohol mixtures, the separa-
tion favors water. Of particular interest is the separation of azeo-
tropic water–ethanol mixtures; see Fig. 3(f).
This work was performed at the Pacific Northwest National
Laboratory (PNNL) and was supported by the U.S. Department
of Energy (DOE). L.X.D. acknowledges funding from the U.S.
Department of Energy, Office of Science, Office of Basic Energy
Sciences, Division of Chemical Sciences, Geosciences, and
Biosciences. PNNL is operated by Battelle for the U.S. Depart-
ment of Energy under Contract DE-AC05-76RL01830.
Notes and references
1 Industrial Technologies Program; US-DOE EERE, 2005, 14.
2 F. P. Mccandless and W. B. Downs, J. Membr. Sci., 1987, 30, 111–116.
3(a) O. K. Farha, C. E. Wilmer, I. Eryazici, B. G. Hauser, P. A. Parilla,
K. O’Neill, A. A. Sarjeant, S. T. Nguyen, R. Q. Snurr and J. T. Hupp,
J. Am. Chem. Soc., 2012, 134, 9860–9863; (b) M. Eddaoudi, J. Kim,
N. Rosi, D. Vodak, J. Wachter, M. O’Keeffe and O. M. Yaghi, Science,
2002, 295, 469–472; (c) H. X. Deng, S. Grunder, K. E. Cordova,
C. Valente, H. Furukawa, M. Hmadeh, F. Gandara, A. C. Whalley,
Z.Liu,S.Asahina,H.Kazumori,M.OKeee,O.Terasaki,J.F.Stoddart
andO.M.Yaghi,Science, 2012, 336, 1018–1023; (d) G. Ferey, Chem. Soc.
Rev., 2008, 37,191214;(e)N.StockandS.Biswas,Chem. Rev., 2012, 112,
933–969; ( f) D. Banerjee, A. J. Cains, J. Liu, R. K. Motkuri, S. K. Nune,
C. A. Fernandez, R. Krishna, D. M. Strachan and P. K. Thallapally, Acc.
Chem. Res., 2015, 48, 211–219.
4(a) J. R. Li, J. Sculley and H. C. Zhou, Chem. Rev., 2012, 112, 869–932;
(b) H. Furukawa, K. E. Cordova, M. O’Keeffe and O. M. Yaghi,
Science, 2013, 341, 1230444; (c) T. Li, M. T. Kozlowski, E. A. Doud,
M. N. Blakely and N. L. Rosi, J. Am. Chem. Soc., 2013, 135, 11688–11691;
(d) R. Haldar, S. K. Reddy, V. M. Suresh, S. Mohapatra,
S. Balasubramanian and T. K. Maji, Chem. – Eur. J., 2014, 20,
4347–4356; (e) R. K. Motkuri, J. Liu, C. A. Fernandez, S. K. Nune,
P. Thallapally and B. P. Mcgrail, Industrial Catalysis and Separations,
Apple Academic Press, 2014, pp. 61–103.
5 H. T. Kwon and H. K. Jeong, J. Am. Chem. Soc., 2013, 135, 10763–10768.
6(a) R. K. Motkuri, H. V. R. Annapureddy, M. Vijaykumar,
H. T. Schaef, P. F. Martin, B. P. McGrail, L. X. Dang, R. Krishna
and P. K. Thallapally, Nat. Commun., 2014, 5, 4368; (b) Z. Y. Gu and
X. P. Yan, Angew. Chem., Int. Ed., 2010, 49, 1477–1480; (c) Z. R. Herm,
B. M. Wiers, J. A. Mason, J. M. van Baten, M. R. Hudson, P. Zajdel,
C. M. Brown, N. Masciocchi, R. Krishna and J. R. Long, Science, 2013,
340, 960–964; (d) M. Latroche, S. Surble, C. Serre, C. Mellot-
Draznieks, P. L. Llewellyn, J. H. Lee, J. S. Chang, S. H. Jhung and
G. Ferey, Angew. Chem., Int. Ed., 2006, 45, 8227–8231; (e) Z. Y. Gu,
C. X. Yang, N. Chang and X. P. Yan, Acc. Chem. Res., 2012, 45,
734–745; ( f) L. C. Lin, J. Kim, X. Q. Kong, E. Scott, T. M. McDonald,
J. R. Long, J. A. Reimer and B. Smit, Angew. Chem., Int. Ed., 2013, 52,
4410–4413; (g) Z. J. Lin, R. Q. Zou, J. Liang, W. Xia, D. G. Xia,
Y. X. Wang, J. H. Lin, T. L. Hu, Q. Chen, X. D. Wang, Y. S. Zhao and
A. K. Burrell, J. Mater. Chem., 2012, 22, 7813–7818; (h) P. Pachfule,
Y. F. Chen, J. W. Jiang and R. Banerjee, Chem. – Eur. J., 2012, 18,
688–694; (i) N. Chang, Z. Y. Gu and X. P. Yan, J. Am. Chem. Soc., 2010,
132, 13645–13647; ( j) R. Krishna, Phys. Chem. Chem. Phys., 2015, 17,
39–59.
7(a) S. Van der Perre, T. Van Assche, B. Bozbiyik, J. Lannoeye, D. E. De
Vos, G. V. Baron and J. F. M. Denayer, Langmuir, 2014, 30,
8416–8424; (b) J. C. Saint Remi, T. Remy, V. Van Hunskerken,
S. van de Perre, T. Duerinck, M. Maes, D. De Vos, E. Gobechiya,
C. E. A. Kirschhock, G. V. Baron and J. F. M. Denayer, ChemSusChem,
2011, 4, 1074–1077.
8 J. K. Sun, M. Ji, C. Chen, W. G. Wang, P. Wang, R. P. Chen and
J. Zhang, Chem. Commun., 2013, 49, 1624–1626.
9 A. Shigematsu, T. Yamada and H. Kitagawa, J. Am. Chem. Soc., 2012,
134, 13145–13147.
10 (a) X. F. Zheng, L. Zhou, Y. M. Huang, C. G. Wang, J. G. Duan,
L. L. Wen, Z. F. Tian and D. F. Li, J. Mater. Chem. A, 2014, 2,
12413–12422; (b) K. Zhang, L. L. Zhang and J. W. Jiang, J. Phys. Chem.
C,2013,117, 25628–25635; (c) T. Borjigin, F. X. Sun, J. L. Zhang,
K. Cai, H. Ren and G. S. Zhu, Chem. Commun.,2012,48, 7613–7615.
11 (a)P.K.Thallapally,J.Tian,M.R.Kishan,C.A.Fernandez,
S.J.Dalgarno,P.B.McGrail,J.E.WarrenandJ.L.Atwood,J. Am. Chem.
Soc., 2008, 130, 16842–16843; (b)R.K.Motkuri,P.K.Thallapally,
S.K.Nune,C.A.Fernandez,B.P.McGrailandJ.L.Atwood,Chem.
Commun., 2011, 47, 7077–7079.
12 M. Sadakiyo, T. Yamada and H. Kitagawa, J. Am. Chem. Soc., 2011,
133, 11050–11053.
13 A. Gupta, S. Chempath, M. J. Sanborn, L. A. Clark and R. Q. Snurr,
Mol. Simul., 2003, 29, 29–46.
14 (a)R.KrishnaandJ.M.vanBaten,Sep. Purif. Technol., 2011, 76, 325–330;
(b)T.Remy,J.C.SaintRemi,R.Singh,P.A.Webley,G.V.Baronand
J. F. M. Denayer, J. Phys. Chem. C, 2011, 115, 8117–8125; (c)R.Krishna,
Microporous Mesoporous Mater., 2014, 185, 30–50.
15 J. J. Gutierrez-Sevillano, S. Calero and R. Krishna, J. Phys. Chem. C,
2015, 119, 3658–3666.
Communication ChemComm
Published on 27 February 2015. Downloaded by PNNL Technical Library on 10/10/2015 00:32:12.
View Article Online
Article
The energy demand for traditional vapor-compressed technology for space cooling continues to soar year after year due to global warming and the increasing human population’s need to improve living and working conditions. Thus, there is a growing demand for eco-friendly technologies that use sustainable or waste energy resources. This review discusses the properties of various refrigerants used for adsorption cooling applications followed by a brief discussion on the thermodynamic cycle. Next, sorbents traditionally used for cooling are reviewed to emphasize the need for advanced capture materials with superior properties to improve refrigerant sorption. The remainder of the review focus on studies using engineered nanoporous frameworks (ENFs) with various refrigerants for adsorption cooling applications. The effects of the various factors that play a role in ENF–refrigerant pair selection, including pore structure/dimension/shape, morphology, open-metal sites, pore chemistry and possible presence of defects, are reviewed. Next, in-depth insights into the sorbent–refrigerant interaction, and pore filling mechanism gained through a combination of characterization techniques and computational modeling are discussed. Finally, we outline the challenges and opportunities related to using ENFs for adsorption cooling applications and provide our views on the future of this technology.
Article
The ethanol/water separation challenge highlights the adsorption capacity/selectivity trade‐off problem. We show that the target guest can serve as a gating component of the host to block the undesired guest, giving molecular sieving effect for the adsorbent possessing large pores. Two hydrophilic/water‐stable metal azolate frameworks were designed to compare the effects of gating and pore‐opening flexibility. Large amounts (up to 28.7 mmol g ⁻¹ ) of ethanol with fuel‐grade (99.5 %+) and even higher purities (99.9999 %+) can be produced in a single adsorption process from not only 95 : 5 but also 10 : 90 ethanol/water mixtures. More interestingly, the pore‐opening adsorbent possessing large pore apertures showed not only high water adsorption capacity but also exceptionally high water/ethanol selectivity characteristic of molecular sieving. Computational simulations demonstrated the critical role of guest‐anchoring aperture for the guest‐dominated gating process.
Article
Full-text available
The ethanol/water separation challenge highlights the adsorption capacity/selectivity trade‐off problem. We show that the target guest can serve as a gating component of the host to block the undesired guest, giving molecular sieving effect for the adsorbent possessing large pores. Two hydrophilic/water‐stable metal azolate frameworks were designed to compare the effects of gating and pore‐opening flexibility. Large amounts (up to 28.7 mmol g⁻¹) of ethanol with fuel‐grade (99.5 %+) and even higher purities (99.9999 %+) can be produced in a single adsorption process from not only 95 : 5 but also 10 : 90 ethanol/water mixtures. More interestingly, the pore‐opening adsorbent possessing large pore apertures showed not only high water adsorption capacity but also exceptionally high water/ethanol selectivity characteristic of molecular sieving. Computational simulations demonstrated the critical role of guest‐anchoring aperture for the guest‐dominated gating process.
Article
High‐purity ethanol is a promising renewable energy resource, however separating ethanol from trace amount of water is extremely challenging. Herein, two ultramicroporous MOFs (UTSA‐280 and Co‐squarate) were used as adsorbents. A prominent water adsorption and a negligible ethanol adsorption identify perfect sieving effect on both MOFs. Co‐squarate exhibits a surprising water adsorption capacity at low pressure that surpassing the reported MOFs. Single crystal X‐ray diffraction and theoretical calculations reveal that such prominent performance of Co‐squarate derives from the optimized sieving effect through pore structure adjustment. Co‐squarate with larger rhombohedral channel is suitable for zigzag water location, resulting in reinforced guest‐guest and guest‐framework interactions. Ultrapure ethanol (99.9%) can be obtained directly by ethanol/water mixed vapor breaking through the columns packed with Co‐squarate, contributing to a potential for fuel‐grade ethanol purification.
Article
Full-text available
High‐purity ethanol is a promising renewable energy resource, however separating ethanol from trace amount of water is extremely challenging. Herein, two ultramicroporous MOFs (UTSA‐280 and Co‐squarate) were used as adsorbents. A prominent water adsorption and a negligible ethanol adsorption identify perfect sieving effect on both MOFs. Co‐squarate exhibits a surprising water adsorption capacity at low pressure that surpassing the reported MOFs. Single crystal X‐ray diffraction and theoretical calculations reveal that such prominent performance of Co‐squarate derives from the optimized sieving effect through pore structure adjustment. Co‐squarate with larger rhombohedral channel is suitable for zigzag water location, resulting in reinforced guest‐guest and guest‐framework interactions. Ultrapure ethanol (99.9 %) can be obtained directly by ethanol/water mixed vapor breaking through the columns packed with Co‐squarate, contributing to a potential for fuel‐grade ethanol purification.
Article
The porous carbon is considered to be one of the most suitable materials for volatile organic compounds (VOCs) adsorption and separation. Revealing the adsorption behavior and molecular interaction mechanism is of great significance for the preparation and synthesis of porous carbon adsorbents. Herein, we investigated detailed mechanistic effects of graphitization degrees and surface oxygen heteroatoms of porous carbon on their benzene and ethanol adsorptive separation performance by combining experiments and theoretical calculations. The results show that graphitized carbon surface has a stronger van der Waals interaction with benzene molecules compared to the disordered carbon surface, and thus exhibit a higher benzene adsorption performance and benzene/ethanol selectivity. While the introduction of oxygen functional groups enhances electrostatic affinity of porous carbon surface, resulting in higher ethanol adsorption capacity and ethanol/benzene selectivity under the entire pressure range. Subsequent theoretical calculations studied the adsorption behaviors of benzene and ethanol molecules in porous carbon models (with different graphitization degrees and oxygen contents) from the perspective of the weak interactions, and found the different molecular configurations of benzene and ethanol is the main leading cause of their different adsorption preferences. This study provided mechanistic insights on how the graphitization degrees and oxygen heteroatoms affected the capture and separation properties of porous carbon, which would further provide a rational alternative strategy in the preparation and synthesis of porous carbons for the effective gas mixture capture and separation.
Article
Crystallinity damage in acid Y zeolite affects the direct relationship between the number of acid sites or conversion of 2-propanol and the zeolite size and the selectivity of 2-propene in nanosized Y zeolite.
Article
Flexible metal-organic frameworks (MOFs) have sparked a lot of interest in view of their appealing features, which emerge from their elasticity and dynamic behaviour, and have a lot of potential applications in a number of sectors. Such MOFs have a unique flexibility that sets them apart from other conventional inorganic porous materials like porous silica and zeolites. These flexible scaffolds are suitable for gas storage or delivery, catalysis, sensing, separation of gases, guest capture and release as well as for scientific investigation. Recent advances in the finding, characterization, and structural research of flexible MOFs are discussed in this paper. Up to now, a novel category of MOFs having flexible frameworks has evolved, labeled as soft porous crystals, flexible MOFs, breathing MOFs, sponge- like MOFs, or dynamic MOFs. We explore many viewpoints and ways to understand and control flexible behaviour in a representative MOF platform in this review with the aim of conveying the fundamental ideas to a broad audience of academics, engineers, and materials scientists. Furthermore, prospective implications of flexible MOFs will be briefly mentioned, as well as advanced analytic methods and simulation that aid in the investigation of MOF flexibility under diverse situations and provide information into the creation of novel dynamic MOFs. The multiscale design of flexible MOF systems is required to translate molecular structural modifications to macroscopic responses. This review also covers advances in this domain over the last ten years, beginning with modular fabrication of basic components and ending with the regulation of dynamics in device architectures. The recent applications which make such MOFs intriguing are also sketched out with their future prospects.
Article
Full-text available
Metal-organic frameworks comprise an important class of solid-state materials and have potential for many emerging applications such as energy storage, separation, catalysis and bio-medical. Here we report the adsorption behaviour of a series of fluorocarbon derivatives on a set of microporous and hierarchical mesoporous frameworks. The microporous frameworks show a saturation uptake capacity for dichlorodifluoromethane of >4 mmol g(-1) at a very low relative saturation pressure (P/Po) of 0.02. In contrast, the mesoporous framework shows an exceptionally high uptake capacity reaching >14 mmol g(-1) at P/Po of 0.4. Adsorption affinity in terms of mass loading and isosteric heats of adsorption is found to generally correlate with the polarizability and boiling point of the refrigerant, with dichlorodifluoromethane >chlorodifluoromethane >chlorotrifluoromethane >tetrafluoromethane >methane. These results suggest the possibility of exploiting these sorbents for separation of azeotropic mixtures of fluorocarbons and use in eco-friendly fluorocarbon-based adsorption cooling.
Article
Full-text available
In this experimental study, the adsorption behavior of the ZIF-68 heterolinked zeolitic imidazolate framework has been explored. Vapor phase adsorption isotherms of linear C1-C6 alcohols, C6 alkane isomers, aromatics (benzene, toluene, xylene isomers, 1,3,5-trimethylbenzene and 1,3,5-triisopropylbenzene) and polar adsorbates (water, acetonitrile and acetone) are reported and discussed. The complex pore structure of ZIF-68, with two one-dimensional channels, each with a different polarity, displays an overall hydrophobic character. Its two-pore system results in S-shaped isotherms for small polar adsorbates (small alcohols, acetone and acetonitrile), while longer alcohols and nonpolar molecules, such as aromatics and C6 alkane isomers, lead to type I adsorption isotherms. Bulky molecules, with a kinetic diameter significantly larger than the pore windows, are adsorbed in large amounts, which gave reason to think that this ZIF-68 material has a certain degree of framework flexibility to enlarge the free aperture of the channels. Besides, diffusion coefficients from vapor phase uptake and infrared experiments point to a different adsorption mechanism for polar and nonpolar adsorbates. Liquid phase adsorption experiments demonstrated the separation of alcohol mixtures (ethanol/1-butanol) at low concentration from water, with a clear preference for 1-butanol.
Article
The selective removal of water from mixtures with methanol, ethanol, and 1-propanol is an important task in the processing industries. With the aid of Configurational-Bias Monte Carlo simulations of unary and mixture adsorption, we establish the potential of CuBTC for this separation task. For operations close to pore saturation conditions, the adsorption is selective to water, that has a significantly higher saturation capacity compared to that of 1-alcohols. The water-selective separation relies on subtle entropy effects that manifest near pore saturation conditions. A further distinguishing feature is that mixture adsorption is determined to be strongly non-ideal, and the activity coefficients of the constituent components deviate strongly from unity as pore saturation is approached. The predictions of the Ideal Adsorbed Solution Theory (IAST), though qualitatively correct, do not predict the component loadings for mixture adsorption with adequate accuracy. Consequently, the activity coefficients, after appropriate parameterization, have been incorporated into the Real Adsorbed Solution Theory (RAST). Transient breakthrough simulations, using the RAST model as basis, demonstrate the capability of CuBTC for selective adsorption of water in fixed bed adsorption devices operating under ambient conditions.
Article
The total world energy demand is predicted to rise significantly over the next few decades, primarily driven by the continuous growth of the developing world. With rapid depletion of nonrenewable traditional fossil fuel, which currently accounts for almost 86% of the worldwide energy output, the search for viable alternative energy resources is becoming more important from a national security and economic development standpoint. Nuclear energy, an emission free, high energy density source, produced by means of controlled nuclear, is often considered as a clean, affordable alternative of fossil fuel. However, the successful installation of an efficient and economically viable industrial scale process to properly sequester and mitigate the nuclear fission related, highly radioactive waste (e.g. used nuclear fuel or UNF) is a prerequisite for any further development of nuclear energy in near future. The reprocessing of UNF is often considered a logical way to minimize the volume of high-level radioactive waste, though the generation of volatile radionuclides during reprocessing raises a significant engineering challenge for its successful implementation. The volatile radionuclides include, but are not limited to, noble gases (predominately isotopes of Xe and Kr) and must be captured during the process to avoid being released into the environment. Currently, energy intensive cryogenic distillation is the primary means to capture and separate radioactive noble gas isotopes during UNF reprocessing. A similar cryogenic process is implemented during commercial production of noble gases though removal from air. In light of high commercial values particularly in lighting and medical industries and associated high production cost, alternate approaches for Xe/Kr capture and storage are of contemporary research interest. The proposed pathways for Xe/Kr removal and capture can essentially be divided in two categories: selective absorption by dissolution in solvents and physisorption on porous materials. Physisorption-based separation and adsorption on highly functional porous materials are promising alternatives for energy intensive cryogenic distillation process, where the adsorbents are characterized by high surface areas for removal capacities and often can be chemically fine-tuned to enhance the adsorbate–adsorbent interactions for optimum selectivity. Several traditional porous adsorbents such as zeolites and activated carbon have been tested for noble gas capture, but have shown low capacity, selectivity, and lack of modularity. Metal-organic frameworks (MOFs) or porous coordination polymers (PCPs) are an emerging class of solid-state adsorbents, tailor-made for applications ranging from gas-adsorption and separation to catalysis and sensing. Herein, we give a concise summary on the background and development of Xe/Kr separation technologies with focus to UNF reprocessing and the prospect of MOF-based adsorbents for that particular application.
Article
We examine mixture separations with microporous adsorbents such as zeolites, metal-organic frameworks (MOFs) and zeolitic imidazolate frameworks (ZIFs), operating under conditions close to pore saturation. Pore saturation is realized, for example, when separating bulk liquid phase mixtures of polar compounds such as water, alcohols and ketones. For the operating conditions used in industrial practice, pore saturation is also attained in separations of hydrocarbon mixtures such as xylene isomers and hexane isomers. Separations under pore saturation conditions are strongly influenced by differences in the saturation capacities of the constituent species; the adsorption is often in favor of the component with the higher saturation capacity. Effective separations are achieved by exploiting differences in the efficiency with which molecules pack within the ordered crystalline porous materials. For mixtures of chain alcohols, the shorter alcohol can be preferentially adsorbed because of its higher saturation capacity. With hydrophilic adsorbents, water can be selectively adsorbed from water-alcohol mixtures. For separations of o-xylene-m-xylene-p-xylene mixtures, the pore dimensions of MOFs can be tailored in such a manner as to allow optimal packing of the isomer that needs to be adsorbed preferentially. Subtle configurational differences between linear and branched alkane isomers result in significantly different packing efficiencies within the pore topology of MFI, AFI, ATS, and CFI zeolites. A common characteristic feature of most separations that are reliant on molecular packing effects is that adsorption and intra-crystalline diffusion are synergistic; this enhances the separation efficiencies in fixed bed adsorbers.
Article
Solvothermal reactions of 5-(4-pyridyl)-isophthalic acid (H2pbdc) and transition-metal centers (Ni2+/Co2+/Zn2+) in the presence (or absence) of N-auxiliary 4,4′-bis(1-imidazolyl)biphenyl (bimb) ligand produce [Ni2(pbdc)2(μ2-H2O)(H2O)2·(DMA)2.7]n (DMA = N,N′-dimethylacetamide, 1), [Ni12(pbdc)12(μ2-H2O)6(py)2(H2O)8(DMA)2·(H2O)5·(DMA)9]n (2), [Co2(pbdc)2(bimb)2·(bimb)0.5·(H2O)4·(DMF)0.25]n (3) and [Zn(pbdc)(bimb)·(H2O)]n (4), which exhibit structural diversity. Both compounds 1 and 2 display a uninodal 8-connected 3D tsi net, but feature different crystal systems and space groups from each other. Compound 3 adopts a 2-fold interpenetrating binodal (3,5)-connected 3D hms net and compound 4 features a rare 2-fold interpenetrating binodal (3,4)-connected 3D fsx architecture. In particular, activated 3 shows high-efficiency for the selective sorption of small molecules, including CO2 over N2 and CH4, H2 over N2, as well as alcohols from water. More importantly, 4 represents the first report on a MOF as a promising luminescent probe for detecting pesticides, and also the very first example for detecting both pesticides and solvent molecules simultaneously.
Article
We report the synthesis, structural characterization, and porous properties of two isomeric supramolecular complexes of ([Cd(NH2 bdc)(bphz)0.5 ]⋅DMF⋅H2 O}n (NH2 bdc=2-aminobenzenedicarboxylic acid, bphz=1,2-bis(4-pyridylmethylene)hydrazine) composed of a mixed-ligand system. The first isomer, with a paddle-wheel-type Cd2 (COO)4 secondary building unit (SBU), is flexible in nature, whereas the other isomer has a rigid framework based on a μ-oxo-bridged Cd2 (μ-OCO)2 SBU. Both frameworks are two-fold interpenetrated and the pore surface is decorated with pendant -NH2 and NN functional groups. Both the frameworks are nonporous to N2 , revealed by the type II adsorption profiles. However, at 195 K, the first isomer shows an unusual double-step hysteretic CO2 adsorption profile, whereas the second isomer shows a typical type I CO2 profile. Moreover, at 195 K, both frameworks show excellent selectivity for CO2 among other gases (N2 , O2 , H2 , and Ar), which has been correlated to the specific interaction of CO2 with the -NH2 and NN functionalized pore surface. DFT calculations for the oxo-bridged isomer unveiled that the -NH2 group is the primary binding site for CO2 . The high heat of CO2 adsorption (ΔHads =37.7 kJ mol(-1) ) in the oxo-bridged isomer is realized by NH2 ⋅⋅⋅CO2 /aromatic π⋅⋅⋅CO2 and cooperative CO2 ⋅⋅⋅CO2 interactions. Further, postsynthetic modification of the -NH2 group into -NHCOCH3 in the second isomer leads to a reduced CO2 uptake with lower binding energy, which establishes the critical role of the -NH2 group for CO2 capture. The presence of basic -NH2 sites in the oxo-bridged isomer was further exploited for efficient catalytic activity in a Knoevenagel condensation reaction.
Article
A molecular simulation study is reported for the adsorption of normal alcohols (methanol, ethanol, propanol, and butanol) in zeolitic imidazolate framework-8 (ZIF-8). The effects of force fields, atomic charges, and framework flexibility are systematically examined and compared with experimental data. Among three force fields (UFF, AMBER, and DREIDING), DREIDING has the best agreement with experiment. The atomic charges and framework flexibility are found to have negligible effects. The four alcohols exhibit S-shaped isotherms without hysteresis loop, as attributed to adsorption at different preferential sites. At a low pressure, cluster formation is observed near the organic linker (2-methylimidazolate) in ZIF-8; with increasing pressure, cage-filling occurs in the large sodalite cage. The interaction between alcohol and ZIF-8 framework is enhanced as the chain length of alcohol increases; thus, the isosteric heat of adsorption rises with chain length. The simulation study provides microscopic insight
Article
The efficacy of nanoporous crystalline materials in separation applications is often influenced to a significant extent by diffusion of guest molecules within the pores of the structural frameworks. The Maxwell–Stefan (M–S) equations provide a fundamental and convenient description of mixture diffusion. The M–S formulation highlights two separate factors that cause mixture diffusion to be intrinsically coupled: correlation effects, and thermodynamic coupling.By careful and detailed analyses of a variety of published experimental data on (a) mixture permeation across nanoporous membranes, (b) transient uptake of mixtures within crystals, and (c) transient breakthrough characteristics of fixed bed adsorbers, we identify conditions that require the use of M–S equations including both correlation effects and thermodynamic coupling. Situations are also identified in which either of the coupling effects can be ignored.Correlation effects cause slowing-down of more-mobile-less-strongly-adsorbed molecules by tardier-more-strongly-adsorbed-partner species; such slowing-down effects are often essential for modeling mixture permeation across nanoporous membranes. Overshoots in the transient uptake of the more mobile partners in single crystals are essentially the consequence of thermodynamic coupling, originating from sizable off-diagonal elements of thermodynamic correction factors Γij.In the case of transient breakthrough of hexane isomers in a fixed bed of MFI zeolite, we show that thermodynamic coupling effects lead to a significant improvement in the separation performance.