ChapterPDF Available

Metal-organic frameworks – synthesis and applications

Authors:

Figures

Content may be subject to copyright.
INDUSTRIAL CATALYSIS
AND SEPARATIONS
Innovations for Process Intensification
Edited by
K. V. Raghavan, PhD, and B. M. Reddy, PhD
Apple Academic Press
TORONTO NEW JERSEY
Apple Academic Press Inc. Apple Academic Press Inc.
3333 Mistwell Crescent 9 Spinnaker Way
Oakville, ON L6L 0A2 Waretown, NJ 08758
Canada USA
©2015 by Apple Academic Press, Inc.
Exclusive worldwide distribution by CRC Press, a member of Taylor & Francis Group
No claim to original U.S. Government works
Printed in the United States of America on acid-free paper
International Standard Book Number-13: 978-1-926895-96-3 (Hardcover)
This book contains information obtained from authentic and highly regarded sources. Reprinted
material is quoted with permission and sources are indicated. Copyright for individual articles
remains with the authors as indicated. A wide variety of references are listed. Reasonable efforts
have been made to publish reliable data and information, but the authors, editors, and the publisher
cannot assume responsibility for the validity of all materials or the consequences of their use. The
authors, editors, and the publisher have attempted to trace the copyright holders of all material
reproduced in this publication and apologize to copyright holders if permission to publish in this
form has not been obtained. If any copyright material has not been acknowledged, please write and
let us know so we may rectify in any future reprint.
Trademark Notice: Registered trademark of products or corporate names are used only for explanation
and identication without intent to infringe.
Library of Congress Control Number: 2014938654
Library and Archives Canada Cataloguing in Publication
Industrial catalysis and separations: innovations for process intensication/edited by
K.V. Raghavan, PhD, and B.M. Reddy, PhD.
Includes bibliographical references and index.
ISBN 978-1-926895-96-3 (bound)
1. Catalysis--Industrial applications. 2. Green chemistry. 3. Chemical processes.
4. Chemical engineering. 5. Sustainable engineering. I. Reddy, B. M., author, editor
II. Raghavan, K. V., author, editor
TP156.C35I54 2014 660'.2995 C2014-902755-9
Apple Academic Press also publishes its books in a variety of electronic formats. Some content that
appears in print may not be available in electronic format. For information about Apple Academic
Press products, visit our website at www.appleacademicpress.com and the CRC Press website at
www.crcpress.com
CHAPTER 3
METAL ORGANIC FRAMEWORKS–
SYNTHESIS AND APPLICATIONS
RADHA KISHAN MOTKURI, JIAN LIU, CARLOS A FERNANDEZ,
SATISH K NUNE, PRAVEEN THALLAPALLY, and B. PETE MCGRAIL
CONTENTS
3.1 Introduction ......................................................................................... 63
3.2 Synthesis and Characterization ...........................................................65
3.3 Sorption applications ........................................................................... 68
3.3.1 Gas Storage and Separation ..................................................... 68
3.3.2 Hydrogen Storage and Separation ........................................... 68
3.3.3 Methane Storage ...................................................................... 70
3.3.4 Carbon Dioxide Capture From Flue Gas ................................. 71
3.3.5 Water Sorption Studies for Adsorption Cooling
Applications .........................................................................................75
3.3.6 Harmful Gas Removal ............................................................. 76
3.4 Heterogeneous Catalysis, Sensor and Other Applications ..................78
3.4.1 Framework Activity ................................................................. 78
3.4.2 Organic or Pseudo-Organic Linkers as Active Sites ...............80
3.4.3 Catalysis with Tailored Microenvironment ............................. 80
3.4.4 MOFS for Sensing Applications ..............................................83
3.4.5 MOF Sensors Based on Piezoresistivity .................................83
3.4.6 Gravimetric-Based MOF Sensors ...........................................84
3.4.7 Surface Acoustic Wave-Based MOF Sensors: ........................ 85
Chapter 3. METAL ORGANIC FRAMEWORKS–SYNTHESIS AND APPLICATIONS
Radha Kishan Motkuri,* Jian Liu, Carlos A . Fernandez, Satish K . Nune, Praveen Thallapally, B . P. McGrail
Citation Information: “Industrial Catalysis and Separations Innovations for Process Intensification “ by K . V . Raghavan and B . M .
Reddy; Apple Academic Press 2014, Pages 61–103
Print ISBN: 978-1-926895-96-3; eBook ISBN: 978-1-4822-3426-8; DOI: 10.1201/b17114-6
62 Industrial Catalysis and Separations: Innovations for Process Intensification
3.5 Sensors Based on MOF Optical Properties ......................................... 86
3.5.1 Photoluminescence (PL)-Based MOF Sensors .......................86
3.5.2 Solvatochromism/Vapochromism ...........................................86
3.6 MOFs for Oil Spill Cleanup ................................................................ 88
3.7 Conclusions ......................................................................................... 92
Keywords ...................................................................................................... 93
Acknowledgments .........................................................................................94
References ..................................................................................................... 94
Metal Organic Frameworks–Synthesis and Applications 63
3.1 INTRODUCTION
Porous materials have received enduring interest in chemical science because
of their suitability for many applications, including host materials for molecu-
lar separation and storage, catalysis, molecular sensing, magnetism, and drug
delivery systems. These porous solids are broadly classified into two major
categories: (i) amorphous and (ii) crystalline. Porous amorphous solids (e.g.,
plastics, gels) do not exhibit any ordered repeated units within their structures,
but it is beneficial to work with them as they are usually less expensive and
easy to prepare. Their primary disadvantages are the potentially wide range
of molecular architectures with nonpredictable channels or topologies and the
lack of long-range order that results in low mechanical stability. In contrast,
the porous crystalline solids are more advantageous because of their ordered
structures with reproducible pores/channels and topologies, which give them
high thermal and mechanical stability. Nanoporous silica and zeolites are
characteristic examples of such ordered porous solids that have predictable
structural features with reproducible pores/channels, dimensions, and topolo-
gies. These porous solids are subdivided into three categories based on their
pore size. According to the International Union of Pure and Applied Chem-
istry, the porosity of crystalline solids is usually given by the diameter of the
pore size and is categorized as microporous (5–20 Å), mesoporous (20–500
Å), and macroporous (>500 Å) [1].
The inorganic microporous zeolites, both natural and synthetic, construct-
ed from aluminosilicates have been extensively studied for decades for a
wide variety of applications. Natural zeolites, which are predominantly ocean
based, have various uses, but their application is limited due to the difculty
in eliminating impurities. However, synthetic zeolites have been used exten-
sively for catalytic applications, adsorption, and separation because they can
be systematically tailored under controlled conditions to meet user require-
ments. For example, zeolite 4Å (Na12Al12Si12O48) with pore size of 4Å was
commercially used as a drying agent, while zeolite ZSM-5 served as a crack-
ing agent. Similarly, Silicalite-1 was extensively used as a porous sorbent for
removal of gasoline from drinking water [2]. Although zeolites have been
widely used, they possess a number of drawbacks that limit their utility. The
disadvantages include difculty in controlling the synthesis conditions, the
very limited number of structure and channel topologies (178 to date), and the
64 Industrial Catalysis and Separations: Innovations for Process Intensification
fact that they are not easily changeable covalently bonded crystalline struc-
tures.
In the last decade, a new class of porous solids called metal-organic
frameworks (MOFs) or porous coordination polymers (PCPs) was discov-
ered. These porous solids exhibit unique and outstanding properties such as
the largest pore volumes and highest surface areas known [3]. The tunable
nanoporosity of MOFs places them far ahead of zeolites and other metal ox-
ides. MOFs are considered as organic analogs of inorganic zeolites in, which
oxygen atoms are replaced by rigid organic ligands/linkers that bridge the
inorganic ions or ion clusters that self-assemble into one-, two-, and three-
dimensional frameworks [4]. The advantage of this class of materials is that
by carefully choosing the building blocks such as inorganic metal ions and
organic ligands, it is possible to tailor the structures, pore sizes, and their to-
pologies [5]. Because of the large selection of transition and lanthanide metal
ions with diverse coordination geometries and the possibility of creating an
even larger number of organic linkers via synthetic organic chemistry, an im-
mensely diverse range of MOF framework architectures can be generated [6].
The schematic representations of the general classication of porous solids
are shown in Fig. 3.1.
FIGURE 3.1 Schematic representations of the general classification of porous solids. A
typical construction procedure for MOFs is given in the bottom panel [7]. (Reproduced
from Li, J. R. et al., Chem Rev 112, 869–932 (2012), with permission).
Metal Organic Frameworks–Synthesis and Applications 65
The coordination polymers were discovered in the 1950s [8], and their
discovery was followed by work reported by Jacobson [9], Bujoli [10], and
Férey [11]. However, the major interest in these materials was boosted in the
late 1990s by a pioneering material chemist Omar Yaghi and his co-workers.
Yaghi and his team have been the most prolic group in synthesizing rigid
coordination networks with metal cations and organic ligands. They started
by replacing the 4,4′-bipyridine moieties with multidentate carboxylates li-
gands and named the resulting neutral frameworks as MOFs [12]. Amongst
them, MOF-5 is the well studied in the literature and its structure aided the
development of new concepts in design and tuning of MOFs. Later, because
of the unique and outstanding properties of these architectures, an exponential
increase in publications related to the development of new materials and their
applications was observed. The historical developments in MOF synthesis are
summarized in Fig. 3.2 [6].
FIGURE 3.2 Historical developments in MOF synthesis [6]. (Reproduced from Stock,
N. et al., Chem Rev 112, 933–969 (2012), with permission).
3.2 SYNTHESIS AND CHARACTERIZATION
The synthesis of MOFs is a moderately young field, and its evolution from
well-developed solid-state/zeolite chemistry is noteworthy. Different types
of synthesis methods were reported in the literature such as classical hydro
(solvo) thermal synthesis, microwave heating [13], diffusion [14], ultrasonic
[15], mechanochemical [16], electrochemical synthesis [17], and more recently
66 Industrial Catalysis and Separations: Innovations for Process Intensification
spray-drying [18]. However, conventional hydrothermal or solvothermal
heating of organic, inorganic precursors dissolved in a suitable solvent system
in closed vessels under autogeneous conditions was well studied for major
MOF synthesis. Both compositional parameters, such as solvent, pH, and mo-
lar concentration, and process parameters, such as temperature, pressure, and
reaction time, will play a large role in the formation of framework structures.
In a clear solution of dissolved reactants, the concentration gradient needs to
be adjusted to exceed the critical nucleation concentration that enables the
formation of crystal growth. The concentration gradient can be achieved by
various methods, including changing the temperature of the system, solvent
evaporation, layering, or slow diffusion of two or multiple solutions contain-
ing the reactants. Along with the powders, various morphologies such as thin
films, membranes, and composites have been successfully synthesized for
various applications (Fig. 3) [6].
FIGURE 3.3 Various synthesis methods and morphologies in MOFs [6]. (Reproduced
from Stock, N. et al., Chem Rev 112, 933–969 (2012), with permission).
Metal Organic Frameworks–Synthesis and Applications 67
In contrast to zeolites, MOFs have signicantly weaker bonds (e.g., co-
ordinative and hydrogen bonds) between the organic–inorganic connections.
These weaker connections are responsible for MOF structural exibility un-
der external stimuli such as temperature [19], mechanical pressure [20], elec-
tric eld, magnetic eld, light, gas [21], and liquid/vapor [22] sorption. More
specically, sorption-induced structural deformation or “breathing” has been
studied extensively to determine its effect on chemical, structural, and topo-
logical changes in MOFs in the presence or absence of the guest [23]. The
different phases of a exible MOF, MIL-53, resulting from increasing the
amount adsorbed are shown in Fig. 3.4 [24]. Recently, it has been demon-
strated that postsynthetic modication methods successfully showed the pos-
sibility to tune the exibility in the material for tailored dynamic behavior in
MOFs [25].
FIGURE 3.4 Different phases of dynamic (flexible) metal-organic framework MIL-53
by increasing the amounts adsorbed. (Reproduced from Llewellyn, P.L. et al., J Am Chem
Soc 131, 13002–13008 (2009), with permission) [24].
Scalability is an important issue for any material for industrial applica-
tions. Large-scale synthesis of MOFs for industrial applications was success-
fully performed by BASF by considering the classical solvothermal synthetic
approach, which is generally used in industrial production of zeolites [26].
In large-scale synthesis (kilotons/year), the availability and cost of reactants
and the space–time-yield (STY) of synthesis play a crucial role. The higher
the STY value (kg of MOF per m3 of reaction mixture per day of synthesis),
the greater the feasibility for MOF production, industrially. The inorganic
salts (nitrates, oxides, sulfates) are readily available and inexpensive, while
the organic linkers are usually expensive and become a signicant compo-
nent of the raw material costs. The STY values for Basolite M050 (Mg(II)/
formic acid) and Basolite A520 (Al(III)/fumaricacid) are >3000 kg/(m3 D)
68 Industrial Catalysis and Separations: Innovations for Process Intensification
while Basolite Z1200 (ZIF-8) has 160 kg/(m3 D). The typical zeolites that
produced by BASF has a STY value of 50–150 kg/(m3 D) respectively [26].
Various characterization techniques have been successfully used to study
the MOFs, most notably single crystal and powder X-ray diffraction (XRD)
studies for getting the crystal morphology, arrangement of atoms and pore
structure; BET and pore size distribution studies for getting the informa-
tion of surface area and pore width and volume; Infrared spectroscopy stud-
ies for getting the information of functional groups in the MOFs structures;
Thermogravimetric analysis for getting information of stability and optimal
activation temperature ; Gas adsorption studies for getting the information
of gas storage capacities, gas adsorption selectivities, and diffusivities of gas
molecules in MOFs.
3.3 SORPTION APPLICATIONS
3.3.1 GAS STORAGE AND SEPARATION
MOFs are widely considered as promising novel adsorbents for gas storage
and separation due to their high surface areas, tunable pore size and structures,
and versatile chemical compositions [28, 29].
3.3.2 HYDROGEN STORAGE AND SEPARATION
Hydrogen is considered to be one of the best alternative fuels to fossil fuels
because of its high energy density, nonpolluting combustion products, and
natural abundance. However, H2 is an extremely volatile gas under ambient
conditions, resulting in a volumetric energy density that is much too low for
practical applications. The goal therefore is to design lightweight materials
that can reversibly and rapidly store H2 near ambient conditions at a density
equal to or greater than liquid hydrogen. MOFs have attracted considerable
attention in the H2 storage area in recent years because of their high working
capacities [27].
One of the rst MOFs investigated for H2 storage was the cubic carbox-
ylate-based framework MOF-5 or IRMOF-1, IRMOF-6, and IRMOF-8 (Fig.
3.5) [28]. With complete activation, MOF-5 can adsorbs 7.1 wt% H2 at 77
K and 40 bar and 10.0 wt% at 100 bar. The latter value is corresponding to
a record volumetric storage density of 66 g L–1, which is near the density of
liquid H2 at 20.4 K and 1 bar [29]. This MOF material is the best cryogenic
storage material currently known. In addition, it was demonstrated that H2 can
Metal Organic Frameworks–Synthesis and Applications 69
be loaded into MOF-5 within 2 min, and the sample can maintain a revers-
ible capacity for at least 24 cycles [30]. Researchers have reported hydrogen
storage data for more than 150 other microporous MOFs since MOF-5 was
investigated as a novel adsorbent for H2 [31]. Various strategies have been
adopted to improve H2 adsorption in MOFs. Open metal coordination sites (or
unsaturated metal centers) in the structures of MOFs can increase the interac-
tion between H2 molecules and MOFs, resulting in an improved performance.
This was rst demonstrated in the MOF Mn3[(Mn4Cl)3(BTT)8]2, which con-
tains open Mn2+ coordination sites and exhibits an isosteric heat of adsorption
of 10.1 kJ mol-1 at zero coverage [32].
FIGURE 3.5 MOFs, topologically the isoreticular MOFs (IRMOFs), are based upon an
augmented simple cubic net (the boron net in CaB6), while MOF-177 is based upon the
augmented form of the (3,6)-coordinated net qom. The large pore of each structure is
represented by a yellow sphere with diameter defined by the distance between the van der
Waals surfaces of the framework atoms. Atom colors: C, black; O, red; Zn, blue tetrahedra;
H, omitted [28a]. (Reproduced from Rowsell, J.L.C. et al., Angew Chem Int Edit 44, 4670–
4679 (2005), with permission).
In addition, surface area plays an important role in H2 adsorption on solid
materials. As a general rule, the maximum or saturated adsorbed amount of
gas on a solid surface is dependent on its surface area [27a]. When the pres-
sure of H2 is not high enough to reach its saturated adsorption, the amount of
the adsorbed H2 is determined mainly by the interactions between molecules
and MOF structures that can be positively related with the heats of adsorp-
tion [33]. Pore size can also affect the interaction between H2 molecules and
MOFs. It was found that MOFs having small pores with walls of high curva-
tures interact with H2 molecules more strongly than large-pore MOFs [34].
The ideal pore size is slightly larger than the kinetic diameter of H2 (2.8 Å) for
low-pressure adsorption because, under that condition, H2 molecules can
70 Industrial Catalysis and Separations: Innovations for Process Intensification
interact with a larger portion of the framework, which increases the interac-
tion energy between the frameworks and H2 molecules [34b].
MOFs have displayed outstanding performance for cryogenic H2 storage
at 77 K and pressures up to 100 bar. Improvements in gravimetric capac-
ity and in volumetric storage density are possible, provided advances in the
synthetic chemistry. The challenge is to design ligands or surface functional
groups that will lead to a high density of strong open metal sites, which are
capable of binding more than one H2 molecule with a binding enthalpy of
about 20 kJ mol–1. Thus, the creation of MOFs for practical applications in H2
storage is still a difcult yet engaging challenge [27b].
On the other hand, MOFs are also promising adsorbents for the purication
of H2 due to their tunable porous structures. Molecular sieving using MOFs
has been used to separate H2 and CH4 due to their different kinetic diameters
(H2: 2.8 Å; CH4: 3.8 Å) [35]. In addition, Long’s group has recently studied
some representative MOFs, including MOF-177, H3[(Cu4Cl)3-(BTTri)8], and
Mg-MOF-74, for separating CO2 and H2 via pressure swing adsorption [36].
These measurements were performed under conditions (at 313 K and pres-
sures up to 40 bar) close to practical application for precombustion CO2 cap-
ture. These MOFs showed excellent separation performances, with evaluated
CO2/H2 selectivities between 2 and 860 on an 20:80 CO2/H2 gas mixture. In
addition to experimental observations, molecular simulations have been used
to investigate the separation of H2, CO2, and CH4 using MOFs [37].
3.3.3 METHANE STORAGE
Natural gas, which consists mainly of CH4, is widely employed as feedstock
for synthesis gas in many countries. It is currently stored as compressed natu-
ral gas (CNG) at 207 bar in pressure vessels, requiring an expensive multi-
stage compression. An attractive alternative to CNG is adsorbed natural gas
(ANG), in, which the gas is stored as an adsorbed phase in a porous solid at
a lower pressure [38]. To promote the vehicular application of methane, the
U.S. DOE has set the target for methane storage at 180 v(STP)/v under 35 bar,
near ambient temperature, with the energy density of ANG being comparable
to that of CNG currently used. An ideal material for CH4 adsorption should
have not only a large accessible surface area but also a large pore volume, a
low framework density, and strong energetic interactions between the frame
work and CH4 molecules. GCMC simulation has been used as a screening
tool to identify new candidates for methane storage and to guide the design
of new materials.
Metal Organic Frameworks–Synthesis and Applications 71
A microporous PCN-14 MOF based on a predesigned anthracene deriva-
tive, 5, 5′-(9,10-anthracenediyl)-di-isophthalate, was synthesized for CH4 up-
take because it contains nanoporous cages as shown in Fig. 3.6. It has an
estimated Langmuir surface area of 2176 m2/g and an estimated pore volume
of 0.87 cm3/g. More important, PCN-14 MOF exhibits an absolute methane-
adsorption capacity of 230 v/v (28% higher than the DOE target of 180v/v
at ambient temperature s) and heats of adsorption of methane of around 30
kJ/mol. These results indicate that PCN-14 is a promising adsorbent for CH4
storage and can be used to store the fuel in the natural-gas-fueled vehicles.
FIGURE 3.6 The crystal structure of the PCN-14 MOF. (a) Squashed cuboctahedral cage
and (b) nanoscopic cage with 18 vertices, 30 edges, and 20 faces. Color scheme: C, gray;
Cu, turquoise; and O, red [39]. (Reproduced from Ma, S.Q. et al., J Am Chem Soc 130,
1012–1016 (2008), with permission).
3.3.4 CARBON DIOXIDE CAPTURE FROM FLUE GAS
The post combustion separation of CO2 from power plant flue gas is of great
interest to mitigate the global warming phenomenon [40]. The sorption of
CO2 using various MOFs and related porous materials is well documented in
the literature [21, 41]. The CO2 partial pressure in a typical flue gas is about
0.15 bar, which is much lower than atmospheric pressure. Therefore, it is im-
portant to study CO2 adsorption in MOFs at low pressures. Yazaydin and col-
leagues used both experiments and simulation to screen MOFs for the highest
CO2 capacities at about 0.1 atm [42]. They found that Mg/DOBDC and Ni/
DOBDC (also known as Mg-MOF-74 and Ni-MOF-74 or CPO-27-Mg and
CPO-27-Ni) have the highest CO2 capacities at 0.1 atm and 298 K, which are
5.95 mol kg–1 and 4.07 mol kg–1, as shown in Fig. 3.7.
72 Industrial Catalysis and Separations: Innovations for Process Intensification
FIGURE 3.7 Experimental CO2 uptake in different MOFs at 0.1 bar. Data were obtained
at 293–298 K [42]. (Reproduced from Yazaydin, A.O. et al., J Am Chem Soc 131, 18198
(2009), with permission).
Instead of the surface area or the free volume, the authors found that MOFs
such as Mg/DOBDC and Ni/DOBDC with a high density of open metal sites
are promising candidates for CO2 capture from ue gas in, which CO2 par-
tial pressure is about 0.1 atm. Caskey et al. [43] found that metal substitu-
tion in the DOBDC series can signicantly impact their CO2 capacities in
the low-pressure region. The metal substitution effect may be caused by the
differences in the ionic character of the metal–oxygen bonds in the DOBDC-
series MOFs. Liu et al. found that Ni/DOBDC has a higher CO2 capacity than
NaX and 5A zeolites at 0.1 atm, and 25°C. In addition, water does not affect
CO2 adsorption in the Ni/DOBDC as much as in NaX and 5A zeolites, and it
is much easier to remove water from Ni/DOBDC by heat regeneration [44].
Therefore, the Ni/DOBDC can adsorb more CO2 than traditional zeolites un-
der the same moist conditions.
Most ue gas is composed of N2, so it is important to have a MOF that
can selectively adsorb CO2 over N2. Motkuri et al. showed that Prussian blue
Metal Organic Frameworks–Synthesis and Applications 73
derivatives represent a class of highly stable ordered crystalline structures
that can selectively adsorb CO2 over N2, CH4, and water present in ue and
natural gas conditions [41]. Seven MOFs, including CuBTC, MIL-47 (V), IR-
MOF-1, IRMOF-12, IRMOF-14, IRMOF-11, and IRMOF-13, were studied
for the separation performance of CO2 over N2 by Liu and Smit using GCMC
simulations [45]. In all the MOFs they considered, CO2 is more preferentially
adsorbed than N2, with CuBTC showing the highest selectivity. Motkuri and
team found that pore size plays an important role in the selective adsorption of
CO2 over N2 [47]. The effect of the chemistry of the materials (i.e., effects of
the electrostatic interaction) becomes less evident compared to the effects of
pore size because both CO2 and N2 molecules have quadruple moments. As a
result, the electrostatic interactions will help increase the adsorption for both
of them. Wu et al. obtained a Li-modied IRMOF-1, chem-4Li MOF, by sub-
stituting all the hydrogen atoms with O–Li groups in the aromatic rings of IR-
MOF-1.The chem-4Li MOF was found to have an extraordinarily high CO2/
N2 selectivity of 395 (CO2:N2 = 15.6:84.4), which is two orders of magnitude
larger than that of the original IRMOF-1 [46]. This selectivity is caused by
the strong electrostatic interactions between the gas molecules and the lithium
atoms in the framework. Thallapally and co-workers synthesized eight dis-
similar metal-organic supramolecular isomers using one exible tetrahedral
organic linker, tectonic acid (TA), with different metals Zn, Cd, Co, Mg, and
Cu [22, 47]. In the MOF constructed with Zn, addition of auxiliary ligands
(pyrazine, 1, 4-diazabicyclo [2.2.2]octane or DABCO) yielded MOFs with no
pores to mesopores. It is interesting to note that the auxiliary ligand bipyri-
dineinvolved in the crystal formation resulted in a exible MOF formation,
TetZB (isomer A in Fig. 3.8) that showed a breathing phenomenon upon sorp-
tion of gasses/vapors. All other TA-based MOFs showed preferable sorption
of CO2 over N2, H2, and methane at room temperature.
Meanwhile, selectivity between H2O and CO2 is important for using
MOFs to separate CO2 from ue gas because a typical ue gas is saturated
with water vapor. Some research work shows that a small amount of water
can help enhance CO2 adsorption in HKUST-1 [48]. The enhanced CO2 up-
take is caused by interactions between the quadruple moment of CO2 and the
electric eld created by the coordinated water molecules. However, further
increasing the water loading on the HKUST-1 will result in considerable un-
coordinated water molecules that block pore space and cause the HKUST-1
to adsorb less CO2 than the dry sample. On the other hand, Liu et al. reported
that neither the HKUST-1 nor the Ni/DOBDC MOF can adsorb any signi-
cant amount of CO2 when water loadings are high, which means that the two
74 Industrial Catalysis and Separations: Innovations for Process Intensification
MOFs preferentially adsorb H2O over CO2 although water does not affect
CO2 adsorption in HKUST-1 and Ni/DOBDC as much as it does in tradi-
tional zeolites [44].
FIGURE 3.8 Eight metal-organic supramolecular isomers formed with the flexible
ligand TA with transition metals Zn, Cd, Co, Cu, and Mg. All the isomers showed the rigid
framework while structure A (TetZB) showed flexible framework towards guest sorption
[21, 22, 47]. Py = pyrazine, DB = 1,4-diazabicyclo[2.2.2]octane, and BP = bipyridine.
(The structures reproduced from Motkuri et al. in referenced RSC and ACS journals, with
permission).
A straightforward approach to reduce water adsorption in MOFs is to cre-
ate more hydrophobic MOF pores. This can be done by either postmodica-
tion synthesis of attaching hydrophobic moieties or generating MOFs with
hydrophobic ligands. Nune et al. reported the synthesis of thermally stable
Metal Organic Frameworks–Synthesis and Applications 75
metal-organic gels (MOGs) using near-supercritical processing conditions;
signicant CO2 uptake capabilities were observed for the rst time [49]. Yang
et al. synthesized some FMOFs in, which hydrogen atoms are replaced by
uorine atoms in all ligands [50]. FMOFs with uoro-lined or uoro-coated
channels or cavities have enhanced thermal stability and hydrophobicity com-
pared to their nonuorous counterparts. Farha et al. replaced the coordinated
solvent molecules in a three-dimensional MOF with various cavity modi-
ers, including pyridine and its derivatives [51]. The resulting tailored cavi-
ties show different adsorption properties, and this postsynthetic modication
method can be adopted to cover hydrophilic surfaces in some MOFs with
hydrophobic molecules, such as pyridine, to reduce H2O effects on CO2 ad-
sorption. An alternative engineering solution to mitigate water effects on CO2
adsorption in MOFs is to install a guard bed loaded with desiccants in front
of the main bed loaded with MOFs to remove most of the water and to take
advantage of the MOFs’ high CO2 capacities and selectivities.
3.3.5 WATER SORPTION STUDIES FOR ADSORPTION
COOLING APPLICATIONS
Technologies based on the sorption of water are detailed in the literature for
a number of applications, including adsorption-based heating and cooling,
air-conditioning systems, desiccant dehumidifiers, and freshwater production.
The main criterion is to find a best adsorbent–adsorbate (solid + vapor) pair
that effectively works on sorption principles. Because of water has the highest
mass-based evaporation enthalpy (2440 kJ/kg at 25°C), it would be a pre-
ferred adsorbate. Currently, inorganic zeolites and silicagel are widely used as
adsorbents for water sorption, but these sorbents suffer with water adsorption
at overly high relative pressures and also at high temperature for desorption,
resulting in a large footprint and high capital cost. An ideal working pair with
water would be a material that has high surface areas that can sorb huge water
molecules at low partial pressures and moderate sorbate–sorbent interactions
for low-temperature desorption in a cycle.
MOFs with extraordinary porosity in combination with tunable pore char-
acteristics make a suitable choice as adsorbent material. Janiak et al. dem-
onstrated a three-dimensional MOF, ISE-1, as an efcient adsorbent for heat
transformation cycles for refrigeration, heat pumping, and storage [52]. The
MOF ISE-1 has approximately 22 water molecules per formula unit with a
potential solvent volume of 1621 Å3 (52%) of the unit cell volume constructed
from benzene-1,3,5-tricarboxylate and 1,2-bis(1,2,4-triazol-4-yl)ethane. ISE-
76 Industrial Catalysis and Separations: Innovations for Process Intensification
1 demonstrated good water stability and achieved a loading spread of about
210 g/kg, which is larger than that of ve zeolites and silica gel that have
been used in commercial heat pump applications. The stability and loading
spread features may be due to the less hydrophilic nature of ISE-1 compared
to silica and zeolites, which release water molecules (desorption) at lower
temperature and lower partial pressures. The Janiak research group recently
also successfully used a thermally stable porous MIL-100 (Fe, Al) for wa-
ter adsorption/desorption of up to 0.75 g of water vapor per gram of MOF.
Because of very good cycle stability and suitable hydrophilicity, MIL-100
proved itself a valuable addition to the pool of sorption materials used in heat
pumps or sorption chillers [53]. Henninger et al. compared the water sorption
capacities of HKUST-1 with that of zeolites and silica gel and showed that
the highest water uptake for driving temperature of 95°C was mesoporous
aluminophosphate AlPO-18 (0.253 g/g), while at 140°C, HKUST-1 showed
a highest water uptake of 0.324 g/g [54]. They also showed the rst results on
integral heats of adsorption in the cycle. Ferey et al. have demonstrated that
hierarchically porous MIL-100 and MIL-101 with mesoporous cages behave
as advanced water adsorbents applicable to energy-efcient dehumidication
systems that showed superior performance over commercial adsorbents such
as zeolite NaX, SAPO-34, and silica gel [55]. On a large-scale evaluation,
a honeycomb rotor coated with these MOFs showed huge sorption uptakes,
even at 40°C, as well as high desorption rates below 80°C and almost two
times higher efciency in energy consumption for adsorptive dehumidica-
tion applications in commercial/residential buildings.
3.3.6 HARMFUL GAS REMOVAL
Harmful gasses in the environment pose a growing international security
threat. Effective capture of these gasses is critical to protect lives and the envi-
ronment [56]. Pioneering work has been published on the performance of six
different MOFs such as MOF-5, IRMOF-3, MOF-74, MOF-177, MOF-199,
and IRMOF-62 for dynamic adsorption capacities for eight harmful gasses:
sulfur dioxide, ammonia, chlorine, tetrahydrothiophene, benzene, dichloro-
methane, ethylene oxide, and carbon monoxide and compared the same with
BPL carbon respectively [56a]. The dynamic adsorption capacity of each
MOF for each gas has been determined using a kinetic breakthrough method
and compared with that of a Calgon BPL-activated carbon. Pore functional-
ity was found to play a dominant role in determining the dynamic adsorp-
tion performance of the MOFs. For example, MOF-199, which featured a
Metal Organic Frameworks–Synthesis and Applications 77
reactive functionality, outperformed BPL carbon in all cases except chlorine
while MOF-74 and IRMOF-3 outperformed in sorption of sulfur dioxide and
ammonia respectively [56a].
In addition, various MOFs and their composites have been used to remove
H2S, SO2, NH3, CO, NO, and benzene [57]. For efcient adsorptive removal,
both the pore structures and open metal sites and functional groups of adsor-
bents are of great importance. The open metal sites of MOFs have been reported
many times as the active sites for the removal of various toxic gasses. Petit and
Bandosz reported composites of MOFs and a graphitic compound (graphite or
graphite oxide, GO) for the adsorptive removal of NH3, H2S, and NO2 under
ambient conditions [58]. The open metal sites of porous MOFs coordinated
with the oxygen groups of GO led to the formation of a new pore space in the
interface, resulting in increases of more than 12% (for NH3), 50% (for H2S), and
4% (for NO2) in the adsorption capacity for the GO/Cu-BTC composite.
A uorinated MOF, FMOF-2 obtained from 2,2′-bis(4-carboxyphenyl)
hexauoropropane and zinc nitrate hexahydrate, was also reported for the ad-
sorptive removal of toxic acidic gasses [59]. FMOF-2 was quite stable for the
adsorption of SO2 and H2S and has adsorption capacities of 14.0% and 8.3%
for SO2 and H2S, respectively, at room temperature and 1 bar. Fernandez et
al. showed that FMOF-2 also manifests “breathing” behavior during SO2 and
CO2 adsorption, an uncommon feature in, which the framework structure ex-
pands and contracts upon adsorption and desorption of guest molecules [59,
60]. In addition to chemically toxic gas adsorption, the removal of radioactive
gases such as Xe and Kr using MOFs has also been investigated. Thallapally
et al. found that the Ni/DOBDC has a Xe adsorption capacity 10% higher than
that of activated carbon at room temperature and 1 bar [61]. Meek et al. found
that the uptakes of Xe and Kr can be affected by the linker polarizability from
their results on a series of mono halogenated isoreticular MOFs [62].
Several MOFs and a benchmark-activated carbon sample were studied as
adsorbents to separate low-concentration Xe and Kr from air [63]. Both the
Ni/DOBDC and the HKUST-1 can selectively adsorb Xe and Kr from air,
even at concentrations at the parts-per-million level. The Ni/DOBDC has a
Xe capacity of 9.3 mmol/kg when the concentration of Xe is 1000 ppm in
air. More important, the Ni/DOBDC is able to separate 400-ppm Xe from 40-
ppm Kr mixture in air with a Xe/Kr overall selectivity of 7.3. The high Xe/Kr
selectivity is due mainly to the strong unsaturated metal centers in the crystal
structures of the Ni/DOBDC. In addition, the uniform cylindrical pores in
the Ni/DOBDC are believed to be favorable to maximize the Xe/Kr selectiv-
ity [64]. Nevertheless, Fernandez et al. demonstrated for the rst time that a
78 Industrial Catalysis and Separations: Innovations for Process Intensification
MOF material can selectively capture and separate Kr from Kr/Xe mixtures
at moderate temperature via a temperature gating mechanism on a copper-
based uorinated MOF [65]. MOFs have shown great application potential in
the gas capture and separations area, which will help in evaluating the MOFs
for sorption, separation-related applications, and selective single-component
sorption/separation applications, respectively.
3.4 HETEROGENEOUS CATALYSIS
Although MOF-based catalysis was proposed and experimentally proven in
the last decade, substantial experimental exploration with considerable prom-
ise in catalysis has been reported very recently [66]. The use of MOFs as cata-
lysts for industrial processes may still be far into the future, but it is likely that
the growing opportunities demonstrated on the laboratory scale may trigger
interest in scaling up for industrial use. The low thermal and chemical stability
of MOFs, certainly limit their use in vapor-phase reactions (reactions carried
out above 300°C) in industrial processes such as oil refining or petrochemi-
cal processes in, which zeolites are currently used. However, in liquid-phase
reactions, MOFs can compete with or even outperform the existing zeolites,
particularly the reactions carried out under mild conditions [67]. The main
advantage of MOFs is their versatility in chemical composition, organic and
inorganic building units, and the bifunctional metal/acid sites for insertion
using isoreticular chemistry. MOFs can complement zeolites to perform the
reactions at lower temperature s. As heterogeneous catalysts, the three known
possibilities for building catalytic activity are (1) framework activity (2) en-
capsulation of active species, and (3) postsynthesis modification.
3.4.1 FRAMEWORK ACTIVITY
One of the most widely explored strategies for configuring MOFs as hetero-
geneous catalysts is to take advantage of the exchangeable coordination posi-
tions around the metal ions. Another methodology would be to introduce the
organic linker with acidic or basic groups as the active sites. Recently, the
structural defects around the metal nodes also were explored for catalytic ap-
plications [68]. The framework activity is further divided into acidic and basic
sites, as discussed below.
3.4.1.1 LEWIS ACID CATALYSIS
The porous material [Cu3(btc)2(H2O)3], also known as HKUST-1, contains
large cavities. The copper metal is coordinated to water molecules that can
Metal Organic Frameworks–Synthesis and Applications 79
be removed easily under heat treatment, leaving open metal centers (Cu(II)
sites act as Lewis acidic centers) [69]. Jacobs et al. investigated the behav-
ior of HKUST-1 as a Lewis acidic catalyst by examining three sets of re-
actions, including isomerization of (i) alpha-pinene oxide (ii) rearrange-
ment of alpha-bromoacetal, and (iii) cyclization of citronellal. Based on the
experimental data and product selectivities, they concluded that HKUST-1
works primarily as a Lewis acid catalyst [70]. Kaskel et al. evaluated the
Lewis acidity of both HKUST-1 and large-pore MIL-101 having the formu-
la [Cr3F(H2O)2O(bdc)3] on cyanosilyzation of benzaldehyde and concluded
that the Cr(III) sites in MIL-101 showed greater activity than Cu(II) sites in
HKUST-1. Moreover, the catalytic sites in MIL-101 are immune to the un-
wanted reduction of benzaldehyde during the reaction [71]. Long et al. used
a manganese-based MOF with the formula Mn3[(Mn4Cl)3BTT8(CH3OH)10]2
(H3BTT = 1,3,5-benzenetris(tetrazol-5-yl) having a three-dimensional pore
structure with a pore diameter of 10 Å having two Mn+2 sites. The MOF ac-
tively engaged in Lewis acid catalytic conversion of selected aldehydes and
ketones with cyanotrimethylsilane to the corresponding cyanosilylated prod-
ucts with conversion yields above 90%, one of the highest yields reported in
metal-organic frameworks (Fig. 3.9). The same MOF also was shown to cata-
lyze the Mukiyama aldol reaction, which generally requires stronger Lewis
acidic sites than cyanosilyzation reactions [72]. In another report, Gandara et
al. documented highly Lewis acidic In (III) sites in a two-dimensional MOF
that catalyzed acetalization of benzaldehyde with trimethyl orthoformate [73].
FIGURE 3.9 (Left) A portion of the crystal structure of Mn3[(Mn4Cl)3BTT8(CH3OH)10]2
(H3BTT = 1,3,5-benzenetris(tetrazol-5-yl) showing the two different types of Mn+2
sites exposed in 10-Å-wide channels. Site I is five-coordinate, while site II is only two-
coordinate; the separation between them is 3.4 Å. (Right) Lewis acid catalyzed conversion
of selected aldehydes and ketones with cyanotrimethylsilane to the corresponding
cyanosilylated products. Orange, green, gray, and blue spheres represent Mn, Cl, C, and
N atoms, respectively [72]. Reproduced from Horike, S. et al., J Am Chem Soc 130, 5854
(2008), with permission).
80 Industrial Catalysis and Separations: Innovations for Process Intensification
3.4.1.2 BRØNSTED ACID CATALYSIS
Ferey et al. reported the Brønsted acid catalysis reaction on Friedal-Crafts
benzylation using two different MIL-100 cations, Fe and Cr. Although the
structure is identical [M3OF0.85(OH)0.15(H2O)2(btc)2] in both cases, the Fe-based
MIL-100 showed superior performance over that the Cr-based MIL-100 and
even surpassed the reported HY and HBEA zeolites [74]. In MIL-100(Cr+3),
the Cr-OH sites showed medium Brønsted acidic strength as confirmed by
low-temperature CO chemisorption studies [75].
3.4.2 ORGANIC OR PSEUDO-ORGANIC LINKERS AS ACTIVE
SITES
The metal complexes with functionalized ligands such as porphyrin building
blocks as organic linkers reported the catalytic active in oxidation reactions.
The porphyrinocarboxylate frameworks constructed from both Mn(III) and
Zn (II) successfully showed for olefin epoxidation reactions [76]. Here, the
metal coordinated to the porphyrin nitrogens performed as active sites when
compared to inorganic nodes of the MOF. Amino-functionalized MOFs such
as IRMOF-3 and [ZnF(Am2Taz)] synthesized from corresponding 2-ami-
noterepthalic acid and 3,5-diamino-1,2,4-triazole (Am2Taz) with zinc has
been successfully used as base catalysts in the Knoevenagel reaction. Both
MOFs showed the activity toward Aza-Michael condensation reactions at
25°C with a turnover number (TON) of 1.4 h–1 and 0.15 h–1 and fatty methyl
ester transesterification with a TON of 3.3 h–1 and 0.3 h–1 at 130°C, respec-
tively [77].
3.4.3 CATALYSIS WITH TAILORED MICROENVIRONMENT
Prepared post synthetic modifications of MOFs as described below have led
to the creation of a new microenvironment inside their porous network for
accelerating the rates of specific reactions and enhancing the selectivity to the
desired product.
3.4.3.1 ENCAPSULATION OF ACTIVE SPECIES
The encapsulation of active species such as metal complexes or metal
nanoparticles (MNPs) inside the porous MOFs is of growing interest in the
catalysis community because MOFs can provide a large pore volume that can
accommodate guest molecules. Here the MOF porous framework is used as a
Metal Organic Frameworks–Synthesis and Applications 81
support or host for the catalytic species, which was positioned in the cavity by
noncovalent interactions using a ship-in-a-bottle or bottle-around-the-ship ap-
proach [68, 78]. The encapsulation of MNPs inside MOFs can be achieved by
a stepwise process of particle infiltration followed by decomposition. Here,
the pore size, shape, and channel structure govern the size and shape of the
nanoparticle. It has also been reported that simple grinding of the organome-
tallic complex and MOF could result in the encapsulation inside the MOF for
aerobic oxidation of alcohols [79].
BASF is industrially producing MOF-5 mainly for encapsulating plati-
num, palladium, copper, and gold nanoparticles for catalysis applications.
The Pd@MOF-5 is used as a heterogeneous catalyst in hydrogenation reac-
tions, specically hydrogenation of cyclooctene, while Cu@MOF-5 is used in
methanol production from syngas [80].
Another example is encapsulation of metal porphyrins into the MOF cavi-
ties to be used for oxidation of cyclohexane to corresponding cyclohexanol and
cyclohexanone. The reactivity was comparable to that of reactions carried out
using a free metal porphyrin complex [81]. The mesoporous MOFs, chromium
(III) terephthalate Cr-MIL-101 [82], and iron anologue Fe-MIL-101 [83] have a
rigid zeotype crystal structure with quasi-spherical cages of 2.9 and 3.4 nm and
have been demonstrated to have very good encapsulation capabilities. Because
of their high thermal stabilities, huge surface areas, and large pore volumes,
these materials played a greater role as host materials for active species. Khold-
eeva et al. showed the encapsulation of polyoxometalates [PW11TiO39]5– and
[PW11CoO39]5–, which accelerates the oxidation of alpha-pinene to correspond-
ing alcohol and ketones using hydrogen peroxide (Fig. 3.10) [84].
FIGURE 3.10 The catalytic activities of polyoxotungstate-encapsulated MIL-101
(PWx/MIL101) for selective oxidation of alkenes with aqueous hydrogen peroxide [85].
(Reproduced from N. V. Maksimchuk, N.V. et al., Inorg Chem 49, 2920–2930 (2010), with
permission).
82 Industrial Catalysis and Separations: Innovations for Process Intensification
3.4.3.2 POST SYNTHETIC MODIFICATION
A new synthetic strategy of incorporating active functional groups using a
postsynthetic modification approach has been proposed recently as an alter-
native to presynthetic approaches discussed above. Cohen et al. used this ap-
proach to modify the amino functional groups in IRMOF-3 with alkyl anhy-
drides [86]. The amino groups within MOFs have been synthetically modified
using carboxylic acid or isocyanates to introduce various functions. Similarly,
Ferey et al. introduced thermally stable amino groups into MIL-101 frame-
work to the unsaturated Cr (III) coordination centers that open new oppor-
tunities for introduction of organic functional groups [87]. He showed the
successful introduction of ethylenediamnie (EDTA) (ED@MIL-101) and
diethylenetriamine (APS@MIL-101) into the MIL-101 framework. The free
terminal NH2 groups hanging in the cavities showed Bronsted basicity in the
Knoevenagel condensation reactions with 99.3% selectivity. Using a similar
approach, Corma et al. introduced a Au(III) complex with a Schiff base, and
the material showed superior activity and selectivity in domino-coupling reac-
tions [88]. The catalyst is completely recyclable and showed higher catalytic
activity when compared to gold-based homogeneous and supported catalysts.
Eddaoudi et al. demonstrated a versatile platform by encapsulation of free-
base porphyrin in a large pore of an In-HImDC-based rho-ZMOF, which post-
synthetically modified by various transition metals to produce a wide array of
encapsulated metalloporphyrins in MOFs and successfully tested for oxida-
tion reactions (Fig. 3.11) [89].
FIGURE 3.11 (a) Eight-coordinate molecular building blocks that could be represented
as tetrahedral building units (b) [H2TMPyP]4+ porphyrin (c) crystal structure of rho-ZMOF,
schematic representation of [H2TMPyP]4+ porphyrin ring enclosed in rho-ZMOF α-cage
(drawn to scale) (right) catalytic oxidation of cyclohexane at 65°C (right) [89]. (Reproduced
with permission from Alkordi, M.H. et al., J Am Chem Soc 130, 12639 (2008)).
Metal Organic Frameworks–Synthesis and Applications 83
3.4.4 MOFS FOR SENSING APPLICATIONS
A sensor is a device that responds to a physical or chemical stimulus (e.g.,
heat, light, sound, pressure, chemical vapor) and transmits a resulting im-
pulse. As described earlier MOFs have high surface area and porosity, struc-
tural diversity, flexibility, and physical and chemical tailorability that render
them excellent candidates for sensing applications. In MOF-based sensors, the
MOF responds to external stimulation such as gravimetric, mechanical, opti-
cal or environmental changes by manifesting a change in structure and prop-
erties or by showing a guest-dependent response. Although the latter is the
most common case, an example of structural change was observed in MOF-5.
The framework volume of MOF-5 changes with external temperature, show-
ing a linear negative thermal expansion (–16 x 10–6 K–1) between 4 K and
600 K. This expansion is due to the large-amplitude transverse vibration of
the carboxylate groups. A large number of examples of MOF-based sensors
show the guest-dependent response [90]. MOF samples have displayed opti-
cal, magnetic, or electronic responses based on the interactions between the
framework structure and a guest molecule, which opens up a large number of
applications of MOFs as sensing devices [91].
Depending on their response mechanism or route, MOF sensors can be
classied into three groups: piezoresistive response, mass response, and opti-
cal response. The rst relies on the stress induced by a guest molecule and
takes advantage of the structural exibility of MOF materials. Mass response
takes advantage of the large surface area and available volume capacity of the
pores to produce an important change in host mass when capturing a given an-
alyte. Optical responses, including photoluminescence quenching, emission
wavelength shifts, or simple color changes due to a guest-induced change in
MOF optical absorption spectrum, are reported on extensively in the litera-
ture.
3.4.5 MOF SENSORS BASED ON PIEZORESISTIVITY
Sensors based on piezoresistive MOFs are promising sensing devices for de-
tecting gas molecules and small organic molecules. One of the attractive fea-
tures of MOF materials is the structural flexibility. A number of MOF materials
exhibit expansion and contraction of the crystal lattice induced by guest–host
interactions. This property can be evidenced by “gating effects” and “breath-
ing effects” due to expansion and contraction of the framework upon incorpo-
ration of guest molecules. The unit cell dimensions of some MOFs can vary
84 Industrial Catalysis and Separations: Innovations for Process Intensification
by as much as 10% when molecules are adsorbed within their pores [92].
Therefore, transduction mechanisms in, which distortions in a MOF thin film
creates stress at the interface with a second material can be used to recognize
and measure a number of analytes. By depositing a MOF film on the surface
of a static microcantilever, stress at the cantilever surface results in bending
that can be detected by means of a built-in piezoresistive sensor [93]. In this
fashion, MOFs can show effective recognition chemistries for a variety of
gases and vapors provided they interact with the MOF pore.
An example is HKUST-1, which, although small, shows adsorption-in-
duced distortions when water, methanol, or ethanol are reversibly adsorbed.
Allendorf et al. showed this property when a thin lm of HKUST-1 was inte-
grated with a microcantiveler surface [94]. The time-dependent responses to
H2O are shown in Fig. 3.12. It was also demonstrated that the sensor responds
to CO2 when the MOF layer is dehydrated. The stress in this case was due to
the coordination of CO2 with unsaturated copper sites in HUKST-1. Because
the framework distortions were quite small, the group concluded that higher
sensitivities can be achieved using more exible MOFs.
FIGURE 3.12 Temporal response of the cantilever piezoresistive sensor to water vapor
diluted in N2 (room temperature, 1 atm) [94]. (Reproduced with permission Allendorf,
M.D. et al., J Am Chem Soc 130, 14404 (2008), with permission).
3.4.6 GRAVIMETRIC-BASED MOF SENSORS
Thin films of MOF materials can also be applied on quartz crystal micro-
balances (QCMs). In this fashion, it is possible to take advantage of the
extremely high sensitivity to mass changes on the QCM surface when an ana-
lyte interacts with the cavities of the MOF. Detection limits on the order of
Metal Organic Frameworks–Synthesis and Applications 85
the nanogram scale have been demonstrated [95]. MOF thin films grown on
a QCM substrate can be widely adopted as a device to probe mass changes
due to gas or vapor adsorption/desorption in the highly porous sorbent mate-
rial. Liu et al. successfully used a stepwise liquid-phase epitaxial method to
synthesize two homochiral thin films of [Zn2(±)cam2dabco]n (where ±cam =
(1R,3S)—(±)-camphoric acid, and dabco = 1,4-diazabicyclo(2.2.2)octane)
on self-assembled monolayers (SAMs) functionalized QCM substrates [96].
In another example, coating a QCM substrate with HKUST-1 showed to de-
tect humidity variations via adsorption/desorption of water molecules on the
sorbent film [97].
3.4.7 SURFACE ACOUSTIC WAVE-BASED MOF SENSORS:
Surface acoustic wave (SAW) sensors [93] are robust devices extensively em-
ployed for chemical sensing. Depositing MOF materials on the sensor surface
can then be useful for detecting and measuring vapors and gases. A frequency
shift of acoustic waves traveling parallel to the MOF surface, which are gen-
erated by an oscillator can be used to detect gas or vapor sorption. By grow-
ing a HKUST-1 film using a layer-by-layer (LBL) method on the quartz of
96.5-MHz devices, Robinson and collaborators reported humidity detection
using SAWs (Fig. 3.13) [98]. Fast and reproducible detection on frost points
(FP) as high as 10°C and as low as −70°C (2.6 ppmv and 12,300 ppmv at an
atmospheric pressure of 625 Torr, respectively) can be obtained. In addition,
the group demonstrated three orders of magnitude better response to humidity
using HKUST-1 as compared to the same coating on QCMs [97, 99]. The in-
fluence of film thickness on the SAW sensor response also was reported. Op-
timum coupling occurred between the MOF film and SAW surface; hence, the
highest response of the HKUST-1 SAW sensor to water vapor was obtained
for LBL coatings between 40 and 50 cycles (150-to 180-nm film thickness).
FIGURE 3.13 Humidity detection (left) and a molecular model of Cu3(BTC)2 (HKUST-1)
SAWs with water molecules (right) [98]. (Reproduced from Robinson, A.L. et al., Anal
Chem 84, 7043 (2012), with permission).
86 Industrial Catalysis and Separations: Innovations for Process Intensification
3.5 SENSORS BASED ON MOF OPTICAL PROPERTIES
3.5.1 PHOTOLUMINESCENCE (PL)-BASED MOF SENSORS
PL-MOFs are the most widely explored type of MOF sensing devices. The
popularity of luminescence over other transduction mechanisms are a con-
sequence of several key elements, including very high detection limits that
can reach molecular level. In addition, there is no need for film fabrication or
other processing, so conventional solvothermal synthesis is used to produce
PL-MOFs. A very important advantage over other PL sensors is the porosity
of MOFs. In addition, the possibility of adjusting MOF sorption properties of-
fers a high degree of molecular specificity. As an example, Li et al. prepared a
luminescent Ln-based MOF displaying efficient turn-on triggered by solvent
vapors, showing good selectivity for DMF vapor [100].
In another example, Lanet al. reported the detection of trace nitroaro-
matic explosives (DNT, 2,4-dinitrotoluene and DMNB, 2,3-dimethyl-2,3-
dinitrobutane) in the vapor phase using a highly luminescent MOF material
[Zn2(bpdc)2(bpee)]n (bpdc = 4,40-biphenyldicarboxylate; bpee = 1,2-bipyri-
dylethene) thin lm where they described the sensing process as a redox
quenching mechanism [101].
3.5.2 SOLVATOCHROMISM/VAPOCHROMISM
A visible change in a material’s color is one of the simplest means of trans-
ducing a sensing signal. Solvatochromism and vapochromism refers to a large
shift in the absorption spectrum of a material in response to a change in the
identity of the solvent or vapor. As in any other optical active material, MOFs
constructed from ligands that are chromosphors should behave in a similar
fashion in the presence of vapors. The electronic transition responsible for
the coloration entails charge transfer (i.e., a change in dipole moment upon
excitation from the ground electronic state to the excited electronic state of
the chromophoric component of the material) in a MOF material, the organic
linker. If the ground state has the larger dipole moment, hypsochromic shifts
(blue shifts) occur with increasing solvent polarity. On the other hand, if the
excited state possesses a larger dipole moment than the ground state, it is pref-
erentially stabilized by polar solvents, and bathochromic shifts (red shifts) are
observed with increasing solvent polarity.
A square shaped nanotubular MOF with the stoichiometry [[(WS4Cu4)-
I2(dptz)3]·DMF]n (124·DMF) in, which each WS4
2− anion chelates four Cu(I)
cations was synthesized by Lu et al. and its optical properties studied. The
Metal Organic Frameworks–Synthesis and Applications 87
sorbent material shows individual WS4Cu4
2+ units that are paired up by pyr-
azines that exhibit π−π interactions at an interplanar distance of 3.578(2)
Å. Also, each Cu(I) center is tetrahedrally coordinated. The 5.4-× 5.3-Å
nanotubes along the c-axis contain DMF guests and, once DMF is removed,
immersion of the MOF into various organic solvents manifested signicant
color changes. The group described a negative solvatochromic effect with a
solvent-induced absorption band shift of 245 nm between CH3CN and CHCl3
(Fig. 3.14) [102].
FIGURE 3.14 (Left) The perspective view of the nanotubular structure of [(WS4Cu4)
I2(dptz)3]n. Right: The UV−vis spectra and photograph of the inclusion compounds 1
solvent [102]. Reproduced from Lu, Z.Z. et al., J Am Chem Soc 133, 4172 (2011), with
permission).
As another example, Long et al. showed that exposing CO2
+-based MOFs
to various vapors could shift the optical absorption across the visible region
[103]. The explanation was a change in coordination environment from the
as-synthesized octahedral to a tetrahedral geometry. Lee et al. proposed a
similar color change mechanism on a MOF material that detects chloride ions
derived from chlorine-containing vapors or gasses. The MOF contains CO2
+
nodes coordinated to 1,2,4,5-tetra(2H-tetrazole-5-yl)-benzene (TTB) struts as
well as Br-anions [104]. The as-synthesized material is characterized by a
visible absorption peak at 475 nm, suggesting octahedral coordination of the
CO2
+ centers. Interestingly, exposure of the MOF material to chlorine-con-
taining gasses, including HCl, SOCl2 (COCl)2, and COCl2 (phosgene) show
color changes from red to blue [104]. The blue color, which is due to a new
88 Industrial Catalysis and Separations: Innovations for Process Intensification
absorption at 670 nm was attributed to tetrahedrally coordinated CO2
+. To
explain the change in metal coordination geometry, it was hypothesized that
the Br-is replaced with Cl-originating from the reactive gasses.
3.6 MOFS FOR OIL SPILL CLEANUP
The world energy dependence on fossil energy continues to increase. Petro-
leum is the largest source of primary energy production in the world. The
world petroleum consumption of 87,135,100 barrels/day (2010) poses a
daunting task to use aggressive methods for producing and transporting the
crude oil by various means including pipeline, ship, or barge to the rest of the
world from production sites [105]. Production, transportation, and storage of
oil present the risk of oil spill.
The rst commercial oil spill dates back to 1967 in the United Kingdom;
the most recent but not the worst (BP Deepwater Horizon Oil spill in the Gulf
of Mexico) occurred in 2010, and its cleanup cost is estimated to be over
$20 billion [106]. Oil spills not only cause nancial burden but also pose an
environmental damage risk. Oil is highly hydrophobic, does not mix with
water, and forms a thick slick or droplets that oat on water. With time the
oil eventually becomes spreadout. Although no two oil spills are considered
the same due to their location, weather conditions, and oil type, current oil
spill cleanup methods include three main methods of response (i) booming
(ii) skimming, and (iii) centrifuge in, which physical collection is routine (see
Ref. [1], of Chapter 8). Use of dispersants that break the oil to speed up the
natural biodegradation and burning of oil are considered as potential alterna-
tives. However, the use of dispersants and burning often results in adverse
effects on the environment. One another approach is to use the porous sorbent
that can effectively absorb the oil [107, 108]. A porous sorbent for oil spill
cleanup should have the following characteristic properties: (i) high rate of
adsorption/absorption (ii) oil retention, and (iii) ease of application. Currently
used sorbents include sand, organic clays, and cotton bers. These adsorbents
are cheap and easy to apply; however, they suffer from their afnity toward
water, limiting their performance during cleanup.
In general, to be effective for oil cleanup, a porous material should be
hydrophobic and olephilic. MOFs with excellent porosities and hydrophobic-
ity are likely to be efcient hydrocarbon absorbers at low to high concentra-
tions. Focusing on the potential industrial importance, Motkuri et al. recently
reported for the rst time breathing and adsorption of hydrocarbons using a
novel exible MOF (TetZB) [22]. They observed framework expansion and
Metal Organic Frameworks–Synthesis and Applications 89
contraction upon uptake and release of hydrocarbons. Although framework
expansion properties in exible MOFs are not of interest for cleanup appli-
cations, they have potential uses in sensing and other applications. In gen-
eral, typical hydrocarbon sorption experiments are conducted on gravimet-
ric analyzers (Intelligent Gravimetric Analyzer (IGA), Hiden Instruments or
duPont Model 990 TGA). Variable-temperature powder x-ray diffraction of
TetZB has revealed that there is framework distortion (probably by contrac-
tion) by removing the coordinated solvent on metals. Figure 3.15 illustrates
the adsorption and desorption of hydrocarbons on TetZBat room temperature.
TetZB exhibits an open structure when the solvent is removed after activa-
tion. Upon exposure to the hydrocarbons, the narrow pores in the framework
are saturated, while at high relative pressures, the framework undergoes ex-
pansion by opening access to the large pores [22].
FIGURE 3.15 Structure of flexible MOF TetZB (left) and their hydrocarbon uptake
measurements (right) [22]. (Reproduced from Motkuri, R.K. et al., Chem Commun 47,
7077 (2011), with permission).
While addressing hydrocarbon sorption properties of MOFs, there has
been intense debate on the commensurate adsorption of hydrocarbons using
MOFs; the mean number of adsorbed hydrocarbon molecules relates to the
symmetry of the framework topology, which is nearly impossible in routinely
studied/used zeolites [109]. Jing Li et al. reported that adsorption capacity
and location of molecules are specically controlled by cavity size, shape,
symmetry, and channel features such as channel segment [109]. Three differ-
ent adsorption mechanisms have been offered for the sorption and separation
of hydrocarbons: (1) equilibrium (2) steric, and (3) kinetic. In some com-
plex materials, it is possible that one or more mechanisms may operate at the
same time. The equilibrium sorption and separation mechanism relies on
90 Industrial Catalysis and Separations: Innovations for Process Intensification
the difference in the relative quantities of various hydrocarbons; the size and
shape of the pores in the adsorbate contribute to the steric mechanism, and the
difference in the rate of adsorption of different hydrocarbons contribute to the
kinetic mechanism [109].
Generally, MOFs have high afnity toward water, limiting their applica-
tions; however, the recent Deepwater Horizon oil spill devastation has in-
creased awareness of the need for new adsorbents. Signicant efforts have
been devoted to develop new hydrophobic MOFs with high thermal stability,
high selectivity, and fast regenerability. A prominent approach to circumvent
this problem is the utilization of postsynthetic modication (PCM) of MOFs to
produce moisture resistant and superhydrophobic MOFs [110]. This approach
garnered signicant attention because it is an important tool to chemically
modify or introduce various functional groups to produce new MOFs with
completely new chemical and physical properties. It means chemical modi-
cation is performed on the nal MOF materials rather than in the precursor
used for MOF construction. This approach is particularly attractive because
(1) the use of solvothermal conditions for MOF synthesis limits the presence
of functional groups on the ligand and (2) unlike any inorganic porous mate-
rials, MOFs have organic component that can be functionally modied, ren-
dering change in physical and chemical properties [110]. To circumvent the
posed by water-unstable MOFs, Yang et al. recently developed stable MOFs
using novel ligands that contain peruorinated groups (Fig. 3.16) [111]. These
new MOFs (FMOFs) are hydrogen free and uorine rich, not only bestowing
greater stability but also offering favorable hydrocarbon sorption properties
while inducing hydrophobic characteristics. FMOF-1 is the rst known MOF
with an exceptional hydrophobic nature while offering afnity for typical aro-
matic hydrocarbons and aliphatic oil components. To the best of our knowl-
edge, this nding is denitely a signicant scientic achievement, considering
the fact that many known MOFs are not stable in the presence of water or they
uptake very large quantities of water that compromise the uptake of hydro-
carbons during cleanup. Water adsorption isotherms clearly illustrate that the
FMOF-1 is signicantly hydrophobic, while routinely studied/used inorganic
porous materials such as activated carbon and zeolite-5A are hydrophilic, ab-
sorbing water even at very low relative humidity [111]. FMOF-1 is not only
hydrophobic but also very stable under aggressive conditions. FMOF-1 has
been soaked in distiled water for several days; powder X-ray diffraction im-
ages of soaked materials resemble the untreated material, indicating the ex-
ceptional stability.
Metal Organic Frameworks–Synthesis and Applications 91
FIGURE 3.16 (Left) Building blocks of FMOF-1 (middle) water sorption studies in
FMOF-1, BPC-Carbon and Zeolite-5A; (right) water and hydrocarbon sorption studies in
FMOF-1 [111]. (Reproduced from Yang, C. et al., J Am Chem Soc 133, 18094 (2011), with
permission).
92 Industrial Catalysis and Separations: Innovations for Process Intensification
In addition to oil cleanup, MOFs have been extensively investigated for
hydrocarbon separation applications due to the fact that currently available
separation strategies are energy-intensive processes and are contributing to
the largest segment of industrial production costs [112]. A recent report using
the exible feature that we reported earlier unveils the exceptional selectivity
of exible MOFs based on the differences in their gate opening pressure [22,
113]. The selectivity depends on the chain length of the hydrocarbon and its
specic interaction with framework [113].
3.7 CONCLUSIONS
Although the synthesis of MOFs was initially established for making new
compounds with interesting properties, the field has matured well and is slow-
ly widening its scope. The surface areas and pore properties of MOFs demon-
strate that these materials have large applicability in both fundamental and in-
dustrial arenas. The chemical versatility, pore tunability and tailoring of active
sites will lead to a broader range of applications. Although the conventional
hydro/solvothermal synthesis was well used in MOF synthesis, alternative
and new methods such as electro, mechano, sonochemical, and microwave-
assisted synthesis are just emerging in this area. However, greater care must
be taken while considering the crucial points in the synthesis, specifically the
reaction conditions and input energy requirements. Recent developments of
high-throughput methods may provide the tools needed in establishing the
standard synthetic protocol for each MOF, but establishing these protocols
takes time. With the developments in myriads of MOFs and their applications,
it is time to take them from the laboratory to the industrial world and focus on
applying this knowledge to industrial scenarios for real-world applications.
Although there is some progress in scale-up of MOFs for industrial produc-
tion and use, much still needs to be learned from zeolite chemistry for com-
mercial success.
In gas capture and separation-related topics, the current status of MOF
research is focused primarily on exploring new materials for single sorption
properties that cannot be viewed as new sorbent materials for practical sepa-
ration applications. The breakthrough experiments with mixed gases (includ-
ing vapors) will be a valuable means for better understanding the materials
for long-term potential and applicability. In the heterogeneous catalysis area,
MOFs still are unable to compete with zeolites in high-temperature reactions,
but these materials are capable of replacing zeolites in catalysis reactions that
use mild reaction conditions in ne chemical and enantio-selective synthesis.
Metal Organic Frameworks–Synthesis and Applications 93
MOF materials can respond to external stimulation in a number of ways
including changes in structure, as well as in their optical or mechanical prop-
erties, which can be employed for the fabrication of a number of sensing de-
vices. Although some examples have been shown, their application in this area
is still in the early stage. Because MOFs offer very attractive properties, there
is a great potential for developing novel MOF-based sensors whose detection
limits will depend on the type of stimulation the material is responsive to. An-
other attractive feature of MOFs is that their physical and chemical properties
can be altered by postsynthesis modication; a technique not possible in any
other class of porous inorganic materials thus will help in developing suitable
materials for required applications.
Most of the current technological options for the recovery and separa-
tion of industrially important gases including CO2 and H2 entail the forma-
tion of molecular complexes, through physical and/or chemical interactions,
that must then be reversed through signicant energy input, which is wasteful
both thermodynamically and dynamically. While continued improvement in
the above technologies can be expected with further R & D, the new options
with novel materials like MOFs and their analogs could provide signicant
breakthroughs in intensication of both chemical processes and separations
by tuning to vary cavity size and functionalizations to create a more effective
microenvironment inside their porous structure and to enhance accessibility
and interactions with molecules conned within the tailored environment.
As there is a great deal of exibility in tailoring their structural framework
and functionalizations, it is likely that MOFs will play much more important
role in the future to achieve higher level of process intensication leading to
greener process.
KEYWORDS
Adsorption chiller
CO2 capture
Gas capture and separation
Heterogeneous catalysis
Metal-organic frameworks
Oil spill cleanup
Sensing applications, sensors
Water sorption
94 Industrial Catalysis and Separations: Innovations for Process Intensification
ACKNOWLEDGMENTS
R. K. Motkuri would like to thank Dr. K.V. Raghavan for the support over the
years. This review was not possible without support from U.S. Department of
Energy’s office of Advanced Research Projects Agency for Energy (ARPA-e)
under Building Energy Efficiency through Innovative Thermodevices (BEET-
IT) program. Pacific North-west National Laboratory is a multiprogramming
laboratory operated by Battelle Memorial Institute for the U.S. Department of
Energy under contract DE-AC05–76RL01830.
REFERENCES
1. Sing, K. S. W., Everett, D. H., Haul, R. A. W., Moscou, L., Pierotti, R. A., Rouquerol,
J., and Siemieniewska, T. Reporting physisorption data for gas solid systems with
special reference to the determination of surface-area and porosity (Recommendations
1984). Pure Appl Chem 57, 603–619 (1985).
2. (a) Corma, A. From microporous to mesoporous molecular sieve materials and their
use in catalysis. Chem Rev 97, 2373–2419 (1997); (b) Erdem-Senatalar, A., Bergen-
dahl, J. A., Giaya, A., and Thompson, R. W. Adsorption of methyl tertiary butyl ether
on hydrophobic molecular sieves. Environ Eng Sci 21, 722–729 (2004).
3. (a) Kitagawa, S., Kitaura, R., and Noro, S. Functional porous coordination polymers.
Angew Chem Int Edit 43, 2334–2375 (2004); (b) Rowsell, J. L. C. and Yaghi, O. M.
Metal-organic frameworks: a new class of porous materials. Micropor Mesopor Mat
73, 3–14 (2004); (c) Ferey, G. Hybrid porous solids: past, present, future. Chem Soc
Rev 37, 191– 214 (2008).
4. (a) Tranchemontagne, D. J., Mendoza-Cortes, J. L., O’Keeffe, M., and Yaghi, O. M.,
Secondary building units, nets and bonding in the chemistry of metal-organic frame-
works. Chem Soc Rev 38, 1257–1283 (2009); (b) Natarajan, S. and Mahata, P. Metal-
organic framework structures-how closely are they related to classical inorganic struc-
tures? Chem Soc Rev 38, 2304–2318 (2009).
5. (a) Furukawa, H., Ko, N., Go, Y. B., Aratani, N., Choi, S. B., Choi, E., Yazaydin,
A. O., Snurr, R. Q., O’Keeffe, M., Kim, J., and Yaghi, O. M. Ultrahigh porosity in
metal-organic frameworks. Science 329, 424–428 (2010); (b) Deng, H. X., Grunder,
S., Cordova, K. E., Valente, C., Furukawa, H., Hmadeh, M., Gandara, F., Whalley, A.
C., Liu, Z., Asahina, S., Kazumori, H., O’Keeffe, M., Terasaki, O., Stoddart, J. F., and
Yaghi, O. M. Large-pore apertures in a series of metal-organic frameworks. Science
336, 1018–1023 (2012).
6. Stock, N. and Biswas, S. Synthesis of metal-organic frameworks (MOFs): Routes to
various MOF topologies, morphologies, and composites. Chem Rev 112, 933–969
(2012).
7. Li, J. R., Sculley, J., and Zhou, H. C. Metal-organic frameworks for separations. Chem
Rev 112, 869–932 (2012).
8. Knobloch, F. W. and Rauscher, W. H. Coordination polymers of Copper(II) prepared
at liquid-liquid interfaces. J Polym Sci 38, 261–262 (1959).
Metal Organic Frameworks–Synthesis and Applications 95
9. Johnson, J. W., Jacobson, A. J., Butler, W. M., Rosenthal, S. E., Brody, J. F., and
Lewandowski, J. T. Molecular recognition of alcohols by layered compounds with
alternating organic and inorganic layers. J Am Chem Soc 111, 381–383 (1989).
10. Drumel, S., Janvier, P., Barboux, P., Bujolidoeuff, M., and Bujoli, B. Synthesis, struc-
ture, and reactivity of some functionalized zinc and Copper(II) phosphonates. Inorg
Chem 34, 148–156 (1995).
11. Riou, D., Roubeau, O., and Ferey, G. Composite microporous compounds. Part I: Syn-
thesis and structure determination of two new vanadium alkyldiphosphonates (MIL-2
and MIL-3) with three-dimensional open frameworks. Micropor Mesopor Mat 23, 23–
31(1998).
12. Li, H., Eddaoudi, M., O’Keeffe, M., and Yaghi, O. M. Design and synthesis of an ex-
ceptionally stable and highly porous metal-organic framework. Nature 402, 276–279
(1999).
13. Klinowski, J., Paz, F. A. A., Silva, P., and Rocha, J. Microwave-assisted synthesis of
metal-organic frameworks. Dalton T 40, 321–330 (2011).
14. Li, C. P., Tian, Y. L., and Guo, Y. M. Solvent-regulated assembly of 1-D and 2-D
Zn(II) coordination polymers with tetrabromoterephthalate. Inorg Chem Commun 11,
1405–1408 (2008).
15. Li, Z. Q., Qiu, L. G., Xu, T., Wu, Y., Wang, W., Wu, Z. Y., and Jiang, X. Ultrasonic
synthesis of the microporous metal-organic framework Cu(3)(BTC)(2) at ambient
temperature and pressure: An efficient and environmentally friendly method. Mater
Lett 63, 78–80 (2009).
16. Klimakow, M., Klobes, P., Thunemann, A. F., Rademann, K., and Emmerling, F.
Mechanochemical synthesis of metal-organic frameworks: A fast and facile approach
toward quantitative yields and high specific surface areas. Chem Mater 22, 5216–5221
(2010).
17. Joaristi, A. M., Juan-Alcaniz, J., Serra-Crespo, P., Kapteijn, F., and Gascon, J. Elec-
trochemical synthesis of some archetypical Zn2+, Cu2+, and Al3+ metal organic frame-
works. Cryst Growth Des 12, 3489–3498 (2012).
18. Carne-Sanchez, A., Imaz, I., Cano-Sarabia, M., and Maspoch, D. A spray-drying strat-
egy for synthesis of nanoscale metal-organic frameworks and their assembly into hol-
low superstructures. Nat Chem 5, 203–211 (2013).
19. Liu, Y., Her, J. H., Dailly, A., Ramirez-Cuesta, A. J., Neumann, D. A., and Brown, C.
M. Reversible structural transition in MIL-53 with large temperature hysteresis. J Am
Chem Soc 130, 11813–11818 (2008).
20. Beurroies, I., Boulhout, M., Llewellyn, P. L., Kuchta, B., Ferey, G., Serre, C., and De-
noyel, R. Using pressure to provoke the structural transition of metal-organic frame-
works. Angew Chem Int Edit 49, 7526–7529 (2010).
21. Thallapally, P. K., Tian, J., Kishan, M. R., Fernandez, C. A., Dalgarno, S. J., McGrail,
P. B., Warren, J. E., and Atwood, J. L. Flexible (breathing) interpenetrated metal-or-
ganic frameworks for CO2 separation applications. J Am Chem Soc 130, 16842–16843
(2008).
22. Motkuri, R. K., Thallapally, P. K., Nune, S. K., Fernandez, C. A., McGrail, B. P., and
Atwood, J. L. Role of hydrocarbons in pore expansion and contraction of a flexible
metal-organic framework. Chem Commun 47, 7077–7079 (2011).
96 Industrial Catalysis and Separations: Innovations for Process Intensification
23. (a) Kitagawa, S. and Uemura, K. Dynamic porous properties of coordination polymers
inspired by hydrogen bonds. Chem Soc Rev 34, 109–119 (2005); (b) Ferey, G. and
Serre, C. Large breathing effects in three-dimensional porous hybrid matter: facts,
analyzes, rules and consequences. Chem Soc Rev 38, 1380–1399 (2009).
24. (a) Llewellyn, P. L., Horcajada, P., Maurin, G., Devic, T., Rosenbach, N., Bourrelly,
S., Serre, C., Vincent, D., Loera-Serna, S., Filinchuk, Y., and Ferey, G. Complex ad-
sorption of short linear alkanes in the flexible metal-organic-framework MIL-53(Fe).
J Am Chem Soc 131, 13002–13008 (2009); (b) Coudert, F. X., Boutin, A., Jeffroy, M.,
Mellot-Draznieks, C., and Fuchs, A. H. Thermodynamic methods and models to study
flexible metal-organic frameworks. Chemphyschem 12, 247–258 (2011).
25. Wang, Z. Q. and Cohen, S. M. Postsynthetic modification of metal-organic frame-
works. Chem Soc Rev 38, 1315–1329 (2009).
26. Yilmaz, B., Trukhan, N., and Muller, U. Industrial outlook on zeolites and metal or-
ganic frameworks. Chinese J Catal 33, 3–10 (2012).
27. (a) Zhao, D., Yuan, D. Q., and Zhou, H. C. The current status of hydrogen storage in
metal-organic frameworks. Energ Environ Sci 1, 222–235 (2008); (b) Murray, L. J.,
Dinca, M., and Long, J. R. Hydrogen storage in metal-organic frameworks. Chem
Soc Rev 38, 1294–1314 (2009); (c) Rowsell, J. L. C. and Yaghi, O. M. Strategies for
hydrogen storage in metal-organic frameworks. Angew Chem Int Edit 44, 4670–4679
(2005).
28. (a) Rowsell, J. L. C., Eckert, J., and Yaghi, O. M. Characterization of H-2 binding sites
in prototypical metal-organic frameworks by inelastic neutron scattering. J Am Chem
Soc 127, 14904–14910 (2005); (b) Ye, Q., Song, Y. M., Wang, G. X., Chen, K., Fu,
D. W., Chan, P. W. H., Zhu, J. S., Huang, S. D., and Xiong, R. G. Ferroelectric metal-
organic framework with a high dielectric constant. J Am Chem Soc 128, 6554–6555
(2006); (c) Rosi, N. L., Eckert, J., Eddaoudi, M., Vodak, D. T., Kim, J., O’Keeffe, M.,
and Yaghi, O. M. Hydrogen storage in microporous metal-organic frameworks. Sci-
ence 300, 1127–1129 (2003).
29. Alvaro, M., Carbonell, E., Ferrer, B., Xamena, F. X. L. I., and Garcia, H. Semicon-
ductor behavior of a metal-organic framework (MOF). Chem-Eur J 13, 5106–5112
(2007).
30. Klontzas, E., Mavrandonakis, A., Froudakis, G. E., Carissan, Y., and Klopper, W. Mo-
lecular hydrogen interaction with IRMOF-1: A multiscale theoretical study. J Phys
Chem C 111, 13635–13640 (2007).
31. (a) Younglove, B. A. Thermo-physical properties of fluids.1. Argon, Ethylene, Para-
Hydrogen, Nitrogen, Nitrogen Trifluoride, and Oxygen. J Phys Chem Ref Data 11,
1–353 (1982); (b) Kaye, S. S., Dailly, A., Yaghi, O. M., and Long, J. R. Impact of
preparation and handling on the hydrogen storage properties of Zn4O(1,4-benzenedi-
carboxylate)(3) (MOF-5). J Am Chem Soc 129, 14176–14177 (2007).
32. van den Berg, A. W. C. and Arean, C. O. Materials for hydrogen storage: current re-
search trends and perspectives. Chem Commun 668–681 (2008).
33. (a) Dinca, M., Dailly, A., Liu, Y., Brown, C. M., Neumann, D. A., and Long, J. R.
Hydrogen storage in a microporous metal-organic framework with exposed Mn2+ co-
ordination sites. J Am Chem Soc 128, 16876–16883 (2006); (b) Hu, Y. H. and Zhang,
L. Hydrogen storage in metal-organic frameworks. Adv Mater 22, E117–E130 (2010).
Metal Organic Frameworks–Synthesis and Applications 97
34. (a) Morris, R. E. and Wheatley, P. S. Gas storage in nanoporous materials. Angew
Chem Int Edit 47, 4966–4981 (2008); (b) Collins, D. J. and Zhou, H. C. Hydrogen
storage in metal-organic frameworks. J Mater Chem 17, 3154–3160 (2007).
35. Cheon, Y. E. and Suh, M. P. Multifunctional fourfold interpenetrating diamondoid
network: Gas separation and fabrication of palladium nanoparticles. Chem-Eur J 14,
3961– 3967 (2008).
36. Herm, Z. R., Swisher, J. A., Smit, B., Krishna, R., and Long, J. R. Metal-organic
frameworks as adsorbents for hydrogen purification and precombustion carbon diox-
ide capture. J Am Chem Soc 133, 5664–5667 (2011).
37. (a) Yang, Q. Y. and Zhong, C. L. Molecular simulation of carbon dioxide/methane/
hydrogen mixture adsorption in metal-organic frameworks. J Phys Chem B 110,
17776–17783 (2006); (b) Babarao, R. and Jiang, J. W. Unprecedentedly high selective
adsorption of gas mixtures in rho zeolite-like metal-organic framework: A molecular
simulation study. J Am Chem Soc 131, 11417–11425 (2009); (c) Liu, Y. H., Liu, D. H.,
Yang, Q. Y., Zhong, C. L., and Mi, J. G. Comparative study of separation performance
of COFs and MOFs for CH4/CO2/H-2 mixtures. Ind Eng Chem Res 49, 2902– 2906
(2010).
38. Duren, T., Sarkisov, L., Yaghi, O. M., and Snurr, R. Q. Design of new materials for
methane storage. Langmuir 20, 2683–2689 (2004).
39. Ma, S. Q., Sun, D. F., Simmons, J. M., Collier, C. D., Yuan, D. Q., and Zhou, H. C.
Metal-organic framework from an anthracene derivative containing nanoscopic cages
exhibiting high methane uptake. J Am Chem Soc 130, 1012–1016 (2008).
40. Zou, R. Q., Abdel-Fattah, A. I., Xu, H. W., Zhao, Y. S., and Hickmott, D. D. Storage
and separation applications of nanoporous metal-organic frameworks. Crystengcomm
12, 1337–1353 (2010).
41. Motkuri, R. K., Thallapally, P. K., McGrail, B. P., and Ghorishi, S. B. Dehydrated
Prussian blues for CO2 storage and separation applications. Crystengcomm 12, 4003–
4006 (2010).
42. Yazaydin, A. O., Snurr, R. Q., Park, T. H., Koh, K., Liu, J., LeVan, M. D., Benin, A. I.,
Jakubczak, P., Lanuza, M., Galloway, D. B., Low, J. J., and Willis, R. R. Screening of
metal-organic frameworks for carbon dioxide capture from flue gas using a combined
experimental and modeling approach. J Am Chem Soc 131, 18198–18199 (2009).
43. Caskey, S. R., Wong-Foy, A. G., and Matzger, A. J. Dramatic tuning of carbon dioxide
uptake via metal substitution in a coordination polymer with cylindrical pores. J Am
Chem Soc 130, 10870–10871 (2008).
44. Liu, J., Wang, Y., Benin, A. I., Jakubczak, P., Willis, R. R., and LeVan, M. D. CO2/
H2O adsorption equilibrium and rates on metal-organic frameworks: HKUST-1 and
Ni/DOBDC. Langmuir 26, 14301–14307 (2010).
45. Liu, B. and Smit, B. Comparative molecular simulation study of CO2/N2 and CH4/
N2 separation in zeolites and metal-organic frameworks. Langmuir 25, 5918–5926
(2009).
46. Wu, D., Xu, Q., Liu, D. H., and Zhong, C. L. Exceptional CO2 capture capability and
molecular-level segregation in a Li-modified metal-organic framework. J Phys Chem
C 114, 16611–16617 (2010).
47. (a) Kishan, M. R., Tian, J., Thallapally, P. K., Fernandez, C. A., Dalgarno, S. J., War-
ren, J. E., McGrail, B. P., and Atwood, J. L. Flexible metal-organic supramolecular
98 Industrial Catalysis and Separations: Innovations for Process Intensification
isomers for gas separation. Chem Commun 46, 538–540 (2010); (b) Tian, J. A., Mot-
kuri, R. K., Thallapally, P. K., and McGrail, B. P. Metal-organic framework isomers
with diamondoid networks constructed of a semirigid tetrahedral linker. Cryst Growth
Des 10, 5327–5333 (2010); (c) Tian, J. A., Motkuri, R. K., and Thallapally, P. K.
Generation of 2D and 3D (PtS, adamantanoid) nets with a flexible tetrahedral building
block. Cryst Growth Des 10, 3843–3846 (2010).
48. Yazaydin, A. O., Benin, A. I., Faheem, S. A., Jakubczak, P., Low, J. J., Willis, R. R.,
and Snurr, R. Q. Enhanced CO2 adsorption in metal-organic frameworks via occupa-
tion of open-metal sites by coordinated water molecules. Chem Mater 21, 1425–1430
(2009).
49. Nune, S. K., Thallapally, P. K., and McGrail, B. P. Metal organic gels (MOGs): a
new class of sorbents for CO2 separation applications. J Mater Chem 20, 7623–7625
(2010).
50. Yang, C., Wang, X. P., and Omary, M. A. Fluorous metal-organic frameworks for high-
density gas adsorption. J Am Chem Soc 129, 15454–15455 (2007).
51. Farha, O. K., Mulfort, K. L., and Hupp, J. T. An example of node-based postassembly
elaboration of a hydrogen-sorbing, metal-organic framework material. Inorg Chem 47,
10223–10225 (2008).
52. Henninger, S. K., Habib, H. A., and Janiak, C. MOFs as adsorbents for low tempera-
ture heating and cooling applications. J Am Chem Soc 131, 2776–2777 (2009).
53. Jeremias, F., Khutia, A., Henninger, S. K., and Janiak, C. MIL-100 (Al, Fe) as water
adsorbents for heat transformation purposes-a promising application. J Mater Chem
22, 10148–10151 (2012).
54. Henninger, S. K., Schmidt, F. P., and Henning, H. M. Water adsorption characteristics
of novel materials for heat transformation applications. Appl Therm Eng 30, 1692–
1702 (2010).
55. Seo, Y. K., Yoon, J. W., Lee, J. S., Hwang, Y. K., Jun, C. H., Chang, J. S., Wuttke, S.,
Bazin, P., Vimont, A., Daturi, M., Bourrelly, S., Llewellyn, P. L., Horcajada, P., Serre,
C., and Ferey, G. Energy-efficient dehumidification over hierachically porous metal-
organic frameworks as advanced water adsorbents. Adv Mater 24, 806–810 (2012).
56. (a) Britt, D., Tranchemontagne, D., and Yaghi, O. M. Metal-organic frameworks with
high capacity and selectivity for harmful gases. P Natl Acad Sci USA 105, 11623–
11627 (2008); (b) Khan, N. A., Hasan, Z., and Jhung, S. H. Adsorptive removal of
hazardous materials using metal-organic frameworks (MOFs): A review. J Hazard
Mater 244, 444–456 (2013); (c) Thallapally, P. K., Motkuri, R. K., Fernandez, C. A.,
McGrail, B. P., and Behrooz, G. S. Prussian blue analogs for CO2 and SO2 capture and
separation applications. Inorg Chem 49, 4909–4915 (2010).
57. (a) Petit, C. and Bandosz, T. J. Enhanced adsorption of ammonia on metal-organic
framework/graphite oxide composites: analysis of surface interactions. Adv Funct
Mater 20, 111–118 (2010); (b) Peterson, G. W., Wagner, G. W., Balboa, A., Mahle,
J., Sewell, T., and Karwacki, C. J. Ammonia vapor removal by Cu-3(BTC)(2) and
its characterization by MAS NMR. J Phys Chem C 113, 13906–13917 (2009); (c)
Sumida, K., Rogow, D. L., Mason, J. A., McDonald, T. M., Bloch, E. D., Herm, Z.
R., Bae, T. H., and Long, J. R. Carbon dioxide capture in metal-organic frameworks.
Chem Rev 112, 724–781 (2012); (d) Zhao, Z. X., Li, X. M., Huang, S. S., Xia, Q. B.,
and Li, Z. Adsorption and diffusion of benzene on chromium-based metal organic
Metal Organic Frameworks–Synthesis and Applications 99
framework MIL-101 synthesized by microwave irradiation. Ind Eng Chem Res 50,
2254–2261 (2011); (e) Glover, T. G., Peterson, G. W., Schindler, B. J., Britt, D., and
Yaghi, O. MOF-74 building unit has a direct impact on toxic gas adsorption. Chem
Eng Sci 66, 163–170 (2011); (f) Hamon, L., Serre, C., Devic, T., Loiseau, T., Millange,
F., Ferey, G., and De Weireld, G. Comparative study of hydrogen sulfide adsorption
in the MIL-53(Al, Cr, Fe), MIL-47(V), MIL-100(Cr), and MIL-101(Cr) metal-organic
frameworks at room temperature. J Am Chem Soc 131, 8775–8777 (2009); (g) Karra,
J. R. and Walton, K. S. Effect of open metal sites on adsorption of polar and nonpolar
molecules in metal-organic framework Cu-BTC. Langmuir 24, 8620–8626 (2008); (h)
Hamon, L., Leclerc, H., Ghoufi, A., Oliviero, L., Travert, A., Lavalley, J. C., Devic, T.,
Serre, C., Ferey, G., De Weireld, G., Vimont, A., and Maurin, G. Molecular insight into
the adsorption of H2S in the flexible MIL-53(Cr) and rigid MIL-47(V) MOFs: Infrared
spectroscopy combined to molecular simulations. J Phys Chem C 115, 2047–2056
(2011).
58. Petit, C. and Bandosz, T. J. Exploring the coordination chemistry of MOF-graphite ox-
ide composites and their applications as adsorbents. Dalton T 41, 4027–4035 (2012).
59. Fernandez, C. A., Thallapally, P. K., Motkuri, R. K., Nune, S. K., Sumrak, J. C., Tian,
J., and Liu, J. Gas-induced expansion and contraction of a fluorinated metal-organic
framework. Cryst Growth Des 10, 1037–1039 (2010).
60. Fernandez, C. A., Thallapally, P. K., and McGrail, B. P. Insights into the temperature
dependent “breathing” of a flexible fluorinated metal-organic framework. Chemphy-
schem 13, 3275–3281 (2012).
61. Thallapally, P. K., Grate, J. W., and Motkuri, R. K. Facile xenon capture and release
at room temperature using a metal-organic framework: a comparison with activated
charcoal. Chem Commun 48, 347–349 (2012).
62. Meek, S. T., Teich-McGoldrick, S. L., Perry, J. J., Greathouse, J. A., and Allendorf,
M. D. Effects of polarizability on the adsorption of noble gases at low pressures in
monohalogenated isoreticular metal-organic frameworks. J Phys Chem C 116, 19765–
19772 (2012).
63. Liu, J., Thallapally, P. K., and Strachan, D. Metal-organic frameworks for removal of
Xe and Kr from nuclear fuel reprocessing plants. Langmuir 28, 11584–11589 (2012).
64. Sikora, B. J., Wilmer, C. E., Greenfield, M. L., and Snurr, R. Q. Thermodynamic
analysis of Xe/Kr selectivity in over 137000 hypothetical metal-organic frameworks.
Chem Sci 3, 2217–2223 (2012).
65. Fernandez, C. A., Liu, J., Thallapally, P. K., and Strachan, D. M. Switching Kr/Xe se-
lectivity with temperature in a metal-organic framework. J Am Chem Soc 134, 9046–
9049 (2012).
66. (a) Lee, J., Farha, O. K., Roberts, J., Scheidt, K. A., Nguyen, S. T., and Hupp, J. T.
Metal-organic framework materials as catalysts. Chem Soc Rev 38, 1450–1459 (2009);
(b) Thallapally, P. K., Fernandez, C. A., Motkuri, R. K., Nune, S. K., Liu, J., and
Peden, C. H. F. Micro and mesoporous metal-organic frameworks for catalysis ap-
plications. Dalton T 39, 1692–1694 (2010).
67. Dhakshinamoorthy, A., Alvaro, M., Corma, A., and Garcia, H. Delineating similarities
and dissimilarities in the use of metal organic frameworks and zeolites as heteroge-
neous catalysts for organic reactions. Dalton T 40, 6344–6360 (2011).
100 Industrial Catalysis and Separations: Innovations for Process Intensification
68. Vermoortele, F., Ameloot, R., Alaerts, L., Matthessen, R., Carlier, B., Fernandez, E.
V. R., Gascon, J., Kapteijn, F., and De Vos, D. E. Tuning the catalytic performance of
metal-organic frameworks in fine chemistry by active site engineering. J Mater Chem
22, 10313– 10321 (2012).
69. Chui, S. S. Y., Lo, S. M. F., Charmant, J. P. H., Orpen, A. G., and Williams, I. D. A
chemically functionalizable nanoporous material [Cu-3(TMA)(2)(H2O)(3)](n). Sci-
ence 283, 1148–1150 (1999).
70. Alaerts, L., Seguin, E., Poelman, H., Thibault-Starzyk, F., Jacobs, P. A., and De Vos,
D. E. Probing the Lewis acidity and catalytic activity of the metal-organic framework
[Cu-3(btc)(2)] (BTC = benzene-1,3,5-tricarboxylate). Chem-Eur J 12, 7353–7363
(2006).
71. Schlichte, K., Kratzke, T., and Kaskel, S. Improved synthesis, thermal stability and
catalytic properties of the metal-organic framework compound CU3(BTC)(2). Micro-
por Mesopor Mat 73, 81–88 (2004).
72. Horike, S., Dinca, M., Tamaki, K., and Long, J. R. Size-selective lewis acid catalysis
in a microporous metal-organic framework with exposed Mn(2+) coordination sites. J
Am Chem Soc 130, 5854–5855 (2008).
73. Gandara, F., Gornez-Lor, B., Gutierrez-Puebla, E., Iglesias, M., Monge, M. A., Proser-
pio, D. M., and Snejko, N. An indium layered MOF as recyclable lewis acid catalyst.
Chem Mater 20, 72–76 (2008).
74. Horcajada, P., Surble, S., Serre, C., Hong, D. Y., Seo, Y. K., Chang, J. S., Greneche, J.
M., Margiolaki, I., and Ferey, G. Synthesis and catalytic properties of MIL-100(Fe),
an iron(III) carboxylate with large pores. Chem Commun 2820–2822 (2007).
75. (a) Vimont, A., Goupil, J. M., Lavalley, J. C., Daturi, M., Surble, S., Serre, C., Mil-
lange, F., Ferey, G., and Audebrand, N. Investigation of acid sites in a zeotypic giant
pores chromium(III) carboxylate. J Am Chem Soc 128, 3218–3227 (2006); (b) Vimont,
A., Leclerc, H., Mauge, F., Daturi, M., Lavalley, J. C., Surble, S., Serre, C., and Ferey,
G. Creation of controlled Bronsted acidity on a zeotypic mesoporous chromium(III)
carboxylate by grafting water and alcohol molecules. J Phys Chem C 111, 383–388
(2007).
76. Suslick, K. S., Bhyrappa, P., Chou, J. H., Kosal, M. E., Nakagaki, S., Smithenry, D.
W., and Wilson, S. R. Microporous porphyrin solids. Accounts Chem Res 38, 283–291
(2005).
77. Savonnet, M., Aguado, S., Ravon, U., Bazer-Bachi, D., Lecocq, V., Bats, N., Pinel,
C., and Farrusseng, D. Solvent free base catalysis and transesterifiation over basic
functionalised Metal-organic frameworks. Green Chem 11, 1729–1732 (2009).
78. Juan-Alcaniz, J., Gascon, J., and Kapteijn, F. Metal-organic frameworks as scaffolds
for the encapsulation of active species: state-of-the-art and future perspectives. J Ma-
ter Chem 22, 10102–10118 (2012).
79. Ishida, T., Nagaoka, M., Akita, T., and Haruta, M. Deposition of gold clusters on po-
rous coordination polymers by solid grinding and their catalytic activity in aerobic
oxidation of alcohols. Chem-Eur J 14, 8456–8460 (2008).
80. Hermes, S., Schroter, M. K., Schmid, R., Khodeir, L., Muhler, M., Tissler, A., Fischer,
R. W., and Fischer, R. A. Metal@MOF: Loading of highly porous coordination poly-
mers host lattices by metal organic chemical vapor deposition. Angew Chem Int Edit
44, 6237–6241 (2005).
Metal Organic Frameworks–Synthesis and Applications 101
81. Rahiman, A. K., Rajesh, K., Bharathi, K. S., Sreedaran, S., and Narayanan, V. Cata-
lytic oxidation of alkenes by manganese(III) porphyrin-encapsulated Al, V, Si-meso-
porous molecular sieves. Inorg Chim Acta 362, 1491–1500 (2009).
82. Ferey, G. A chromium terephthalate-based solid with unusually large pore volumes
and surface area (vol 309, pg 2040, 2005). Science 310, 1119–1119 (2005).
83. Taylor-Pashow, K. M. L., Della Rocca, J., Xie, Z. G., Tran, S., and Lin, W. B. Post-
synthetic modifications of iron-carboxylate nanoscale metal-organic frameworks for
imaging and drug delivery. J Am Chem Soc 131, 14261–14263 (2009).
84. Maksimchuk, N. V., Timofeeva, M. N., Melgunov, M. S., Shmakov, A. N., Chesalov,
Y. A., Dybtsev, D. N., Fedin, V. P., and Kholdeeva, O. A. Heterogeneous selective
oxidation catalysts based on coordination polymer MIL-101 and transition metal-sub-
stituted polyoxometalates. J Catal 257, 315–323 (2008).
85. Maksimchuk, N. V., Kovalenko, K. A., Arzumanov, S. S., Chesalov, Y. A., Melgunov,
M. S., Stepanov, A. G., Fedin, V. P., and Kholdeeva, O. A. Hybrid polyoxotungstate/
MIL-101 materials: synthesis, characterization, and catalysis of H2O2-based alkene ep-
oxidation. Inorg Chem 49, 2920–2930 (2010).
86. (a) Wang, Z. Q. and Cohen, S. M. Postsynthetic covalent modification of a neutral
metal-organic framework. J Am Chem Soc 129, 12368–12369 (2007); (b) Tanabe, K.
K., Wang, Z. Q., and Cohen, S. M. Systematic functionalization of a metal-organic
framework via a postsynthetic modification approach. J Am Chem Soc 130, 8508–
8517 (2008).
87. Hong, D. Y., Hwang, Y. K., Serre, C., Ferey, G., and Chang, J. S. Porous chromium
terephthalate MIL-101 with coordinatively unsaturated sites: Surface functionaliza-
tion, encapsulation, sorption and catalysis. Adv Funct Mater 19, 1537–1552 (2009).
88. Zhang, X., Xamena, F. X. L. I., and Corma, A. Gold(III) – metal organic framework
bridges the gap between homogeneous and heterogeneous gold catalysts. J Catal
265,155– 160 (2009).
89. Alkordi, M. H., Liu, Y. L., Larsen, R. W., Eubank, J. F., and Eddaoudi, M. Zeolite-like
metal-organic frameworks as platforms for applications: On metalloporphyrin-based
catalysts. J Am Chem Soc 130, 12639–12641 (2008).
90. Zhou, W., Wu, H., Yildirim, T., Simpson, J. R., and Walker, A. R. H. Origin of the
exceptional negative thermal expansion in metal-organic framework-5 Zn(4)O(1,4-
benzenedicarboxylate)(3). Phys Rev B 78, 054114 (2008).
91. (a) Takashima, Y., Martinez, V. M., Furukawa, S., Kondo, M., Shimomura, S., Uehara,
H., Nakahama, M., Sugimoto, K., and Kitagawa, S. Molecular decoding using lumi-
nescence from an entangled porous framework. Nat Commun 2 (2011); (b) Harbuzaru,
B. V., Corma, A., Rey, F., Atienzar, P., Jorda, J. L., Garcia, H., Ananias, D., Carlos, L.
D., and Rocha, J. Metal-organic nanoporous structures with anisotropic photolumi-
nescence and magnetic properties and their use as sensors. Angew Chem Int Edit 47,
1080–1083 (2008); (c) Achmann, S., Hagen, G., Kita, J., Malkowsky, I. M., Kiener,
C., and Moos, R. Metal-organic frameworks for sensing applications in the gas phase.
Sensors-Basel 9, 1574–1589 (2009); (d) Southon, P. D., Liu, L., Fellows, E. A., Price,
D. J., Halder, G. J., Chapman, K. W., Moubaraki, B., Murray, K. S., Letard, J. F., and
Kepert, C. J. Dynamic interplay between spin-crossover and host-guest function in a
nanoporous metal-organic framework material. J Am Chem Soc 131, 10998–11009
(2009); (e) Maspoch, D., Ruiz-Molina, D., Wurst, K., Domingo, N., Cavallini, M.,
102 Industrial Catalysis and Separations: Innovations for Process Intensification
Biscarini, F., Tejada, J., Rovira, C., and Veciana, J. A nanoporous molecular magnet
with reversible solvent-induced mechanical and magnetic properties. Nat Mater 2,
190–195 (2003).
92. Serre, C., Mellot-Draznieks, C., Surble, S., Audebrand, N., Filinchuk, Y., and Ferey,
G. Role of solvent-host interactions that lead to very large swelling of hybrid frame-
works. Science 315, 1828–1831 (2007).
93. Goeders, K. M., Colton, J. S., and Bottomley, L. A. Microcantilevers: Sensing chemi-
cal interactions via mechanical motion. Chem Rev 108, 522–542 (2008).
94. Allendorf, M. D., Houk, R. J. T., Andruszkiewicz, L., Talin, A. A., Pikarsky, J., Choud-
hury, A., Gall, K. A., and Hesketh, P. J. Stress-induced chemical detection using flex-
ible metal-organic frameworks. J Am Chem Soc 130, 14404–14405 (2008).
95. Liu, B., Ma, M. Y., Zacher, D., Betard, A., Yusenko, K., Metzler-Nolte, N., Woll, C.,
and Fischer, R. A. Chemistry of SURMOFs: Layer-selective installation of functional
groups and postsynthetic covalent modification probed by fluorescence microscopy. J
Am Chem Soc 133, 1734–1737 (2011).
96. Liu, B., Shekhah, O., Arslan, H. K., Liu, J. X., Woll, C., and Fischer, R. A. Enantiopure
metal-organic framework thin films: Oriented SURMOF growth and enantioselective
adsorption. Angew Chem Int Edit 51, 807–810 (2012).
97. Biemmi, E., Darga, A., Stock, N., and Bein, T. Direct growth of Cu-3(BTC)(2)(H2O)
(3)center dot xH(2)O thin films on modified QCM-gold electrodes-Water sorption
isotherms. Micropor Mesopor Mat 114, 380–386 (2008).
98. Robinson, A. L., Stavila, V. L., Zeitler, T. R., White, M. I., Thornberg, S. M., Great-
house, J. A., and Allendorf, M. D. Ultrasensitive humidity detection using metal-or-
ganic framework-coated microsensors. Anal Chem 84, 7043–7051 (2012).
99. Ameloot, R., Stappers, L., Fransaer, J., Alaerts, L., Sels, B. F., and De Vos, D. E.
Patterned growth of metal-organic framework coatings by electrochemical synthesis.
Chem Mater 21, 2580–2582 (2009).
100. Li, Y., Zhang, S. S., and Song, D. T. A luminescent metal-organic framework as a
turn-on sensor for DMF vapor. Angew Chem Int Edit 52, 710–713 (2013).
101. Lan, A. J., Li, K. H., Wu, H. H., Olson, D. H., Emge, T. J., Ki, W., Hong, M. C., and
Li, J. A luminescent microporous metal-organic framework for the fast and reversible
detection of high explosives. Angew Chem Int Edit 48, 2334–2338 (2009).
102. Lu, Z. Z., Zhang, R., Li, Y. Z., Guo, Z. J., and Zheng, H. G. Solvatochromic behavior
of a nanotubular metal-organic framework for sensing small molecules. J Am Chem
Soc 133, 4172–4174 (2011).
103. Beauvais, L. G., Shores, M. P., and Long, J. R. Cyano-bridged Re(6)Q(8) (Q = S, Se)
cluster-Cobalt(II) framework materials: Versatile solid chemical sensors. J Am Chem
Soc 122, 2763–2772 (2000).
104. Lee, H., Jung, S. H., Han, W. S., Moon, J. H., Kang, S., Lee, J. Y., Jung, J. H., and
Shinkai, S. A Chromo-fluorogenic tetrazole-based CoBr2 coordination polymer gel as
a highly sensitive and selective chemosensor for volatile gases containing chloride.
Chem-Eur J 17, 2823–2827 (2011).
105. Orcajo, B., Muruzabal, F., Isasmendi, M. C., Gutierrez, N., Sanchez, M., Orive, G.,
and Anitua, E. The use of plasma rich in growth factors (PRGF-Endoret) in the treat-
ment of a severe mal perforant ulcer in the foot of a person with diabetes. Diabetes
Res Clin Pr 93, E65–E67 (2011).
Metal Organic Frameworks–Synthesis and Applications 103
106. Dave, D. and Ghaly, A. E. Remediation technologies for marine oil spills: A critical
review and comparative analysis. American Journal of Environmental Sciences 7,
423–440 (2011).
107. (a) Adebajo, M. O., Frost, R. L., Kloprogge, J. T., Carmody, O., and Kokot, S. Porous
materials for oil spill cleanup: A review of synthesis and absorbing properties. J Po-
rous Mat 10, 159–170 (2003); (b) Tao, S. Y., Wa’ng, Y. C., and An, Y. L. Superwet-
ting monolithic SiO2 with hierarchical structure for oil removal. J Mater Chem 21,
11901– 11907 (2011); (c) Venkatanarasimhan, S. and Raghavachari, D. Epoxidized
natural rubber-magnetite nanocomposites for oil spill recovery. J Mater Chem A 1,
868–876 (2013); (d) Wang, X. S., Liu, J., Bonefont, J. M., Yuan, D. Q., Thallapally,
P. K., and Ma, S. Q. A porous covalent porphyrin framework with exceptional uptake
capacity of saturated hydrocarbons for oil spill cleanup. Chem Commun 49, 1533–
1535 (2013); (e) Zhu, Q., Pan, Q. M., and Liu, F. T. Facile removal and collection of
oils from water surfaces through superhydrophobic and superoleophilic sponges. J
Phys Chem C 115, 17464–17470 (2011).
108. USE OF SORBENT MATERIALS IN OIL SPILL RESPONSE Technical Report
from The International Tanker Owners Pollution Federation Limited (ITOPF) (2013).
109. Wu, H. H., Gong, Q. H., Olson, D. H., and Li, J. Commensurate adsorption of hy-
drocarbons and alcohols in microporous metal organic frameworks. Chem Rev 112,
836– 868 (2012).
110. Cohen, S. M. Postsynthetic methods for the functionalization of metal-organic frame-
works. Chem Rev 112, 970–1000 (2012).
111. Yang, C., Kaipa, U., Mather, Q. Z., Wang, X. P., Nesterov, V., Venero, A. F., and Om-
ary, M. A. Fluorous metal-organic frameworks with superior adsorption and hydro-
phobic properties toward oil spill cleanup and hydrocarbon storage. J Am Chem Soc
133, 18094– 18097 (2011).
112. Maes, M., Alaerts, L., Vermoortele, F., Ameloot, R., Couck, S., Finsy, V., Denayer,
J. F. M., and De Vos, D. E. Separation of C-5-hydrocarbons on microporous ma-
terials: Complementary performance of MOFs and Zeolites. J Am Chem Soc 132,
2284–2292 (2010).
113. Nijem, N., Wu, H. H., Canepa, P., Marti, A., Balkus, K. J., Thonhauser, T., Li, J., and
Chabal, Y. J. Tuning the gate opening pressure of metal-organic frameworks (MOFs)
for the selective separation of hydrocarbons. J Am Chem Soc 134, 15201–15204
(2012).
INDEX
... 38−41 This is a desirable feature since it means that families (platforms) of related materials can be prepared 42−48 and their functional properties can be studied systematically. 46,49,50 An increasingly important subset of MOMs is chiral MOMs (CMOMs), which are composed of homochiral ligands or chiral channels that arise from the crystal packing of achiral components. Since the report of POST-1 in 2000, 51 it has been realized that CMOMs offer potential utility in asymmetric catalysis, chiral detection, and enantiomeric separations. ...
Article
Full-text available
Chiral metal−organic materials (CMOMs), can offer molecular binding sites that mimic the enantioselectivity exhibited by biomolecules and are amenable to systematic fine-tuning of structure and properties. Herein, we report that the reaction of Ni(NO 3) 2 , S-indoline-2-carboxylic acid (S-IDECH), and 4,4′-bipyridine (bipy) afforded a homochiral cationic diamondoid, dia, network, [Ni(S-IDEC)(bipy)(H 2 O)][NO 3 ], CMOM-5. Composed of rod building blocks (RBBs) cross-linked by bipy linkers, the activated form of CMOM-5 adapted its pore structure to bind four guest molecules, 1-phenyl-1-butanol (1P1B), 4-phenyl-2-butanol (4P2B), 1-(4-methoxyphenyl)ethanol (MPE), and methyl mandelate (MM), making it an example of a chiral crystalline sponge (CCS). Chiral resolution experiments revealed enantiomeric excess, ee, values of 36.2−93.5%. The structural adaptability of CMOM-5 enabled eight enantiomer@CMOM-5 crystal structures to be determined. The five ordered crystal structures revealed that host−guest hydrogen-bonding interactions are behind the observed enantioselectivity, three of which represent the first crystal structures determined of the ambient liquids R-4P2B, S-4P2B, and R-MPE.
... (20,21) The mechanisms involved in the formation of MOF's having unusual morphologies have yet to fully understood. (22)(23)(24)(25) In this work, we report a unique paradoxical crystallographic phenomenon: single crystals with a multidomain polyhedron morphology and concave (re-entrant) angles. According to accepted crystallographic conventions, these characteristics should not coexist: it is thought to be impossible to generate structures featuring re-entrant angles with any of the unit cell geometries. ...
Preprint
Full-text available
The symmetry of a crystal’s morphology usually reflects the symmetry of the crystallographic packing. For single-crystals, the space and point groups allow only a limited number of mathematical descriptions of the morphology (forms), all of which are convex polyhedrons. In contrast, concave polyhedrons are a hallmark of twinning and polycrystallinity and are inconsistent with single crystallinity. Here we report unique metallo-organic crystals that exhibit, on the one hand, single crystallinity and, on the other hand, a concave, multidomain appearance and a rare space group (P622). Despite these contradictions, it is puzzling that the symmetry of these crystals is linked at different hierarchies, from the molecular to the microscopic levels. They exhibit a deceptive combination of the three characteristics that define a crystal at different length scales: the unit cell, the organization of these unit cells, and morphology. The structural properties of this system are intrinsic, as no other materials (additives) were used to shape the facets. The unusual concave morphology of the crystals can be a direct consequence of the open molecular packing often found in metal-organic frameworks. This fact has not been recognized before in these and other porous materials. Such crystals offer new opportunities for the formation of 3D objects where the physical properties can be designed and predicted based on the point group symmetry.
... According to IUPAC recommendations 2013, Metal-Organic Frameworks (MOFs) are Coordination Polymers with organic ligands, extended in two or three dimensions, containing potential voids [8]. The possibility to choose the organic ligand or linker and the metal ions, allows tuning the structure, pore size, surface area and multiple functionalities in a rational way [9]. Due to their crystalline structure, tailorable ligands, presence of pores and metallic centers, applications of MOFs have been proposed for many applications like: gas storage and separation [10], drug delivery [11], sensors [12], catalysis [13] and optoelectronic devices [14]. ...
Article
Full-text available
In order to obtain platinum-group-metal-free catalysts for the oxygen evolution reaction (OER) and the hydrogen evolution reaction (HER) in alkaline electrolyzers, nitrogen-doped mesoporous carbons were prepared from pyrolysis of two cobalt metal organic frameworks (MOFs), one linear coordination polymer and one complex. The catalyst derived from cobalt 2,3-pyrazinedicarboxylate polymer (700 °C) had a Tafel slope of 90 mV dec⁻¹ and 130 mV dec⁻¹ for the HER and OER respectively, a current density of 10 mA cm⁻² was reached at − 0.23 V and 1.56 V vs RHE for the HER and OER respectively. In order to prove the bifunctional catalytic activity towards the overall water splitting, a two electrode electrolyzer was constructed depositing the catalyst on carbon paper. H2 and O2 evolved volumes followed the Faraday law, showing efficiency very close to 100%. The same materials also showed catalytic activity towards the oxygen reduction reaction (ORR), reaching an electron transfer number close to 4 and low H2O2 yields. The acid leached catalyst derived from Cobalt 2,3-pyrazinedicarboxylate polymer (700 °C) reached a ∆E = Ej = 10(OER) −E½(ORR) = 0.80 V, making it an useful oxygen catalyst for metal-air batteries.
... [61,62] Si bien la IUPAC aún no ha promulgado formalmente una definición del término o de las características que deben cumplir los films para clasificarse como SURMOFs, es ampliamente aceptado que se trata de multicapas cristalinas, orientadas y contínuas de MOFs, preparadas por LbL sobre un sustrato debidamente funcionalizado. [48,63] Los SURMOFs, deben cumplir con lo siguiente: ...
Thesis
Full-text available
RESUMEN: La Tesis se centró en el uso de ZIF-8, una Red Metal-Orgánica (MOF) en una configuración de film delgado sobre superficies electroactivas. Se analizó su crecimiento, estabilidad y propiedades funcionales, así como el efecto de diversas modificaciones tanto en el sustrato (uso de capas de silanos, monocapas autoensambladas de tioles y brushes poliméricos) como en los films desarrollados (inclusión de polielectrolitos). ABSTRACT: The focus of this thesis was the use of ZIF-8, a Metal-Organic Framework (MOF) in thin-film configuration, as a modifier of electroactive surfaces. It includes the analysis of the growth, stability, and functional properties of the native films. Additionally, the effect of different chemical modifications of the substrate (silane layers, thiol Self-Assembled Monolayers, and polymer brushes) and the final films (polyelectrolyte inclusion) was also studied.
Chapter
The properties of materials are governed by the interactions among its associated sub-units. This chapter will describe the bonding motifs of both crystalline and amorphous solids. Details of common archetypical crystal structures will be given, as well as introductory X-ray crystallography. The various types of defects in solids are also described, which are critical in understanding electrical conductivity and optical properties of crystalline solids. The structure vs. property relationship for key classes of materials such as ceramics, glasses, semiconductors, insulators, and gemstones will be described for a myriad of applications.
Article
Full-text available
Metal ions or clusters that have been bonded with organic linkers to create one- or more-dimensional structures are referred to as metal-organic frameworks (MOFs). Reticular synthesis also forms MOFs with properly designated components that can result in crystals with high porosities and great chemical and thermal stability. Due to the wider surface area, huge pore size, crystalline nature, and tunability, numerous MOFs have been shown to be potential candidates in various fields like gas storage and delivery, energy storage, catalysis, and chemical/biosensing. This study provides a quick overview of the current MOF synthesis techniques in order to familiarize newcomers in the chemical sciences field with the fast-growing MOF research. Beginning with the classification and nomenclature of MOFs, synthesis approaches of MOFs have been demonstrated. We also emphasize the potential applications of MOFs in numerous fields such as gas storage, drug delivery, rechargeable batteries, supercapacitors, and separation membranes. Lastly, the future scope is discussed along with prospective opportunities for the synthesis and application of nano-MOFs, which will help promote their uses in multidisciplinary research.
Preprint
Full-text available
Nanostructured materials such as metal--organic frameworks and perovskites can be easily tuned towards applications ranging from sensors to photovoltaic devices. However, the key to unlock their design potential, namely causal relations between a material's atomic structure and its macroscopic function, is currently still missing. Therefore, we herein introduce strain engineering as a general approach to rationalize and design how atomic-level structural modifications induce dynamically interacting strain fields that dictate these material's macroscopic mechanical behavior. We demonstrate the potential of strain engineering by consciously designing shear instabilities in UiO-66, leading to intriguing, counterintuitive mechanical behavior. The strain-engineered structures exhibit time- and space-dependent crumple zones that instill flexibility in the otherwise rigid material and that locally focus the strain, partially preserving the porosity of the material under compression. This example demonstrates how strain engineering can be adopted to design, from the atomic level onwards, state-of-the-art materials for challenging applications.
Article
Three isomeric metal-organic frameworks (MOFs) such as MAF-5, -6, and -32 (with the same composition of [Zn(2-ethylimidazole)2]) were carbonized and, for the first time, activated further with KOH to prepare highly porous MOF-derived carbons (MDCs). Importantly, MDC-32 derived from non-porous MAF-32 had the highest porosity among the three MDCs although it has the lowest porosity when no KOH activation was done. Adsorption of sulfanilamide and chloroxylenol from water was investigated with the MDCs. Among the MDCs, MDC-32 showed the best adsorptive performance for sulfanilamide and chloroxylenol. Moreover, MDC-32, had the highest adsorption capacity (256 mg/g) for removing sulfanilamide from water, compared with any adsorbent reported so far. Based on the observed adsorption and properties of the adsorbate and adsorbent, π–π and hydrogen bonding interactions, with a slight contribution of repulsive electrostatic interaction, could be suggested as the mechanism for the sulfanilamide adsorption over the MDC-32. Moreover, the MDC-32 could be recycled easily for up to four cycles. It could be suggested that non-porous MOFs can be a good precursor for highly porous MDCs, if activated well using KOH, for example. Finally, MAF-32-derived carbon, MDC-32, might be suggested as a plausible adsorbent to eliminate organics such as sulfanilamide from water.
Chapter
The preparation of functional solids where the macroscopic properties are governed by photochromic molecular components is enabled by nanoporous metal–organic frameworks (MOFs). The photoresponsive molecules can be incorporated in a high density in or at the crystalline MOF lattice as well as in the pores. The molecular photoisomerization goes along with substantial modifications of the host–guest interaction, yielding light‐responsive MOFs. These materials demonstrated unprecedented functionalities, ranging from photo‐switching the adsorption and diffusion properties of the guests, over the photo‐modulation of the membrane separation as well as the singlet oxygen generation up to the remote control of proton and electron conduction properties.
Article
Phenolic compounds would be the emerging pollutant by 2050, because of their wide spread applicability in daily life and therefore the adoption of suitable detection methods in which identification and separation of isomers is highly desirable. Owing to the fascinating features, Metal-organic framework (MOF), a class of reticular materials holds a large surface area with tunable shape and adjustable porosity will provide strong interaction with analytes through abundant functional groups resulting in high selectivity towards electrochemical determination of phenolic isomers. Nevertheless, the sensing performance can still be further improved by building MOF network (intrinsic resistance) with functional (conducting) materials, resulting in MOF based nanocomposite. Herein, this review provides the summary of MOF based nanocomposites for electrochemical sensing of phenolic compounds developed from 2015. In this review, we discussed the demerits of pristine MOF as electrode materials, and the requirement of new class of MOF with functional materials such as nanomaterials, carbon nanotubes, graphene and MXene. The history and evolution of MOF nanocomposite-based materials are discussed and also featured the impressive physical and chemical properties. Besides this review discusses the factors influencing the conducting pathway and mass transport of MOF based nanocomposite for enhanced sensing performance of phenolic compounds with suitable mechanistic illustrations. Finally, the major challenges governing the determination of phenolic compounds and the future advancements required for the development of MOF based electrodes for various applications are highlighted.
Article
Full-text available
Metal–organic frameworks (MOFs) are porous crystals with the potential to improve many industrial gas separation processes. Because there is a practically unlimited number of different MOFs, which vary in their pore geometry and chemical composition, it is challenging to find the best MOF for a given application. Here, we applied high-throughput computational methods to rapidly explore thousands of possible MOFs, given a library of starting materials, in the context of Xe/Kr separation. We generated over 137000 structurally diverse hypothetical MOFs from a library of chemical building blocks and screened them for Xe/Kr separation. For each MOF, we calculated geometric properties via Delaunay tessellation and predicted thermodynamic Xe/Kr adsorption behavior via multicomponent grand canonical Monte Carlo simulations. Specifically, we calculated the pore limiting diameter, largest cavity diameter, accessible void volume, as well as xenon and krypton adsorption at 1.0, 5.0 and 10 bar at 273 K. From these data we show that MOFs with pores just large enough to fit a single xenon atom, and having morphologies resembling tubes of uniform width, are ideal for Xe/Kr separation. Finally, we compare our generated MOFs to several known structures (IRMOF-1, HKUST-1, ZIF-8, Pd-MOF, & MOF-505) and conclude that significantly improved materials remain to be synthesized. All crystal structure files are freely available for download and browsing in an online database.
Article
Problem statement: Anthropogenic activities pollute the oceans with oil through land run off, vessels accidents, periodic tanker discharges and bilge discharges. Oil spills are environmental disasters that impact human, plants and wild life including birds, fish and mammals. Approach: In this study, the International Guidelines for Preventing Oils Spills and Response to Disasters were reviewed and the characteristics of oil spills were discussed. The advantages and disadvantages of various oil spill response methods were evaluated. A comparative analysis were performed on the currently available remediation technologies using 10 evaluation criteria that included cost, efficiency, time, impact on wild life, reliability, level of difficulty, oil recovery, weather, effect on physical/chemical characteristics of oil and the need for further treatment. The advantages and disadvantages of each response method were used to determine the score assigned to that method. Result: There are many government regualtions for individual countries that serve as prevention mesures for oil spills in the offshore environment. They have to do with the design of equipment and machinery used in the offshore environment and performing the necessary safety inspections. The primary objectives of response to oil spill are: to prevent the spill from moving onto shore, reduce the impact on marine life and speed the degradation of any unrecovered oil. There are several physical, chemical, thermal and biological remediation technologies for oil spills including booms, skimmers, sorbents, dispersants, insitu burning and bioremediation. Each technique has its advantages and disadvantages and the choice of a particular technique will depend on: Type of oil, physical, biological and economical characteristics of the spill, location, weather and sea conditions, amount spilled and rate of spillage, depth of water column, time of the year and effectiveness of technique. Coclusion: Based on the comparative analysis, oil recovery with mechanical methods and the application of dispersants followed by bioremediation is the most effective response for marine oil spill.
Article
Face-capped octahedral clusters of the type [Re6Q8(CN)6]4- (Q = S, Se) are used to space apart partially hydrated Co2+ ions in extended solid frameworks, creating porous materials that display dramatic color changes upon exposure to certain organic solvents. The clusters react with cobaltous ions in aqueous solution to precipitate the new solid phases [Co2(H2O)4][Re6S8(CN)6]·10H2O (1), Cs2[Co(H2O)2][Re6S8(CN)6]·2H2O (2), and [Co(H2O)3]4[Co2(H2O)4][Re6Se8(CN)6]3·44H2O (3). The structures of 1·2H2O and 3 were determined by single-crystal X-ray analysis. The former consists of an expanded Prussian blue type framework with [Re6S8]2+ and [Co2(μ-OH2)2]4+ cluster cores occupying alternate metal ion sites, and features cubelike cages enclosing water-filled cavities approximately 258 Å3 in volume. The latter structure exhibits a network of Co2+ ions and [Co2(μ-OH2)2]4+ cores connected through [Re6Se8(CN)6]4- clusters, defining an array of one-dimensional channels with minimum internal diameters of 4.8 Å. A Rietveld refinement against X-ray powder diffraction data established compound 2 as isostructural to an analogous Fe-containing phase with a two-dimensional framework reminiscent of the Hoffman clathrates. Thermogravimetric analyses show that all three compounds are fully dehydrated by ca. 100 °C, with no further significant loss of mass below 500 °C. Upon exposure to diethyl ether vapor, the color of compounds 1 and 3 immediately changes from orange to an intense blue-violet or blue; other polar solvents induce somewhat different colors. These (reversible) changes are associated with the emergence of an envelope of new absorption features at wavelengths between 500 and 650 nm, and the magnitude of the response to a solvent can be estimated by measuring the relative intensity of a band with a maximum near 600 nm. We propose that the vapochromic response is due to solvent molecules entering the pores of the solid, where they disrupt the hydrogen-bonded water network, prompting the release of bound water from the [Co2(H2O)4]4+ clusters and conversion of their Co centers from octahedral to tetrahedral coordination. Significantly, this process does not destroy the three-dimensional connectivity in either structure, but rather creates a much more flexible framework that can expand to accommodate the incoming solvent molecules. Spectroscopic and magnetic data confirm the change in coordination geometry, and the trends in solvent responses (e.g., methanol < ethanol < n-propanol < i-propanol) are consistent with a decreased ability to support the bridging water ligands of the clusters as steric bulk increases. Size-selective sensing is demonstrated with methyl tert-butyl ether, which causes a color change in compound 3, but not in compound 1. X-ray powder diffraction experiments indicate that the vapochromic response in both compounds is affiliated with a reversible change in the bulk crystal structure of the material. Variable-temperature magnetic susceptibility data for compound 1 suggest a weak antiferromagnetic coupling interaction between the water-bridged Co2+ ions of the dinuclear cluster units. Finally, a simple chemical sensing device employing these solids is described, along with some properties relevant to its function.
Article
Several archetypical metal organic frameworks (MOFs), namely, HKUST-1, ZIF-8, MIL-100(Al), MIL-53(Al), and NH2-MIL-53(Al), were synthesized via anodic dissolution in an electrochemical cell. The influence of different reaction parameters such as solvent, electrolyte, voltage–current density, and temperature on the synthesis yield and textural properties of the MOFs obtained was investigated. The characterization of the samples involved X-ray diffraction, gas adsorption, atomic force microscopy, diffuse reflectance infrared Fourier transform spectroscopy, and scanning electron microscopy. In the present article, we demonstrate that electrochemical synthesis is a robust method offering additional degrees of freedom in the synthesis of these porous materials. The main advantages are the shorter synthesis time, the milder conditions, the facile synthesis of MOF nanoparticles, the morphology tuning and the high Faraday efficiencies. The synthesized MIL-53 and NH2-MIL-53 samples exhibit suppressed framework flexibility compared to samples synthesized solvothermally.
Article
A systematic investigation of the effects of linker polarizability on the adsorption properties of weakly interacting gases (N2, Ar, Kr, and Xe) is reported. Experimental and simulated adsorption properties were measured for a complete isoreticular series of monohalogenated metal–organic frameworks (MOFs). Variations on IRMOF-2, in which one linker hydrogen is replaced with −F, −Cl, −Br, or −I, comprise the series. Both experimental and simulated results indicate that increasing linker polarizability correlates with increased gas uptake. Evidence of increased adsorbate interaction with increased linker polarizability is also observed in the Kr/N2, Xe/N2, and Xe/Kr selectivity data and in isosteric heats of adsorption. Unexpectedly, comparison between simulated and experimental isotherms reveals that the agreement between the two improves with the size of the adsorbate, with essentially identical results for Xe. This is apparently due to the creation of regions inaccessible to any of the noble gases as a result of halogen functionalization. Simulated adsorption isotherms are also reported for radon, which is difficult to measure experimentally due to its radioactivity.
Article
Two vanadium diphosphonates were hydrothermally synthesized in a dilute medium. (VIVO)2(VIVO(H2O))4{O3P–CH2–PO3}4(NH4)4(H2O)4 (MIL-2) crystallizes in the orthorhombic system (space group Pmmn (no. 59)) with lattice parameters of a=13.512(7) Å, b=16.194(8) Å, c=4.677(2) Å, V=1023.4(9) Å3 and Z=1. Its structure, built up from corner-shared VIVO5 square pyramids, VIVO5(H2O) octahedra and PO3CH2PO3 tetrahedral units, is characterized by the presence of 14-membered tunnels inserting ammonium cations and water molecules. (VVO(H2O))(VIVO)O{O3P–(CH2)2–PO3}(NH4) (MIL-3) crystallizes in the P−1 (no. 2) triclinic space group with lattice parameters of a=7.4548(3) Å, b=8.0825(3) Å, c=10.1660(4) Å, α=75.244(1)°, β=68.883(1)°, γ=80.648(1)°, V=550.85(6) Å3 and Z=2. Its structure is built up from chains of VIVO6 octahedra and tetrahedral groups related by VVO4(H2O) square pyramids. The dehydration and the stability of both compounds are followed by thermodiffractometric measurements.
Article
The adsorption/desorption of up to 0.75 g of water vapour per g of the porous MOFs 3D-{M3O(X)(H2O)2[btc]2·nH2O}, MIL-100 (M = Al, Fe; X = OH, F, btc = benzene-1,3,5-tricarboxylate, trimesate), occurs at small relative pressures of p/p0 < 0.4 and a comparatively small hysteresis. Together with very good cycle stability, these properties render both MIL-100(Al and Fe) very suitable candidates for thermally driven heat pumps or adsorption chillers.
Article
The use of metal–organic frameworks (MOFs) for the encapsulation of different active entities is thoroughly reviewed. Either by following ship in a bottle or bottle around a ship approaches, active species can be encapsulated in the porous framework of different MOFs. Encapsulated species vary from polymers to organometallics and from polyoxometalates to metal nanoparticles and metal oxides. The main advantages and limitations of the use of MOFs together with the synthetic approaches followed are evaluated.