You are on page 1of 11

Materials Today  Volume 17, Number 3  April 2014

RESEARCH

RESEARCH: Review

Composites of metalorganic
frameworks: Preparation and application
in adsorption
Imteaz Ahmed and Sung Hwa Jhung*
Department of Chemistry and Green-Nano Materials Research Center, Kyungpook National University, Daegu 702-701, Republic of Korea

Metalorganic frameworks (MOFs) are one of the most discussed materials of the last decade. Their
extraordinary porosity and functionality from metals and organic linkers make them one of the most
promising materials for a vast array of applications. The easy tunability of their pore size and shape from
the micro- to meso-scale, by changing the connectivity of the inorganic moiety and the nature of the
organic linkers, makes these materials special. Moreover, by combining with other suitable materials,
the properties of MOFs can be improved further for enhanced functionality/stability, ease of preparation
and selectivity of operation. In this review, various methods and paths for the preparation of composites
are discussed, especially for those which have been successfully applied to gas and liquid phase
adsorptions. In the second part of this paper, several applications in adsorptive processes are discussed.
Introduction
Remarkable progress on porous materials has been achieved
because of developments in mesoporous materials [1,2] and
metalorganic frameworks (MOFs) [35], which are one of the
most important and rapidly growing groups of porous materials.
MOF materials are composed of metal ions (or clusters) and
coordinating linkers which impart high porosity to the MOF
structures. The particular interest in MOF materials is due to the
easy tunability of their pore size and shape from a microporous to a
mesoporous scale, by changing the connectivity of the inorganic
moiety and the nature of the organic linkers. Recently, MOF
materials have been used in many applications including gas
adsorption/storage, separation, catalysis, adsorption of organic
molecules, drug delivery, luminescence, electrode materials, carriers for nanomaterials, magnetism, polymerization, imaging,
membranes and so on [311]. Their potential increases every year
due to the easy modification of MOFs, which make them a
prominent group of materials in a vast number of published
reports.
As mentioned, MOFs show very promising physical and chemical properties for various applications; moreover, their properties can be further improved by several means. Some of these
are grafting active groups [12], changing organic linkers [13],
*Corresponding author. Jhung, S.H. (sung@knu.ac.kr)

136

impregnating suitable active materials [14], postsynthetic ligand


and ion exchange [15], making composites with suitable materials
[1620], and so on. A composite is a multi-component material
with multiple phases which has at least one continuous phase [21].
Recently, MOF composites have been receiving tremendous interest due to their various applications, including adsorption. This is a
relatively new concept and several reports have been published
recently with successful syntheses and promising applications of
MOF composites. By composing MOFs with suitable materials, the
synthesis kinetics [22], morphology [15,23], physicochemical
properties [24,25], stability [2628] and potential applications
[24,25] can be largely improved.
Adsorption receives more attention every year for its superiority
to other techniques in the removal of hazardous materials, purification of fuels or water and storage of gases such as hydrogen
and methane. Because adsorption can be carried out at low temperatures, it is less energy intensive and hence of comparatively
low cost. It also has the advantages of a wide range of applications,
simplicity of design, easy operation, and low harmful secondary
products. MOF-type materials are very well known for their applicability in adsorption processes due to their high surface area with
adequate pore openings [29,30]. Additionally, MOFs contain special and specific chemical functionalities which are useful to
selectively adsorb some species, and these functionalities are
one of the great properties of MOFs. Recently MOF composite

1369-7021/06/$ - see front matter 2014 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.mattod.2014.03.002

materials have been used to improve functionality and porosity


and, hence, improve the performance of adsorbents for their
selectivity or adsorption capacity. These materials have been
applied to both gas/vapor phase and liquid phase adsorptions.
Gas phase adsorptions are carried out mainly to remove toxic
gases. There are a large number of reports showing high adsorption
capacity for the adsorption of gases over MOFs. Recently, MOF
composites have also been used for gas phase adsorption to
improve the capacity and selectivity of adsorbents [30]. A large
number of studies have shown promising gas storage capabilities
with MOF composites which are also based on the adsorptive
phenomenon [3133]. By imparting special functionality, selective adsorption of specific gases can be achieved. Therefore, it is
very important to improve the selectivity of MOF materials by
combining them with suitable materials for removal, storage and
separation of gaseous materials. Several toxic gases such as H2S,
NH3 and NO2 could be adsorbed and hence separated by, for
example, graphite oxide/MOF composites [24,25,34,35]. MOF
composites have also been used for gas storage. A few reports have
been published for hydrogen storage using carbon nanotube
(CNT)/MOF composites, and in some cases, metals were also
loaded to these composites to improve the storage capacity
[16,36].
Liquid phase adsorptions have also been widely discussed and
applied to the various fields of MOF composites. Presently, fuel
processing is an important area of research, and every year MOFs
are being used to purify fuels or to remove harmful components
from fuels. In most cases, organic compounds are removed or
separated through MOFs using their selective interaction behavior. One of the most discussed adsorption processes is the adsorptive removal of sulfur or nitrogen containing compounds from
fossil fuels as shown by a few recent review articles [37,38]. Besides
these, water purification using MOFs is also gaining interest every
year. Understanding their true potential, researchers are turning
their interests to the adsorptive nature of MOF materials, and a vast
number of publications related to MOFs have been published in
the last decade. The opportunity to study pure MOFs has become
narrower due to the broadening field, and hence, researchers now
are experimenting with modified MOF materials with plenty of
room for studies on MOF composites.
In this review, we will discuss various methods and aspects for
the synthesis of MOF composites. The improvements in MOFs
made by combining with different materials will also be discussed.
Applications of MOF composites in adsorption will be explained
through the enhancement of porosity (surface area and pore
volume) or improved interaction between adsorbents and adsorbates (by acidbase interaction or p-complexation, and so on).

Discussion
Preparation of MOF composites
Several methods and approaches have been applied for the synthesis of MOF composites. In general, there are two basic types of
MOF composites according to the formation of the composite [39].
In MOF composites, MOFs can be in both the discontinuous phase
and the continuous phase. However, the first type of composites
(discontinuous MOFs) is not commonly used for application in
adsorption and are prepared mainly for purposes such as composites that require different sizes/shapes and that demand easy

RESEARCH

handling. These include monoliths, beads, fibers, membranes/


films, and so on. Examples of these materials are HKUST-1 monoliths [40], ZIF-8/PVP (polyvinylpyrrolidone) fiber mats [41], composite membranes of MOF-5 and ZIF-8 [42,43], and HKUST-1/PAM
(polyacrylamide) beads [27]. The second type of MOF composites
is the well-known MOF composites with continuous MOFs for
applications such as adsorption, and our discussion will be focused
on these materials.
For the preparation of MOF composites where MOFs are in a
continuous phase, a wide variety of methods have been applied.
The methods can be generally categorized into a few major classes.
In one of these methods, MOF precursors are mixed with presynthesized composing materials and then the synthesis procedure is carried out. Graphite oxide/MOF (GO/MOF) is one of these
types of materials that has been widely reported for various applications [25,35,44,45]. Bandosz et al. published several reports on
GO/MOF composites prepared using this procedure and successfully applied them in gas phase adsorptive removal of toxic gases
such as NH3, H2S, NO2, and so on [25,35,44,45]. In these composites, the GO layers are somewhat separated and stacked inbetween the planar cage structure of the MOFs (Fig. 1). In some
cases, for this type of synthesis, the particles of the composing
materials can be trapped in the pores of the MOF and hence can be
stabilized or immobilized. This concept is known as bottle around
ship (BAS) and is shown in Fig. 2(a). In the BAS method, large
particles of the composing materials are immobilized inside the
cage of the MOF, which is only possible by in situ synthesis of the
MOF. The MOF is built from the precursors around the composing
materials, and after complete formation of the MOF, the composing materials are trapped inside the pores. Several reports about the
encapsulation of polyoxometalates (POM) in MOFs have also been
published [22,46,47].
Liu, Su and coworkers also prepared several MOF composites
(the preparation method was very similar to the BAS method),
named NENU-n, from the in situ synthesis of MOFs (HKUST-1) in
the presence of POMs. Fig. 3 shows the structure of two pores of
NENU-n, which was unambiguously obtained by single crystal Xray analysis [48]. It was later demonstrated that the NENU-11 (a
sodalite-type porous MOF with a POM template as shown in
Fig. 3(c)) possessed a good adsorption/decomposition capability
for dimethyl methyl phosphonate (from 6 molecules in MOF-5 to
15.5 molecules in NENU-11 per unit formula) [49].
In the other preparation method, the MOF composite is supposed to form from the precursors of the composing materials and
the preformed MOF. This method is called sometimes the ship in a
bottle (SIB) method, as explained by Gascon et al. [19] (Fig. 2(b)).
In this case, the composing material forms inside the cages of the
MOF and the diffusion of the precursors is usually carried out in
the cages by a solvent system. Due to the increased size, the
particles stay stable inside of the cages. In most cases, the precursors of the composing materials are added through solutions
and diffused into the pores. Then the composite is formed by
stabilizing the composing moiety inside the pores (of the MOFs) by
means of chemical or thermal methods. One of the best examples
of this, shown in Fig. 4, is direct encapsulation of different porphyrines into ZMOFs, and this composite was successfully applied
in catalytic oxidation [50]. Although SIB techniques are widely
known for zeolite composites [51], their study within the MOF
137

RESEARCH: Review

Materials Today  Volume 17, Number 3  April 2014

RESEARCH

Materials Today  Volume 17, Number 3  April 2014

RESEARCH: Review
FIGURE 1

Schematic comparison of the coordination between GO carbon layers and MOF units for different types of MOF networks (MOFs are MOF-5, HKUST-1 and
MIL-100). Adapted from Ref. [44].

field has recently started and there are vast opportunities for
utilizing this encapsulation method for MOFs.
If the solubility of composing materials is not an issue for
applications, the composing procedure may be carried out using
postsynthesis methods. Simple loading, or impregnation, is one
widely applied method of this type. Jhungs group loaded soluble
salts and polyoxometalates (POM) and used them for liquid phase
adsorptions [52,53]. One of the best examples of these materials is
the loaded metal oxides or insolubilized materials inside the pores,
achieved through chemical means (e.g. changing the oxidation
state, and so on). This method has been used for the preparation of
MOFs composed of metal oxide (e.g. Cu2O) [54]. There have also
been reports on the insolubilization of composing materials in
which precipitates were formed from the solution by changing the
oxidation state of the soluble salts [55,56]. Lu et al. [57] reported a
controlled encapsulation strategy that enables surfactant-capped
nanostructured objects of various sizes, shapes and compositions
to be enshrouded by a zeolitic imidazolate framework (ZIF-8). This
strategy worked for the incorporation of several types of functional
nanoparticles into the MOF pores. The incorporated nanoparticles
are well dispersed and fully confined within the ZIF-8 crystals.
They used polyvinylpyrrolidone (PVP) as a surfactant, and the
procedure is shown in Fig. 5. Canioni et al. [58] synthesized
138

POM-inserted MIL-100(Fe) with various methods including the


in situ synthesis of MIL-100(Fe) and showed the stability of the
composite in water.
An important type of MOF composite is the metal/MOF composite. Several studies on MOFs loaded with different metal particles are available in the literature. Most of the metals used are dblock transition metals which demonstrate nice functionality in
terms of their complex forming capability and acid-base interactions, thanks to their d-block electrons. There have been reports on
Au, Ag, Pd, Ni, Ga, Co, Pt, Ru and some other metals being
composed with MOFs [12,5966]. It is also possible to use more
than one type of metal atom in composites [67]. The encapsulation
method differs widely for different metals and MOF pairs. Ishida
et al. used a solid grinding technique to prepare Au-MOF composites by grinding the Au particles along with MIL-53(Al), MOF-5
and HKUST-1 [59,60]. Other techniques used for metal/MOF
composites are direct encapsulation [61], impregnation
[12,62,63], gas phase infiltration [64,65], coprecipitation [66],
and so on.
Some composing materials have templating effects for the
structures of MOFs. Graphene oxide is one of such example.
According to Jahan et al. [23], graphene can be decorated with
functional groups on either side of its basal plane, giving rise to a

RESEARCH

RESEARCH: Review

Materials Today  Volume 17, Number 3  April 2014

FIGURE 2

Composite formation by encapsulation with different strategies: (a) bottle around ship approach in which the porous host materials are produced around
the encapsulated materials; (b) ship in a bottle in which guest moieties are prepared in the presence of porous host and; and (c) the composite prepared
by any of the two methods. Adapted from Ref. [19].

bifunctional nanoscale building block that can undergo face-toface assembly. They demonstrated that benzoic acid-functionalized graphene (BFG) can act as a structure-directing agent (SDA) in
influencing the crystal growth of an MOF. Moreover, BFG can also
be imparted into the MOF structure as a linker (Fig. 6). New
electrical properties were also imparted into the MOF by intercalation with graphene oxide while the original MOF is insulating.
Bajpe et al. showed that a POM/HKUST-1 composite can remarkably improve the kinetics of the synthesis. The interaction

between the two materials even reduced the energy requirement


for the synthesis as shown by the fact that room temperature
synthesis was possible in a short period of time [22,68,69]. Another
study by Bandoszs group showed not only the enhancement of
the properties of virgin MOFs, but also the distinct properties of
the composites compared to the parent materials [45]. Interaction
between the metal sites of MOFs and the oxygen groups of GO
creates new pores in the interfaces between the carbon layers and
the MOF units. There is a report where activated carbon (AC)
139

RESEARCH

Materials Today  Volume 17, Number 3  April 2014

RESEARCH: Review
FIGURE 3

(a) Components of the crystal structures of NENU-n (n = 16): (a) a cube of five truncated-octahedral cages sharing square faces; (b) the pore A
accommodating the Keggin polyanion, Cu2+ cluster (blue polyhedra), Keggin polyanion (polyhedral in yellow circle), C (gray), and O (red); (c) (1) a Keggin
type of POM; (2) three-connected node and hexagonal face (blue) defined by a BTC ligand linked to six adjacent Cu2+ ions; (3) secondary building unit (SBU)
and square face (green) defined by four Cu2+ ions; and (4) cube of eight sodalite-like truncated octahedral cages sharing square faces. Reproduced with
permission from Refs. [48,49]. Copyright 2009 and 2011 American Chemical Society, respectively.

generated a large proportion of voids in MIL-101 and hence


improved the porosity to enhance the hydrogen storage of MIL101 [70]. In this way, the functionality and applicability of MOFs
can be improved to a great extent.

we will discuss the applications of MOF composites in different


areas through adsorption processes. All the adsorption techniques
can be divided into two general categories: (a) gas phase adsorption and (b) liquid phase adsorption.

Applications of MOF composites in adsorption

Gas phase adsorption

The applications of MOF composites not only include all those


of individual MOFs, but also include additional fields. Their
applicability includes adsorption (both in liquid and gas phases),
separation, purification, catalysis, drug delivery, structure modification and so on. The potential applications of MOF composites
are summarized in Fig. 7. The mechanisms for the applications of
MOF composites may be similar to or different from that of virgin
MOFs. Moreover, adsorption is observed in many applications and
hence it is important to understand the applicability of MOF
composites through adsorption phenomena. In the next part,

All over the world, toxic gases are an alarming environmental


concern. One of the best ways to remove these gases is by utilizing
an adsorption process. There have been a vast amount of reports on
the adsorptive removal of harmful gases. Recently, the use of MOFs
has been reported for the removal of gases through their enhanced
surface area and distinct interactions [9,29,7176]. Additionally, gas
storage has also been gaining wide attention every year, and MOFs
have found their way into gas storage applications due to their
extraordinary properties, including a high surface area and pore
volume. Recently, MOF composites, similar to the virgin MOFs,

140

FIGURE 4

(a) Eight-coordinated InN4 represented as a tetrahedral building unit; (b)


[H2TMPyP]4+ porphyrin; and (c) crystal structure of rho-ZMOF (left), hydrogen
atoms omitted for clarity, and schematic representation of the [H2TMPyP]4+
porphyrin ring enclosed in rho-ZMOF a-cage (right, drawn to scale).
Reproduced with permission from Ref. [50]. Copyright 2008 American
Chemical Society.

RESEARCH

have been used for gas phase adsorption/storage because of their


improved functionality, stability, selectivity and capacity.
Table 1 shows the applications of MOF composites in gas phase
adsorption with detailed improvements and plausible mechanisms. There are many applications for gas removal/storage through
adsorption. In some studies, graphite oxide combined with different MOFs has been used to improve adsorption capacities, as
well as other properties. It was found that a GO/MOF composite
demonstrates a strong dispersive force which enhances the retention of small molecules such as ammonia. The results showed that
composites of MOF-5 and GO possess a synergetic effect in the
adsorption of ammonia, compared to the individual materials,
and hence the composite material demonstrates good performance as an ammonia adsorbent [24]. The adsorption capacity
was improved from 6 mg/g (pristine MOF) to 82 mg/g over the
composite surface. Ammonia adsorption was also improved in a
GO/HKUST-1 composite. It was observed that the GO/HKUST-1
composite had a maximum adsorption of NH3 compared with
other materials [24]. The improved adsorption arises from the
development of new porosity at the interface between the GO
and the MOF units. It was also reported that different adsorption
sites are introduced upon combining the two materials (MOF and
GO) [25]. NO2 and H2S removals have also been reported with GO/
HKUST-1 and GO/MIL-100(Fe), respectively [35,44].
Gas storage through gas phase adsorption also has been widely
studied using MOF composites. It was observed that MOF-5 has a
good storage capability for H2; however, this material is not stable
when exposed to atmospheric moisture. Its stability in moisture

FIGURE 5

Scheme of the controlled encapsulation of nanoparticles in ZIF-8 crystals. Through surface modification with surfactant PVP, nanoparticles of various sizes,
shapes and compositions can be encapsulated in a well-dispersed fashion in ZIF-8 crystals, themselves formed by assembling zinc ions with imidazolate
ligands. Reproduced with permission from Ref. [57]. Copyright 2012 Nature Publishing Group.
141

RESEARCH: Review

Materials Today  Volume 17, Number 3  April 2014

RESEARCH

Materials Today  Volume 17, Number 3  April 2014

RESEARCH: Review
FIGURE 6

Schematic of proposed bonding between (a) MOF and graphene and (b) MOF and BFG. Reproduced with permission from Ref. [23]. Copyright 2010
American Chemical Society.

was improved by combining with multi-walled carbon nanotubes


(MWCNTs), and the composite was denoted as MOFMC (MOF
multi-walled CNT) [16]. This material has an interpenetrating
structure and hierarchical nanopores that improve the storage
capability of the composite. The MOFMC is also mesoporous,
which further improves the storage capacity (2.02 wt% at 77 K
and 1 bar). Another report was published using Pt-loaded
MWCNT/MOF-5 with a great improvement in H2 storage (from
1.20 wt% in virgin MOF to 1.89 wt% in the composite) [77]. Yang
et al. also published a report on H2 storage using MWCNTs
composed with MOF-5 [78]. They reported a 50% improvement

in storage capacity compared to virgin MOF-5. They measured the


H2 storage capacity for various conditions, and in every case, they
found improved adsorption results compared to the virgin MOF-5.
Activated carbon (AC) composed with MOF was also reported for
H2 storage [70]. Up to 10.1 wt% of H2 storage has been reported
with the AC/MIL-101 material where only 0.63% of AC was
incorporated. A theoretical study showed that the hydrogen storage capacity is improved to 6.3 wt% or 42 g L 1 at 100 bar and
243 K over Li-decorated IRMOF-10 with Li-coated fullerenes (C60)
[79]. POM/MOF composites have also been used in hydrogen
storage. The introduction of POMs and lithium ions in MOFs leads
to strong hydrogen adsorption [80].
CO2 capture is one of the important applications of MOF
composites, and an MWCNT (multi-walled carbon nanotube)/
MIL-101 composite has been used to improve CO2 capture [81].
The MWCNT/MIL-101 has the same morphology as the virgin
MIL-101, but a 60% improvement in CO2 capture has been noticed
(from 0.84 to 1.35 mmol g 1). The improvement of CO2 capture
with the composite is attributable to the increased micropore
volume of the MIL-101 through the MWCNTs. An MOF composite, polyethyleneimine loaded MIL-101 (PIM/MIL-101), showed
improved CO2 capture although the porosity of the MIL-101 was
drastically reduced by the PIM [82] due to the strong interaction
between the alkylamine groups and CO2.

Liquid phase adsorption


FIGURE 7

Potential applications of MOF composites.


142

The most prominent liquid phase adsorption using MOFs is the


purification of petroleum by removing NCCs (nitrogen containing
compounds) and SCCs (sulfur containing compounds) from fuels.

Gas adsorption properties of MOF composites.


MOF

Composing material

Application

Mechanism

Experiment conditions

Result a

Ref.

MOF-5

Multiwalled carbon
nanotube (MWCNT)

H2 storage through
adsorption

Improved surface area by improving nanopores


and improved stability through interaction

77 K temp. and up to 1 bar pressure


were used

[16]

MOF-5

Graphite oxide (GO)

Adsorption of NH3

Improved adsorption sites and additional porosity


by composing the MOF and GO

HKUST-1

Graphite oxide (GO)

Adsorption of NH3

Improved adsorption sites and additional porosity


by composing the MOF and GO

HKUST-1

Graphite oxide

Adsorption of NH3, H2S and


NO2

Synergetic effects between HKUST-1 and GO

NH3 adsorption was done by diluting


with air in a breakthrough analyzing
apparatus at ordinary temp. and
pressure
At a temperature of 2575 8C.
Pressure was maintained at
100120 kPa
Flow of the gases diluted by air was
used for breakthrough experiments

H2 storage was improved from


1.2 wt% to 1.52 wt% in the
composite
Adsorption of NH3 was improved
from 6 mg/g to 82 mg/g

MIL-101

Activated carbon (AC)

H2 storage through
adsorption

Reduction of pore size and improvement of pore


volume. Additional micropores were formed

77.4 K temp. and up to 600 kPa


pressure were used

MOF-5

H2 storage through
adsorption

Secured porosity which acted as H2 spillover


receptor

77 K temp. and up to 1 bar pressure


was used

IRMOF-10

Multiwalled carbon
nanotube (MWCNT)
loaded with Pt
Li coated fullerene

H2 storage application

HKUST-1

Polyoxometalate (POM)

H2 adsorption

Electrostatic charge quadrupole and dipole


interactions between H2 and Li
Interaction between H2 molecule and O atom of
POM

243 K temp. and 100 bar pressure


were studied
77 and 298 K temp. along with
various pressures up to 100 bar

MIL-101

Multiwalled carbon
nano-tube (MWCNT)

CO2 storage through


adsorption

Increase of micropore volume of the MOF by


MWCNT incorporation

298 K temp. and 10 bar pressure

MIL-101

Polyethyleneimine (PEI)

CO2 storage by improving


the selectivity

Selectivity of CO2 adsorption improved due to


strong interaction between the alkylamine groups
of PEI

298 K and 323 K temp. along with


0.15 bar and 1.0 bar were applied

[24]

Adsorption improved from


6.2 mmol/g to 8 mmol/g on average

[25]

Around 10% and 30% improvement


for NH3 and H2S. Little improvement
was found for NO2.
Nearly 58% improvement of H2
storage was observed in the
composite
H2 storage was improved from
1.2 wt% to 1.89 wt% in the
composite
An improved H2 uptake up to
6.3 wt%
Slight improvement in H2
adsorption. Basically the mechanism
was studied.
60% increase (from 0.84 to
1.34 mmol/g) in CO2 uptake was
found
140% more amount of CO2 intake
was observed

[45]

Materials Today  Volume 17, Number 3  April 2014

TABLE 1

[70]

[77]

[79]
[80]

[81]

[82]

Improvements compared to the corresponding virgin MOFs.

RESEARCH

143
RESEARCH: Review

RESEARCH

RESEARCH: Review
FIGURE 8

(a) Proposed mechanism for the adsorption of benzothiophene (BT) over


Cu+ site; and (b) adsorption isotherms for BT over MIL-47 and CuCl2/MIL-47.
Adapted from Ref. [55].

These materials have to be removed before commercial use due to


environmental problems and catalyst poisoning. As in gas-phase
removal, MOF composites are also being widely applied to the
removal of contaminates from fossil fuels.
MOFs, not in the form of composites, have a large and accessible
porosity inside their cage; therefore, this empty space can be
utilized by suitable entities for adsorptions [37,38]. Adsorptive
removal of obnoxious materials can be further improved by combining MOFs with suitable materials. Khan et al. showed improved
adsorption of benzothiophene (BT) over a polyoxometallate
(POM)-loaded porous HKUST-1 [52] to understand the effect of
POM on the adsorption/removal of BT. The maximum adsorption
capacity was improved up to 26% by the loading although the
porosity of the composite decreased. Therefore, it was concluded
that the chemical interaction, through an acid-base effect
(between the acidic POM and slightly basic BT), improved the
adsorption.
As reported, several metal ions such as Cu+, Ag+, Pd2+ and Pt2+
show p-complexation behavior and hence can improve the
adsorptive removal of SCCs. [83,84]. Yang and coworkers have
developed adsorbents based on the p-complex for desulfurization
[83]. As shown in Fig. 8, Khan et al. [55] demonstrated a remarkable
adsorption of BT over CuCl2-loaded MIL-47 through p-complexation. Cu2+ sites are not useful as p-complexating sites; however,
144

Materials Today  Volume 17, Number 3  April 2014

upon loading, the Cu2+ sites readily turn to Cu+ sites by reacting
with the V3+ of MIL-47 and hence a p-complexation site is automatically created through the loading. By contrast, other analogous MOFs [85] having a similar structure, such as MIL-53(Al) and
MIL-53(Cr), do not show a reduction ability to form Cu+ because of
the lack of reducibility of Al3+ and Cr3+, respectively [56]. In
another study the same group reported on BT adsorption using
Cu2O/MIL-100(Fe), presenting improved performance [54]. Very
recently, they showed that ionic liquid/MOFs are effective in
removing SCCs compared to pristine MOFs [86]. Ahmed et al.
showed the improved adsorption of quinoline (QUI), one of the
typical NCCs, over POM-loaded MIL-101 [53]. It was found that a
1% loading of POM could improve the adsorption of QUI up to
20%. They showed that the adsorption of a neutral NCC, indole,
could not be improved by the loading of POM, although basic QUI
adsorption was nicely improved through an acid-base interaction.
Very recently, it was also reported that GO/MIL-101 was very
effective in removing NCCs partly because of the improved porosity of the adsorbent [87].
MOF composites have also been used for water purification,
similar to that performed with virgin MOFs [8891]. GO/HKUST1 composites were successfully applied in the removal of methylene blue (MB) with improved adsorptive performance. The study
showed a maximum adsorption capacity of 183 mg/g for MB
in water [92]. It was found that the efficiency of the adsorption
was high for a low concentration of MB in water for both
virgin HKUST-1 and GO/HKUST-1 composites with fast kinetics.
However, in the case of a high MB concentration, the efficiency
of HKUST-1 decreased below 50% of the original value, while
GO/HKUST-1 still removed the MB up to 80% within 30 min,
which indicates the much better adsorption ability of the
composite compared to the virgin MOF alone. Another report
shows the effective removal of several toxic chemicals (NH3, H2S,
and so on) from effluent water with Zr(OH)4/HKUST-1 composites [93].
Besides these applications, there are other adsorptive applications of MOF composites, such as separation with magnetic properties [94], drug delivery [95], static phase material for liquid
chromatography [96] and so on. In particular, drug delivery is
currently a widely studied topic in the field of MOFs [9799]. In
this case, the concern is more probably desorption rather than
adsorption. In many cases, the steady, slow and targeted release of
a drug in the body is essential for efficient therapeutics. MOF
composites have promising and prospective applications in this
field. Several studies on drug delivery have recently been reported
using MOF composites [95]. Ke et al. synthesized HKUST-1 with
incorporated Fe3O4 nanorods which could be potentially used for
drug delivery (Fig. 9(a)) [95]. This kind of material provides an
efficient platform for a new strategy of delivering an imaging
contrast agent and an anticancer drug by postsynthetic modifications of a highly porous nanoscale metalorganic framework
(NMOF). Another report on drug delivery was published by Taylor-Pashow et al. [100]: they demonstrated the synthesis of an
ironcarboxylate NMOF with a MIL-101 structure, and post-modification was performed by loading an organic fluorophore and an
anticancer drug via covalent modifications of the as-synthesized
nanoparticles for targeted drug delivery, magnetic resonance imaging and magnetic separation (Fig. 9(b)) [99].

RESEARCH

RESEARCH: Review

Materials Today  Volume 17, Number 3  April 2014

FIGURE 9

(a) Illustration of the synthesis and action of Fe3O4/HKUST-1 nanocomposites adapted from Ref. [95]; and (b) postsynthetic modifications of iron-carboxylate
nanoscale MOFs for imaging and drug delivery. Reproduced with permission from Ref. [100]. Copyright 2008 American Chemical Society.

Summary and conclusion


For over a decade, individual MOFs have been serving the scientific
community with their remarkable properties and wide applications. Additionally, composite materials are currently opening the
door for new opportunities and prospects by extending the applicability of MOFs. There are many different ways to prepare an MOF
composite, such as encapsulation, impregnation, infiltration, solid
grinding, coprecipitation, and so on. Different types of preparation method can produce different properties in the MOF composites and hence can improve the applicability of the composite
materials. One of the major improvements in MOF composites is
the enhanced porosity and the subsequently improved adsorption
capability. Other composites show special functionality and
improved practical applications. Some composites show structural
modifications and enhanced kinetics in the synthesis of MOFs and
have imparted new properties which improve the versatility of the
materials in different aspects.
The most utilized application of MOFs is adsorption, which can
be carried out for different purposes. These include the removal of
harmful materials from liquid/gas phases, gas storage, separation,
purification, catalysis, drug delivery and so on. In every case, the
applicability and performance of an MOF can be improved by
forming a composite with suitable MOFs, mainly because of the

imparted functionality and enhanced porosity. Specially, in such


cases where selective adsorption or separation is necessary, MOF
composites can show spectacular effects. MOF composites can be
applied in every field where MOFs can be applied, and additional
new possibilities can also be found in applications where individual MOFs cannot be used. With suitable functional materials,
limitless composites are bound only by ones imagination, and
can be prepared for new applications. In the future, there will be
many more requirements on the specific capabilities of such
materials, and these can be fulfilled by MOF composites. By
combining MOFs with suitable materials, the functionality, porosity, ease of synthesis, and thermal/magnetic/electric properties
can be improved to meet specific requirements. Therefore, it can
be concluded that though the exploration of these composite
materials has only just begun, the prospects are vast and will
continue to increase.

Acknowledgements
This research was supported by Basic Science Research Program
through the National Research Foundation of Korea (NRF) grant
funded by the Korea government (MSIP) (grant number:
2013R1A2A2A01007176). Authors express their sincere thanks to
Dr. N.A. Khan and Dr. Z. Hasan for their helpful discussion.
145

RESEARCH

References

RESEARCH: Review

[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]

146

A. Vinu, et al. J. Porous Mater. 13 (2006) 379383.


K. Ariga, et al. Bull. Chem. Soc. Jpn. 85 (2012) 132.
G. Ferey, Chem. Soc. Rev. 37 (2008) 191214.
S. Kitagawa, R. Kitaura, S.-I. Noro, Angew. Chem. Int. Ed. 43 (2004) 23342375.
O.M. Yaghi, et al. Nature 423 (2003) 705714.
ller, Chem. Soc. Rev. 38 (2009) 12841293.
A.U. Czaja, N. Trukhan, U. Mu
S.T. Meek, J.A. Greathouse, M.D. Allendorf, Adv. Mater. 23 (2011) 249267.
H. Wu, et al. Chem. Rev. 112 (2012) 836868.
J.-R. Li, J. Sculley, H.-C. Zhou, Chem. Rev. 112 (2012) 869932.
C. Sanchez, et al. Chem. Soc. Rev. 40 (2011) 696753.
M. Shah, et al. Ind. Eng. Chem. Res. 51 (2012) 21792199.
Y.K. Hwang, et al. Angew. Chem. Int. Ed. 47 (2008) 41444148.
M. Eddaoudi, et al. Science 295 (2002) 469472.
A.W. Thornton, et al. J. Am. Chem. Soc. 131 (2009) 1066210669.
M. Kim, et al. J. Am. Chem. Soc. 134 (2012) 1808218088.
S.J. Yang, et al. Chem. Mater. 21 (2009) 18931897.
C. Petit, T.J. Bandosz, Adv. Mater. 21 (2009) 47534757.
F. Yanyan, Y. Xiuping, Prog. Chem. (Chin.) 25 (2013) 221232.
iz, J. Gascon, F. Kapteijn, J. Mater. Chem. 22 (2012) 1010210118.
J. Juan-Alcan
D. Bradshaw, A. Garai, J. Huo, Chem. Soc. Rev. 41 (2012) 23442381.
W.J. Work, et al. Pure Appl. Chem. 76 (2004) 19852007.
L.H. Wee, et al. Chem. Commun. 46 (2010) 81868188.
M. Jahan, et al. J. Am. Chem. Soc. 132 (2010) 1448714495.
C. Petit, T.J. Bandosz, Adv. Funct. Mater. 20 (2010) 111118.
C. Petit, et al. Langmuir 27 (2011) 1304313051.
D. Mustafa, et al. Chem. Commun. 47 (2011) 80378039.
L.D. ONeill, H. Zhang, D. Bradshaw, J. Mater. Chem. 20 (2010) 57205726.
D.-D. Liang, et al. Adv. Synth. Catal. 353 (2011) 733742.
H. Furukawa, et al. Science 329 (2010) 424428.
J.-R. Li, R.J. Kuppler, H.-C. Zhou, Chem. Soc. Rev. 38 (2009) 14771504.
P. Nugent, et al. Nature 495 (2013) 8084.
S.M. Luzan, et al. Int. J. Hydrogen Energy 34 (2009) 97549759.
R. Pedicini, et al. Int. J. Hydrogen Energy 36 (2011) 90629068.
B. Levasseur, C. Petit, T.J. Bandosz, ACS Appl. Mater. Interfaces 2 (2010) 36063613.
C. Petit, B. Mendoza, T.J. Bandosz, ChemPhysChem 11 (2010) 36783684.
H. Jiang, et al. Int. J. Hydrogen Energy 38 (2013) 1095010955.
N.A. Khan, Z. Hasan, S.H. Jhung, J. Hazard. Mater. 244245 (2013) 444456.
N.A. Khan, Z. Hasan, S.H. Jhung, Adv. Porous Mater. 1 (2013) 91102.
N. Stock, S. Biswas, Chem. Rev. 112 (2012) 933969.
sgens, et al. J. Am. Ceram. Soc. 93 (2010) 24762479.
P. Ku
R. Ostermann, et al. Chem. Commun. 47 (2011) 442444.
E.V. Perez, et al. J. Membr. Sci. 328 (2009) 165173.
ez, J. Membr. Sci. 361 (2010) 2837.
M.J.C.ET-AL> Ordon
C. Petit, T.J. Bandosz, Adv. Funct. Mater. 21 (2011) 21082117.
C. Petit, T.J. Bandosz, Dalton Trans. 41 (2012) 40274035.
L.H. Wee, et al. Catal. Today 171 (2011) 275280.
iz, et al. J. Catal. 269 (2010) 229241.
J. Juan-Alcan
C.-Y. Sun, et al. J. Am. Chem. Soc. 131 (2009) 18831888.
F.-J. Ma, et al. J. Am. Chem. Soc. 133 (2011) 41784181.
M.H. Alkordi, et al. J. Am. Chem. Soc. 130 (2008) 1263912641.

Materials Today  Volume 17, Number 3  April 2014

[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]
[83]
[84]
[85]
[86]
[87]
[88]
[89]
[90]
[91]
[92]
[93]
[94]
[95]
[96]
[97]
[98]
[99]
[100]

A. Corma, H. Garcia, Eur. J. Inorg. Chem. (2004) 11431164.


N.A. Khan, S.H. Jhung, Fuel Process. Technol. 100 (2012) 4954.
I. Ahmed, et al. J. Hazard. Mater. 250251 (2013) 3744.
N.A. Khan, S.H. Jhung, J. Hazard. Mater. 237238 (2012) 180185.
N.A. Khan, S.H. Jhung, Angew. Chem. Int. Ed. 51 (2012) 11981201.
N.A. Khan, S.H. Jhung, J. Hazard. Mater. 260 (2013) 10501056.
G. Lu, et al. Nat. Chem. 4 (2012) 310316.
R. Canioni, et al. J. Mater. Chem. 21 (2011) 12261233.
M. T. Ishida, et al. Chem. Eur. J. 14 (2008) 84568460.
T. Ishida, N. Kawakita, T.A.M. Haruta, Gold Bull. 42 (2009) 267274.
K. Sugikawa, Y. Furukawa, K. Sada, Chem. Mater. 23 (2011) 31323134.
R.J.T. Houk, et al. Nano Lett. 9 (2009) 34133418.
B. Yuan, et al. Angew. Chem. Int. Ed. 49 (2010) 40544058.
ller, et al. Eur. J. Inorg. Chem. (2011) 18761887.
M. Mu
S. Hermes, et al. Chem. Mater. 19 (2007) 21682173.
S. Opelt, et al. Catal. Commun. 9 (2008) 12861290.
der, et al. Eur. J. Inorg. Chem. 21 (2009) 31313140.
F. Schro
S.R. Bajpe, et al. Chem. Eur. J. 16 (2010) 39263932.
S.R. Bajpe, et al. J. Mater. Chem. 21 (2011) 97689771.
P.B.S. Rallapalli, et al. J. Energy Res. 37 (2013) 746753.
G.W. Peterson, et al. J. Phys. Chem. C 113 (2009) 1390613917.
D. Britt, D. Tranchemontagne, O.M. Yaghi, Proc. Natl. Acad. Sci. U.S.A. 105
(2008) 1162321162.
J. Liu, et al. Chem. Soc. Rev. 41 (2012) 23082322.
K. Sumida, et al. Chem. Rev. 112 (2012) 724781.
A.R. Millward, O.M. Yaghi, J. Am. Chem. Soc. 127 (2005) 1799817999.
Z. Zhao, et al. Ind. Eng. Chem. Res. 50 (2011) 22542261.
S.J. Yang, et al. Int. J. Hydrogen Energy 35 (2010) 1306213067.
O. Landau, A. Rothschild, E. Zussman, Chem. Mater. 21 (2009) 911.
D. Rao, et al. Chem. Commun. 47 (2011) 76987700.
F. Ma, et al. Eur. J. Inorg. Chem. (2010) 37563761.
M. Anbia, V. Hoseini, Chem. Eng. J. 191 (2012) 326330.
Y. Lin, et al. Sci. Rep. 3 (2013) 18591865.
R.T. Yang, A.J. Hernandez-Maldonado, F.H. Yang, Science 301 (2003) 7981.
Z.Y. Zhang, et al. Appl. Catal. B: Environ. 82 (2008) 110.
S.H. Jhung, N.A. Khan, Z. Hasan, CrystEngComm 14 (2012) 70997109.
N.A. Khan, Z. Hasan, S.H. Jhung, Chem. Eur. J. 20 (2014) 376380.
Z. Hasan, J. Jeon, S.H. Jhung, J. Hazard. Mater. 209210 (2012) 151157.
I. Ahmed, N.A. Khan, S.H. Jhung, Inorg. Chem. 52 (2013) 1415514161.
E. Haque, et al. J. Hazard. Mater. 181 (2010) 535542.
E. Haque, J.W. Jun, S.H. Jhung, J. Hazard. Mater. 185 (2011) 507511.
E.Y. Park, et al. J. Nanosci. Nanotechnol. 13 (2013) 27892794.
L. Li, et al. J. Mater. Chem. A 1 (2013) 1029210299.
G.W. Peterson, et al. Ind. Eng. Chem. Res. 52 (2013) 54625469.
S.-H. Huo, X.-P. Yan, Analyst 137 (2012) 34453451.
F. Ke, et al. J. Mater. Chem. 21 (2011) 38433848.
R. Ameloot, et al. Eur. J. Inorg. Chem. (2010) 37353738.
P. Horcajada, et al. Nat. Mater. 9 (2010) 172178.
P. Horcajada, et al. J. Am. Chem. Soc. 130 (2008) 67746780.
P. Horcajada, et al. Chem. Rev. 112 (2012) 12321268.
K.M.L. Taylor-Pashow, et al. J. Am. Chem. Soc. 131 (2009) 1426114263.

You might also like