SlideShare a Scribd company logo
1 of 7
Download to read offline
This journal is ©The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 New J. Chem., 2015, 39, 931--937 | 931
Cite this: New J. Chem., 2015,
39, 931
Improved photocatalytic activity in a surfactant-
assisted synthesized Ti-containing MOF
photocatalyst under blue LED irradiation†
Sedigheh Abedi and Ali Morsali*
The Ti-incorporated metal–organic framework structure, MIL-125, its amine-functionalized form, MIL-NH2
and its surfactant-assisted version, SMIL-NH2 were synthesized via a solvothermal route. The samples were
characterized by XRD, SEM, EDS, UV-vis spectroscopy, and N2 adsorption–desorption measurements. In
the presence of a nonionic surfactant, cylindrical SMIL-NH2, with 50–250 nm particle size, was obtained.
The photocatalytic properties of the materials were evaluated by deoximation reaction. Compared with
MIL-NH2, the nanostructured SMIL-NH2 MOF has more than double the surface area and is able to
catalyze the regeneration of carbonyl compounds from various oximes at twice the rate in an energy-
efficient photocatalytic system. Furthermore, it has been successfully shown that this new nanostructured
photocatalyst is significantly durable under the reaction conditions and can be easily recycled after the
reaction, with significant retention of its photocatalytic activity.
Introduction
Today, the rapid everyday rise of waste associated with fossil
fuels is one of the major concerns of ecologists. On the other
hand, the rapid reduction in fossil sources of energy has led the
attention of many enthusiasts to developing ways of utilizing
carbon-neutral energy sources. Extensive efforts have been
made to replace these kinds of nonrenewable sources with
widely available, abundant solar energy. Photocatalytic systems
are one the most promising routes to achieve this worthy global
objective. Among them, heterogeneous photocatalytic systems,
using the most widely used TiO2-based photocatalysts, have
attracted much attention, due to their biological and chemical
inertness as well nontoxic and inexpensive characteristics.1–3
The unique and tunable architectural properties of metal–
organic frameworks (MOFs) have made them promising candidates
for heterogeneous catalysis applications.4,5
The extraordinarily high
surface areas, porosities and well-defined pore structures of these
materials are the cause of recent demands for their use as
catalysts.6–8
Recently, a variety of modified MOFs and their
applications in light harvesting9–12
, in driving various photo-
catalytic reactions such as H2 production,13–15
O2 production,16
and carbon dioxide reduction,17
in gas separation and storage
and in the catalytic transformation of organic molecules, have
been developed.18–20
The porous channels in MOF structures
can be used as photocatalysis sites. Theoretical calculations
show that MOFs are semiconductors or insulators with band
gaps between 1.0 and 5.5 eV, which can be altered by changing
the degree of conjugation in the ligands.21
Experimental results
show that the band gaps of IRMOF-type samples can be tuned
by varying the functionality of the linker.22
Among numerous scientific attempts to investigate the
unexplored catalytic properties of crystalline MOFs,7,23–30
the
development of robust MOFs with the potential for photo-
catalytic conversion of light energy into chemical energy has
had only limited success. A few investigations have been made
into their application in the photocatalytic degradation of
organic compounds.9,31
Garcia and coworkers revealed the
semiconductor behaviour of MOF-5 and showed the potential
performance of this MOF in the photocatalyzed degradation of
phenol in aqueous solutions.32
The development of special
photocatalytic systems for selective organic transformations is
a further step in this field that has been more rarely investi-
gated. Very recently, this barrier has been broken using MOFs
containing other co-catalysis species, such as metal complex parti-
cles. Lin’s group used functionalized dicarboxylic UiO-67 containing
phosphorescent [IrIII
(ppy)2(dcbpy)]Cl and [RuII
(bpy)2(dcbpy)]Cl2
(where ppy is 2-phenylpyridine and bpy is 2,20
-bipyridine) in
three photocatalytic organic transformations (the aza-Henry
reaction, aerobic amine coupling, and aerobic oxidation of
thioanisole).33
A visible-light responsive Ti-based MOF photo-
catalyst (Ti-MOF-NH2) with 1.5 wt% Pt deposited into it was
also used for the photocatalytic reduction of nitrobenzene
Department of Chemistry, Faculty of Sciences, Tarbiat Modares University,
P.O. Box 14115-4838, Tehran, Iran. E-mail: morsali_a@modares.ac.ir;
Fax: +98 21 8009730; Tel: +98 21 82884416
† Electronic supplementary information (ESI) available. See DOI: 10.1039/
c4nj01536c
Received (in Montpellier, France)
9th September 2014,
Accepted 5th November 2014
DOI: 10.1039/c4nj01536c
www.rsc.org/njc
NJC
PAPER
Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16.
View Article Online
View Journal | View Issue
932 | New J. Chem., 2015, 39, 931--937 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015
under visible-light irradiation. The co-catalytic role of the Pt
particles was revealed by ESR measurement and, with optimum
amounts of Pt deposited in Ti-MOF-NH2, the reduction reaction
was carried out successfully within 70 h under visible light
irradiation.34
To our knowledge, there are no reports of using visible-light
responsive MOF materials without any additives in selective
organic transformations. Since morphology-controlled synthesis
of porous materials is very important for their adsorption,
separation, and catalytic selectivity,35,36
in this study the changes
in the structural properties of the Ti-based MOF photocatalyst,
MIL-NH2, which is an amine-functionalized version of MIL-125,
prepared by the conventional solvothermal method have been
compared with those of the same MOF after surfactant-assisted
solvothermal synthesis. Fortunately we could expose the visible
light photocatalytic activity of the resultant Ti-based MOF photo-
catalyst (Ti-MOF-NH2) in the regeneration of carbonyl com-
pounds from oximes (deoximation reaction) in very mild
photoreaction conditions under blue LED light irradiation.
It will be revealed that the Ti-containing MOF structure synthe-
sized in the presence of a nonionic amphiphilic surfactant,
Pluronic P123, has a more uniform cylindrical morphology with
50–250 nm particle size and more than double the surface area
of that synthesised by the conventional solvothermal method,
and that it can consequently catalyze the regeneration of carbonyl
compounds from various oximes in an energy-efficient photo-
catalytic system. Furthermore, we will successfully show that this
nanostructured catalyst is significantly stable under the reaction
conditions and can be easily recycled after the reaction, retaining
its photocatalytic reactivity.
Experimental section
Materials
Poly(ethylene oxide)–poly(propylene oxide)–poly(ethylene oxide) tri-
block copolymer template (EO20PO70EO20, Pluronic P123 from
Aldrich), 2-amino-benzenedicarboxylic acid (H2BDC-NH2; C8H7NO4
from Aldrich), 1,4-benzenedicarboxylic acid (H2BDC; C8H6O4, from
Merck), tetrapropyl orthotitanate (TPOT; Ti-(OC3H7)4, from
Aldrich), N,N-dimethylformamide (DMF, from DAE JUNG
Reagents Chemicals), and methanol (CH3OH) were used with-
out additional purification. CH3CN (from Merck) was distilled
and dried by refluxing in the presence of sodium hydride under
an argon atmosphere and kept on 4 Å molecular sieves.
All oximes were synthesized according to known procedures37
from the corresponding carbonyl compounds (from Merck), and
were characterized by nuclear magnetic resonance spectroscopy
(1
H-NMR and 13
C-NMR spectra; see the ESI†).
Instrumentation
Powder X-ray diffraction (PXRD) data were acquired on a Philips
X-Pert diffractometer using Cu Ka radiation. SEM images were
captured on a Philips XL-300 instrument. The nitrogen sorption
isotherms were collected on a Belsorp (BELMAX, Japan). The GC
analyses were performed on an Echrom A90 using a flame
ionization detector (FID). NMR spectra were recorded on a
Bruker 250 MHz spectrometer.
Synthesis of MOF structures
Synthesis of MIL-125 and MIL-NH2. MIL-125, or Ti8O8(OH)4-
[O2C–C6H4–CO2]6, and MIL-NH2 were prepared according to the
solvothermal methods reported by Serre et al.38
and Matsuoka
et al. respectively.39
Typically, a mixture of terephthalic acid, or
H2BDC, (3 mmol, 498 mg), 2 mmol of TPOT (0.6 ml), DMF
(9 ml), and methanol (1.0 ml) was transferred to a Teflon-lined
stainless-steel autoclave and heated for 48 h at 423 K under
autogenous pressure. Back at room temperature, the white
solid was recovered by filtration, washed twice with DMF and
dried under air at 80 1C. The free solvent was removed by
calcination at 200 1C overnight for 12 hours. MIL-NH2 was
synthesized in a similar way with H2BDC-NH2 instead of H2BDC.
Synthesis of SMIL-NH2. This structure was synthesized
according to a known procedure39
with a little modification.
In a typical experiment, 2.0 g of Pluronic P123 (Aldrich, average
Mw D 5800), was dissolved in 40 ml of DMF at room temperature.
The mixture was stirred magnetically, dissolving the surfactant
completely and was subsequently sonicated for 15 minutes. Then,
2 mmol (0.6 ml) of TPOT and 3 mmol (0.54 g) of H2BDC-NH2 were
added rapidly and the mixture was stirred at 35 1C for 5 h under
an Ar atmosphere. Then, for the aging process, the reaction
mixture was kept at 140 1C without stirring for 48 hours. The
mixture was cooled to room temperature and the resultant
yellow solid was filtered off, washed thoroughly with DMF and
dried in air at 80 1C. Both the free solvent and the surfactant
molecules were removed by calcination at 200 1C overnight
for 12 hours.
Photocatalytic reaction using the photocatalytic system
A mixture of oxime (20 mg) and catalyst (30 mg) in dry CH3CN
(10 ml) was prepared in a two-necked flask. One neck was used
for the oxygen inlet and outlet and the other was equipped with
a condenser with a bladder at the end. The flask was filled with
pure oxygen, and sonicated for 5 minutes. Subsequently, it was
irradiated under blue LED lamps (4 Â 3 W). Then the catalyst
was filtered off through a 0.4 mm filter and the reaction mixture
was analyzed using gas chromatography.
After the reactions of different oximes, the catalyst was
collected by centrifugation, stirred in 1 : 1 CH3CN : CH3OH for
24 hours and then dried in an oven at 75 1C and denoted as
SMIL-NH2R.
Results and discussion
After the synthesis of MIL-125 (or Ti8O8(OH)4[O2C–C6H4–CO2]6)
and MIL-NH2 according to the known procedures,38,39
the
surfactant-assisted synthesized MIL-NH2, denoted as SMIL-NH2,
was prepared using a solvothermal method in the presence of
a nonionic structure-directing agent, Pluronic P123. The powder
XRD pattern recorded for the synthesized MIL-125 control
sample matches with the literature and the pattern simulated
Paper NJC
Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16.
View Article Online
This journal is ©The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 New J. Chem., 2015, 39, 931--937 | 933
based on the X-ray data of the MIL-125 single crystal (Fig. 1a and d).
The major diffraction pattern is assigned to the rhombohedral
structure. In addition, a comparison of the simulated MIL-125
single crystal and experimental powder X-ray diffraction patterns of
both the synthesized MIL-NH2 and the surfactant-assisted synthe-
sized structure, which is named SMIL-NH2, confirmed the parent
structure estimated from the basic crystallographic data (Fig. 1a–c).
As Fig. 2 shows, SEM images of MIL-NH2 and SMIL-NH2 indicate
that in the presence of the nonionic block copolymer, Pluronic
P123, the angular plate structures of MIL-NH2, with nonuniform
crystal sizes between 400 and 1000 nm, clearly convert into the
soft plate structures of SMIL-NH2, with smaller crystals of about
50–250 nm. Moreover, more uniform cylindrical structures
have been constructed from the surfactant-assisted synthesis
system. Recently, the effect of the surfactant-assisted medium
on the morphology and particle size of [Cu3(btc)2], known as
HKUST-1 has been studied.40
As has been well demonstrated,
in the presence of the cationic surfactant CTAB in aqueous
solution, the triethylammonium salt of the BTC ligand, btc3À
, is
surrounded by CTA+
species. Therefore, in the presence of
higher amounts of CTAB as an additive, nucleation is retarded
significantly, which causes a larger particle size for the final
crystals. In contrast, it has been shown that via common routes
of MOF synthesis in organic solvents, increased concentration
of additives leads to smaller crystal size, due to suppression of
crystal growth.40,41
It seems that in the presence of the nonionic
surfactant P123, in DMF as a solvent, controlled particle
formation of carboxylic acid-based MOFs in solution generally
occurs and this affects the ultimate crystal size.41
However, the
particle size decrease and morphology alteration observed in
the presence of P123 is a result of competition between the
surfactant and the ligand for coordination to Ti6+
sites in this
medium, instead of the surfactant being assembled to make
micelles and play its usual templating role in the solution. Very
recently, we showed that in the presence of the same concen-
tration of P123 in DMF solvent, but with copper salt and BDC
ligand, the precursors of the HKUST-1 structure, the template
effect of the nonionic surfactant directed the final structure.42
The typical octahedral morphology of the crystals remained
intact while some new ordered domains were constructed
within the original framework. Here, the existence of the
oxophilic Ti6+
metal centers in the relatively harsh conditions
of the solvothermal reactor makes a completely different path
for dictating the final MOF structure. A comparison of N2
adsorption–desorption analysis of MIL-NH2 and SMIL-NH2
clearly verified this statement (Fig. 3). As shown in Fig. 3a,
the hysteresis loop in the N2 adsorption–desorption isotherm of
MIL-NH2 indicates a type IV mesoporous material, and this has
completely disappeared in the isotherm of SMIL-NH2, which
shows a type I isotherm, characteristic of microporous materials.
This observation that is quite the opposite of what was seen in
the case of HKUST-1, evidently proving the discrepancy of the
surfactant role in the two cases.
To investigate the photocatalytic activity of MIL-NH2 and its
surfactant-assisted synthesized analogue, SMIL-NH2, regenera-
tion of carbonyl compounds from oximes, or the deoximation
reaction, was chosen as a probe. This is a useful organic
transformation because of its use in the purification and
characterization of carbonyl compounds and also its importance
in organic synthesis.43–47
This reaction has proceeded in crystal-
line TiO2-based photocatalytic systems under UV and also solar
light.48,49
Recently, a new binary photocatalyst system including
a periodic mesoporous organosilicate (PMO) decorated with
amorphous TiO2 was developed for the deoximation reaction
of various types of aldoximes and ketoximes under sunlight
irradiation.50
Nowadays, the quest for the use of high photon
Fig. 1 Powder XRD patterns: (a) simulated from X-ray data of MIL-125
single crystal, (b) MIL-NH2, (c) SMIL-NH2, (d) MIL-125, (e) SMIL-NH2R.
Fig. 2 SEM images of solvothermally synthesized MIL-NH2, (a) and (b),
and surfactant-assisted synthesized SMIL-NH2, (c) and (d).
NJC Paper
Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16.
View Article Online
934 | New J. Chem., 2015, 39, 931--937 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015
efficiency, low voltage of electricity, and high power stability LED
lamps51
as a light source has attracted wide attention, owing to
their simplified practical aspects as well from the standpoint of
environmental and economic concerns for cleaner and cheaper
organic transformations.52–55
Fortunately, both MIL-NH2 and
SMIL-NH2 could successfully catalyze the deoximation reaction
of acetophenone oxime in the presence of blue LED lamps (4 Â 3 W)
under an oxygen atmosphere. At the beginning of the reaction,
when the oxime concentration is high, in the presence of either
MIL-NH2 or SMIL-NH2 the rate of acetophenone production is
the same (Fig. 4). Interestingly, after 4 h, the concentration
curves for the product formation with respect to time begin to
separate and the rate of product formation with SMIL-NH2 is
about twice as fast as that with MIL-NH2. Lattice parameters of
the two structures have been provided in Table 1. As these
clearly show, and as is confirmed by the N2 adsorption–
desorption analysis, MIL-NH2 and SMIL-NH2 are entirely dis-
similar in pore size distribution and eventually in their surface
areas. While MIL-NH2 has a 3.3 nm average mesopore size and
a 566 m2
gÀ1
surface area, SMIL-NH2 has a smaller pore size of
2.4 nm and a larger surface area of 1132 m2
gÀ1
, which is more
than two times that of the former. Although the fact that both
reactions have the same induction period needs more precise
investigation, the improved lattice parameters may be one of
the reasons for the higher catalytic activity of SMIL-NH2.
In confirmation of the SEM results, these data evidently are
in contrast with the previous results reported for HKUST-1.42
Here the nonionic surfactant, P123, acts as a capping agent
instead of being assembled, making micelles and acting as
scaffold for construction of the MOF structure. The discrepancy
may be due to the different affinities of the two metal centers,
copper and titanium ions, to the hydrophilic head groups of the
surfactant in the organic solution.
No change was observed in the absence of irradiation,
catalyst or oxygen for runs carried out under the same experi-
mental conditions (Table 2, entry 2). Since no reaction took
place in the presence of MIL-125 as a catalyst (Table 2, entry 3)
even after 48 h irradiation, the role of the amine-functionalized
versions of the MIL can be described as an antenna effect.
UV-vis spectra of MIL-NH2 and SMIL-NH2 are shown in Fig. 5.
For both structures, the major absorption edge is similar and in
the visible light region extends to around 500 nm, relative to the
350 nm absorption edge reported for MIL-125. The observed
distinction is known to be induced by O–Ti ligand-to-metal
charge transfer in TiO5(OH) inorganic clusters, apparently
influenced by H2BDC being replaced with H2BDC-NH2.6,8
Therefore, the antenna role of the amine-functionalized MILs
in the photocatalytic deoximation reaction of oximes in an
oxygen atmosphere can be described as in Scheme 1. In this
known manner, an electron and a hole are generated from light
excitation, and transfer to the oxygen and oxime molecules,
respectively.7,56–58
Moreover, a Ti-leaching test has been per-
formed after a 10 h reaction of acetophenone oxime under blue
Fig. 3 (a) N2 adsorption–desorption analysis and (b) BJH analysis of
MIL-NH2, SMIL-NH2 and SMIL-NH2.
Fig. 4 Yield-versus-time profiles of the deoximation reaction in the
presence of (a) MIL-NH2, (b) SMIL-NH2 and (c) Degussa P25; y-axis: yield
of acetophenone in %, x-axis: time in hours.
Table 1 Lattice parameters of MIL-NH2 structures prepared in the
absence or presence of P123
Sample SBET
a
(m2
gÀ1
) Dmicro
b
(nm) Dmeso
c
(nm) Vmeso
c
(cm3
gÀ1
)
MIL-NH2 566 0.7 3.3 0.179
SMIL-NH2 1132 0.6 2.4 0.151
SMIL-NH2Rd
1066 0.6 2.4 0.151
a
The Brunauer–Emmett–Teller (BET) surface area. b
The micropore
size distributions were calculated using an MP plot. c
The mesopore
volume and the mesopore size distributions were calculated using the
Barrett–Joyner–Halenda (BJH) method. d
The recycled catalyst.
Paper NJC
Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16.
View Article Online
This journal is ©The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 New J. Chem., 2015, 39, 931--937 | 935
LED light irradiation. The ICP analysis of the reaction solution
showed that no Ti species has been leached, which validates the
stability of the MOF structure as well as the heterogeneous
character of our catalytic system.
The remarkable photocatalytic activity of SMIL-NH2 in such
mild reaction conditions encouraged us to examine its photo-
catalytic ability for other structurally diverse oximes. Since photo-
catalytic deoximation has been known to proceed via a cationic
intermediate formed via the reaction of the oxime substrate with
photogenerated holes,49
activated ketoximes with electron-
donating substituents such as methyl and methoxy were equally
well converted into the corresponding ketones in relatively short
times under mild reaction conditions (Table 2, entries 4 and 5).
Very recently, a deoximation reaction of 4-methoxyacetophenone
oxime to the corresponding ketone in the presence of C–N
codoped sonofabricated crystalline titania (anatase) as well as a
sample of bicrystalline framework titania, (anatase/brookite,
78/22) was examined under blue LED irradiation.59
In similar
reaction conditions to our photocatalytic system, with the
former catalyst the reaction proceeded quantitatively within 2 h,
Table 2 Regeneration of carbonyl compounds from oximes in blue LED
light irradiation using SMIL-NH2
Entry Substrate Product
Time
(h)
Yielda
(%)
1 10 100
2
48b
—
12c
100
3d
48 —
4 6 100
5
4.5 100
4.5e
86
6 48 57
7 48 74
8 24 100
9 24 100
10 48 33
11 48 38
a
GC yield, reaction conditions: oxime (20 mg), catalyst (30 mg), 10 ml dry
CH3CN, O2 atmosphere (1 atm), blue LED light (4 Â 3 W lamps). b
In the
dark, in the absence of O2 and/or catalyst the same results were obtained.
c
In the presence of the recycled catalyst, SMIL-NH2R. d
In the presence
of MIL-125 as a catalyst. e
In the presence of P25 as a catalyst.
Fig. 5 UV-vis spectra of (a) MIL-NH2 and (b) SMIL-NH2.
Scheme 1 The antenna role of MIL-NH2 in transferring the generated
electron and hole to the oxygen and oxime molecules during light
excitation.
NJC Paper
Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16.
View Article Online
936 | New J. Chem., 2015, 39, 931--937 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015
whereas just 31 percent of the corresponding carbonyl compound
was obtained with the latter catalyst. With the most reactive
commercial crystalline TiO2, Degussa P25 (anatase/rutile, 75/25),
86 percent of 4-methoxyacetophenone was obtained in the same
reaction conditions mentioned in Table 2, in comparison with
a quantitative amount of the product formed in the presence of
SMIL-NH2 within 4.5 h irradiation (Table 2, entry 5). It is worth
mentioning that there is no report of the deoximation reaction
with MOFs in such mild reaction conditions. Moreover, the
crystalline structure of these materials is completely different
and there are no diffraction peaks of any TiO2 crystalline phase
(anatase, rutile or brookite) in their structure. Meanwhile,
after 48 h irradiation, deactivated oximes containing electron
withdrawing groups such as 4-chloroacetophenone oxime,
3-bromoacetophenone oxime and the more passive substrate,
methyl 2-pyridyl ketone oxime gave weak to moderate yields of
the corresponding carbonyl products (Table 2, entries 6, 7 and 11).
Furthermore, the reaction of sterically hindered oxime substrates
proceeded over a long time and was not completed in the case of
deactivated substrates such as (4-chlorophenyl)(phenyl)methanone
oxime (Table 2, entries 9 and 10). This observation shows the
preference of the substrate to pass into the channels of the
photoactive MOF framework rather than to react on the surface.
In other words, the reaction proceeds within the channels of
the MOF framework.
For practical purposes, recycling of the catalyst is an important
problem to solve. Considering the potential application of
SMIL-NH2, we collected the catalyst used in the deoximation
reactions of diverse substrates, thoroughly washed it with 1 :1 of
CH3CN and MeOH, and subsequently dried it in an oven at 80 1C
for 12 h. A powder XRD pattern of this recycled catalyst, denoted
as SMIL-NH2R, is shown in Fig. 1e, and obviously indicates that
the crystallinity of the MOF structure has been reduced slightly,
while the original XRD pattern of the structure remains intact and
matches with the X-ray data of the MIL-125 single crystal. The
deoximation reaction of acetophenone oxime in the presence of
SMIL-NH2R in the same optimum conditions mentioned in
Table 2 gave a quantitative yield of acetophenone after 12 h
irradiation (Table 2, entry 2). SEM images of SMIL-NH2R before
and after washing the catalyst, shown in Fig. S1 (ESI†), not only
verify the importance and effectiveness of the washing stage to
disconnect the agglomerated particles of the MOF structure
formed during the reaction, but also confirm the basic morpho-
logy remains nearly intact after the deoximation reaction. In
addition, some residual nitrogen content (about 7%) seen in
the energy dispersive X-ray analysis (EDS) of the unwashed
catalyst, which has been eliminated in the EDS analysis of
the washed catalyst, clearly validates the SEM results (Fig. S1
and S2, ESI†). Accordingly, partial agglomeration of the cylindrical
structures of SMIL-NH2 and trapping of a few organic molecules
in the MOF structure (existence of carbon content in the EDS of
SMIL-NH2R, Fig. S3, ESI†), can occupy active Ti centers and
reduce the photocatalytic activity of the catalyst. It is also
worth mentioning that despite the structural changes observed,
SMIL-NH2R is more reactive than MIL-NH2. A comparison of
the porosity parameters shown in Table 1 shows that just a little
decrease in the surface area of SMIL-NH2R relative to SMIL-NH2
does not affect its activity, which is still greater than that of
MIL-NH2.
Conclusion
We succeeded in the preparation of a new morphology of
MIL-NH2 in the presence of Pluronic P123, as a structure-
directing agent and in solvothermal conditions. Adsorption–
desorption isotherm, BET, and SEM analyses clearly established
the effectiveness of the procedure for the synthesis of uniform
cylindrical SMIL-NH2 nanoparticles possessing double the surface
area of MIL-NH2, with 50–250 nm particle sizes. Furthermore,
regeneration of carbonyl compounds from various oximes in an
energy-efficient photocatalytic system was successfully catalyzed
by this stable nanostructured photocatalyst in an O2 atmosphere
and under blue LED light irradiation. In addition, it was shown
that this nanostructured MOF catalyst is stable in the reaction
conditions and can be easily recycled after the reaction, retaining
its photocatalytic reactivity.
Acknowledgements
This work was supported by Tarbiat Modares University.
Notes and references
1 A. Mills and S. Le Hunte, J. Photochem. Photobiol., A, 1997,
108, 1–35.
2 N. J. Peill, L. Bourne and M. R. Hoffmann, J. Photochem.
Photobiol., A, 1997, 108, 221–228.
3 M. R. Hoffmann, S. T. Martin, W. Choi and D. W. Bahnemann,
Chem. Rev., 1995, 95, 69–96.
4 R. J. Kuppler, D. J. Timmons, Q.-R. Fang, J.-R. Li, T. A.
Makal, M. D. Young, D. Yuan, D. Zhao, W. Zhuang and H.-C.
Zhou, Coord. Chem. Rev., 2009, 253, 3042–3066.
5 I. Fechete, Y. Wang and J. C. Ve´drine, Catal. Today, 2012,
189, 2–27.
6 S.-N. Kim, J. Kim, H.-Y. Kim, H.-Y. Cho and W.-S. Ahn, Catal.
Today, 2013, 204, 85–93.
7 J. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. T. Nguyen and
J. T. Hupp, Chem. Soc. Rev., 2009, 38, 1450–1459.
8 N. D. McNamara, G. T. Neumann, E. T. Masko, J. A. Urban
and J. C. Hicks, J. Catal., 2013, 305, 217–226.
9 J.-L. Wang, C. Wang and W. Lin, ACS Catal., 2012, 2,
2630–2640.
10 C. A. Kent, D. Liu, L. Ma, J. M. Papanikolas, T. J. Meyer and
W. Lin, J. Am. Chem. Soc., 2011, 133, 12940–12943.
11 C. A. Kent, D. Liu, T. J. Meyer and W. Lin, J. Am. Chem. Soc.,
2012, 134, 3991–3994.
12 K. C. Stylianou, R. Heck, S. Y. Chong, J. Bacsa, J. T. A. Jones,
Y. Z. Khimyak, D. Bradshaw and M. J. Rosseinsky, J. Am.
Chem. Soc., 2010, 132, 4119–4130.
13 C. Gomes Silva, I. Luz, F. X. Llabre´s i Xamena, A. Corma and
H. Garcı´a, Chem. – Eur. J., 2010, 16, 11133–11138.
Paper NJC
Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16.
View Article Online
This journal is ©The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 New J. Chem., 2015, 39, 931--937 | 937
14 A. Fateeva, P. A. Chater, C. P. Ireland, A. A. Tahir, Y. Z.
Khimyak, P. V. Wiper, J. R. Darwent and M. J. Rosseinsky,
Angew. Chem., 2012, 124, 7558–7562.
15 C. Wang, K. E. deKrafft and W. Lin, J. Am. Chem. Soc., 2012,
134, 7211–7214.
16 C. Wang, J.-L. Wang and W. Lin, J. Am. Chem. Soc., 2012,
134, 19895–19908.
17 Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu and Z. Li,
Angew. Chem., Int. Ed., 2012, 51, 3364–3367.
18 S. Horike and S. Kitagawa, Metal-Organic Frameworks,
Wiley-VCH Verlag GmbH & Co. KGaA, 2011, pp. 1–21.
19 Y.-X. Tan, Y.-P. He and J. Zhang, Chem. Mater., 2012, 24,
4711–4716.
20 M. C. Das, S. Xiang, Z. Zhang and B. Chen, Angew. Chem.,
Int. Ed., 2011, 50, 10510–10520.
21 M. Fuentes-Cabrera, D. M. Nicholson, B. G. Sumpter and
M. Widom, J. Chem. Phys., 2005, 123, 124713.
22 J. Gascon, M. D. Herna´ndez-Alonso, A. R. Almeida, G. P. M.
van Klink, F. Kapteijn and G. Mul, ChemSusChem, 2008, 1,
981–983.
23 A. Corma, H. Garcı´a and F. X. Llabre´s i Xamena, Chem. Rev.,
2010, 110, 4606–4655.
24 S. Horike, M. Dinca˘, K. Tamaki and J. R. Long, J. Am. Chem.
Soc., 2008, 130, 5854–5855.
25 F. Song, C. Wang, J. M. Falkowski, L. Ma and W. Lin, J. Am.
Chem. Soc., 2010, 132, 15390–15398.
26 M. Yoon, R. Srirambalaji and K. Kim, Chem. Rev., 2012, 112,
1196–1231.
27 K. Leus, G. Vanhaelewyn, T. Bogaerts, Y.-Y. Liu, D. Esquivel,
F. Callens, G. B. Marin, V. Van Speybroeck, H. Vrielinck and
P. Van Der Voort, Catal. Today, 2013, 208, 97–105.
28 G. Calleja, R. Sanz, G. Orcajo, D. Briones, P. Leo and
F. Martı´nez, Catal. Today, 2014, 227, 130–137.
29 L. Chen, Z. Gao and Y. Li, Catal. Today, DOI: 10.1016/
j.cattod.2014.03.074.
30 E. E. Macias, P. Ratnasamy and M. A. Carreon, Catal. Today,
2012, 198, 215–218.
31 L. Wen, L. Zhou, B. Zhang, X. Meng, H. Qu and D. Li,
J. Mater. Chem., 2012, 22, 22603–22609.
32 M. Alvaro, E. Carbonell, B. Ferrer, F. X. Llabre´s i Xamena
and H. Garcia, Chem. – Eur. J., 2007, 13, 5106–5112.
33 C. Wang, Z. Xie, K. E. deKrafft and W. Lin, J. Am. Chem. Soc.,
2011, 133, 13445–13454.
34 T. Toyao, M. Saito, Y. Horiuchi, K. Mochizuki, M. Iwata,
H. Higashimura and M. Matsuoka, Catal. Sci. Technol.,
2013, 3, 2092–2097.
35 Y. Wan, Y. Shi and D. Zhao, Chem. Mater., 2008, 20, 932–945.
36 S. G. Rudisill, Z. Wang and A. Stein, Langmuir, 2012, 28,
7310–7324.
37 B. S. Furniss, Vogel’s textbook of practical organic chemistry,
Pearson Education India, 1989.
38 M. Dan-Hardi, C. Serre, T. Frot, L. Rozes, G. Maurin,
C. Sanchez and G. Fe´rey, J. Am. Chem. Soc., 2009, 131,
10857–10859.
39 Y. Horiuchi, T. Toyao, M. Saito, K. Mochizuki, M. Iwata,
H. Higashimura, M. Anpo and M. Matsuoka, J. Phys. Chem.
C, 2012, 116, 20848–20853.
40 T. Uemura and S. Kitagawa, Chem. Lett., 2005, 34, 132–137.
41 S. Hermes, T. Witte, T. Hikov, D. Zacher, S. Bahnmu¨ller,
G. Langstein, K. Huber and R. A. Fischer, J. Am. Chem. Soc.,
2007, 129, 5324–5325.
42 S. Abedi and A. Morsali, ACS Catal., 2014, 4, 1398–1403.
43 D. H. R. Barton, J. M. Beaton, L. E. Geller and M. M. Pechet,
J. Am. Chem. Soc., 1961, 83, 4076–4083.
44 D. H. R. Barton and J. M. Beaton, J. Am. Chem. Soc., 1961, 83,
4083–4089.
45 R. S. Varma, Green Chem., 1999, 1, 43–55.
46 W. Chrisman, M. J. Blankinship, B. Taylor and C. E. Harris,
Tetrahedron Lett., 2001, 42, 4775–4777.
47 A. Khazaei, R. Ghorbani Vaghei and M. Tajbakhsh, Tetra-
hedron Lett., 2001, 42, 5099–5100.
48 C. Saravanselvi, N. Somasundaram, S. Vijaikumar and
C. Srinivasan, Photochem. Photobiol. Sci., 2002, 1, 607–608.
49 K. Muthu, K. Selvam, B. Krishnakumar and M. Swaminathan,
Appl. Catal., A, 2009, 358, 259–263.
50 S. Abedi, B. Karimi, F. Kazemi, M. Bostina and H. Vali,
Org. Biomol. Chem., 2013, 11, 416–419.
51 N. G. Yeh, C.-H. Wu and T. C. Cheng, Renewable Sustainable
Energy Rev., 2010, 14, 2161–2166.
52 S. Fu¨ldner, R. Mild, H. I. Siegmund, J. A. Schroeder,
M. Gruber and B. Ko¨nig, Green Chem., 2010, 12, 400–406.
53 S. Fu¨ldner, T. Mitkina, T. Trottmann, A. Frimberger,
M. Gruber and B. Ko¨nig, Photochem. Photobiol. Sci., 2011,
10, 623–625.
54 S. Fu¨ldner, P. Pohla, H. Bartling, S. Dankesreiter, R. Stadler,
M. Gruber, A. Pfitzner and B. Ko¨nig, Green Chem., 2011, 13,
640–643.
55 Z. Zand, F. Kazemi and S. Hosseini, Tetrahedron Lett., 2014,
55, 338–341.
56 D. Farrusseng, S. Aguado and C. Pinel, Angew. Chem.,
Int. Ed., 2009, 48, 7502–7513.
57 U. Ravon, M. E. Domine, C. Gaudillere, A. Desmartin-
Chomel and D. Farrusseng, New J. Chem., 2008, 32, 937–940.
58 L. Alaerts, E. Se´guin, H. Poelman, F. Thibault-Starzyk,
P. A. Jacobs and D. E. De Vos, Chem. – Eur. J., 2006, 12,
7353–7363.
59 M. Abdollahi Kakroudi, F. Kazemi and B. Kaboudin, J. Mol.
Catal. A: Chem., 2014, 392, 112–119.
NJC Paper
Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16.
View Article Online

More Related Content

What's hot

Preparation, characterization and application of sonochemically doped fe3+ in...
Preparation, characterization and application of sonochemically doped fe3+ in...Preparation, characterization and application of sonochemically doped fe3+ in...
Preparation, characterization and application of sonochemically doped fe3+ in...eSAT Journals
 
Characterization of Clay/Chitosan Nanocomposites and their Use for Adsorption...
Characterization of Clay/Chitosan Nanocomposites and their Use for Adsorption...Characterization of Clay/Chitosan Nanocomposites and their Use for Adsorption...
Characterization of Clay/Chitosan Nanocomposites and their Use for Adsorption...Editor IJCATR
 
Chelating ion exchange and antimicrobial studies
Chelating ion exchange and antimicrobial studiesChelating ion exchange and antimicrobial studies
Chelating ion exchange and antimicrobial studiesIJECSJournal
 
Applications of metal nanoparticles in photocatalysis
Applications of metal nanoparticles in photocatalysisApplications of metal nanoparticles in photocatalysis
Applications of metal nanoparticles in photocatalysisChamudithaBenaragama
 
Degradation of Paracetamol by Electro-Fenton and Photoelectro-Fenton Processe...
Degradation of Paracetamol by Electro-Fenton and Photoelectro-Fenton Processe...Degradation of Paracetamol by Electro-Fenton and Photoelectro-Fenton Processe...
Degradation of Paracetamol by Electro-Fenton and Photoelectro-Fenton Processe...Oswar Mungkasa
 
Solid base catalyzed depolymerization (liquifaction/valorization) of lignin i...
Solid base catalyzed depolymerization (liquifaction/valorization) of lignin i...Solid base catalyzed depolymerization (liquifaction/valorization) of lignin i...
Solid base catalyzed depolymerization (liquifaction/valorization) of lignin i...Richa Chaudhary
 
Mesoporous Catalysis
Mesoporous CatalysisMesoporous Catalysis
Mesoporous CatalysisAndrew Finkle
 
Lignin-depolymerization-aromatic monomers-solid acid-heterogeneous catalyst-A...
Lignin-depolymerization-aromatic monomers-solid acid-heterogeneous catalyst-A...Lignin-depolymerization-aromatic monomers-solid acid-heterogeneous catalyst-A...
Lignin-depolymerization-aromatic monomers-solid acid-heterogeneous catalyst-A...Deepa A K
 
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...IOSRJAC
 
Synthesis and Characterization of Schiff Base from Aromatic Amine and Aromati...
Synthesis and Characterization of Schiff Base from Aromatic Amine and Aromati...Synthesis and Characterization of Schiff Base from Aromatic Amine and Aromati...
Synthesis and Characterization of Schiff Base from Aromatic Amine and Aromati...ijtsrd
 
Development of Mesoporous Materials and Noble Metal Based Hydrodesulfurizatio...
Development of Mesoporous Materials and Noble Metal Based Hydrodesulfurizatio...Development of Mesoporous Materials and Noble Metal Based Hydrodesulfurizatio...
Development of Mesoporous Materials and Noble Metal Based Hydrodesulfurizatio...Lebong Andalaluna
 
Schiff base and complexe
Schiff base and complexeSchiff base and complexe
Schiff base and complexewael alharbi
 
Facultative methylotrophic bacterium вrevibacterium
Facultative methylotrophic bacterium вrevibacteriumFacultative methylotrophic bacterium вrevibacterium
Facultative methylotrophic bacterium вrevibacteriumAlexander Decker
 

What's hot (19)

Preparation, characterization and application of sonochemically doped fe3+ in...
Preparation, characterization and application of sonochemically doped fe3+ in...Preparation, characterization and application of sonochemically doped fe3+ in...
Preparation, characterization and application of sonochemically doped fe3+ in...
 
Characterization of Clay/Chitosan Nanocomposites and their Use for Adsorption...
Characterization of Clay/Chitosan Nanocomposites and their Use for Adsorption...Characterization of Clay/Chitosan Nanocomposites and their Use for Adsorption...
Characterization of Clay/Chitosan Nanocomposites and their Use for Adsorption...
 
Chelating ion exchange and antimicrobial studies
Chelating ion exchange and antimicrobial studiesChelating ion exchange and antimicrobial studies
Chelating ion exchange and antimicrobial studies
 
Applications of metal nanoparticles in photocatalysis
Applications of metal nanoparticles in photocatalysisApplications of metal nanoparticles in photocatalysis
Applications of metal nanoparticles in photocatalysis
 
Degradation of Paracetamol by Electro-Fenton and Photoelectro-Fenton Processe...
Degradation of Paracetamol by Electro-Fenton and Photoelectro-Fenton Processe...Degradation of Paracetamol by Electro-Fenton and Photoelectro-Fenton Processe...
Degradation of Paracetamol by Electro-Fenton and Photoelectro-Fenton Processe...
 
Nanocatalyst
NanocatalystNanocatalyst
Nanocatalyst
 
Solid base catalyzed depolymerization (liquifaction/valorization) of lignin i...
Solid base catalyzed depolymerization (liquifaction/valorization) of lignin i...Solid base catalyzed depolymerization (liquifaction/valorization) of lignin i...
Solid base catalyzed depolymerization (liquifaction/valorization) of lignin i...
 
Mesoporous Catalysis
Mesoporous CatalysisMesoporous Catalysis
Mesoporous Catalysis
 
Lignin-depolymerization-aromatic monomers-solid acid-heterogeneous catalyst-A...
Lignin-depolymerization-aromatic monomers-solid acid-heterogeneous catalyst-A...Lignin-depolymerization-aromatic monomers-solid acid-heterogeneous catalyst-A...
Lignin-depolymerization-aromatic monomers-solid acid-heterogeneous catalyst-A...
 
Bech2007
Bech2007Bech2007
Bech2007
 
SYNTHESIS, CHARACTERIZATION AND ANTIMICROBIAL ACTIVITIES OF{FE(II),CO(II),NI(...
SYNTHESIS, CHARACTERIZATION AND ANTIMICROBIAL ACTIVITIES OF{FE(II),CO(II),NI(...SYNTHESIS, CHARACTERIZATION AND ANTIMICROBIAL ACTIVITIES OF{FE(II),CO(II),NI(...
SYNTHESIS, CHARACTERIZATION AND ANTIMICROBIAL ACTIVITIES OF{FE(II),CO(II),NI(...
 
Nanocatalysis
NanocatalysisNanocatalysis
Nanocatalysis
 
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
Synthesis, Characterization, Antibacterial and DNA Binding Studies of Mn (II)...
 
He2513001307
He2513001307He2513001307
He2513001307
 
Synthesis and Characterization of Schiff Base from Aromatic Amine and Aromati...
Synthesis and Characterization of Schiff Base from Aromatic Amine and Aromati...Synthesis and Characterization of Schiff Base from Aromatic Amine and Aromati...
Synthesis and Characterization of Schiff Base from Aromatic Amine and Aromati...
 
Development of Mesoporous Materials and Noble Metal Based Hydrodesulfurizatio...
Development of Mesoporous Materials and Noble Metal Based Hydrodesulfurizatio...Development of Mesoporous Materials and Noble Metal Based Hydrodesulfurizatio...
Development of Mesoporous Materials and Noble Metal Based Hydrodesulfurizatio...
 
CHEMICAL ENG JOURNAL
CHEMICAL ENG JOURNALCHEMICAL ENG JOURNAL
CHEMICAL ENG JOURNAL
 
Schiff base and complexe
Schiff base and complexeSchiff base and complexe
Schiff base and complexe
 
Facultative methylotrophic bacterium вrevibacterium
Facultative methylotrophic bacterium вrevibacteriumFacultative methylotrophic bacterium вrevibacterium
Facultative methylotrophic bacterium вrevibacterium
 

Similar to New J. Chem., 2015,39, 931-937

Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Pawan Kumar
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Pawan Kumar
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Pawan Kumar
 
Spatial charge separation on the (110)/(102) facets of cocatalyst-free ZnIn2S...
Spatial charge separation on the (110)/(102) facets of cocatalyst-free ZnIn2S...Spatial charge separation on the (110)/(102) facets of cocatalyst-free ZnIn2S...
Spatial charge separation on the (110)/(102) facets of cocatalyst-free ZnIn2S...Pawan Kumar
 
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...Pawan Kumar
 
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...Pawan Kumar
 
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...Pawan Kumar
 
Mixed Ligand, Palladium(II) and Platinum(II) Complexes of Tertiary Diphosphin...
Mixed Ligand, Palladium(II) and Platinum(II) Complexes of Tertiary Diphosphin...Mixed Ligand, Palladium(II) and Platinum(II) Complexes of Tertiary Diphosphin...
Mixed Ligand, Palladium(II) and Platinum(II) Complexes of Tertiary Diphosphin...Karwan Omer
 
Octahedral rhenium k4[re6 s8(cn)6] and
Octahedral rhenium k4[re6 s8(cn)6] andOctahedral rhenium k4[re6 s8(cn)6] and
Octahedral rhenium k4[re6 s8(cn)6] andPawan Kumar
 
Octahedral rhenium K4[Re6S8(CN)6] and Cu(OH)2cluster modifiedTiO2for the phot...
Octahedral rhenium K4[Re6S8(CN)6] and Cu(OH)2cluster modifiedTiO2for the phot...Octahedral rhenium K4[Re6S8(CN)6] and Cu(OH)2cluster modifiedTiO2for the phot...
Octahedral rhenium K4[Re6S8(CN)6] and Cu(OH)2cluster modifiedTiO2for the phot...Pawan Kumar
 
8100 10718-1-pb adsorption study
8100 10718-1-pb adsorption study8100 10718-1-pb adsorption study
8100 10718-1-pb adsorption studyTaghreed Al-Noor
 
Photoreduction of CO2 to methanol with hexanuclear molybdenum [Mo6Br14]2 clu...
Photoreduction of CO2 to methanol with hexanuclear molybdenum [Mo6Br14]2 clu...Photoreduction of CO2 to methanol with hexanuclear molybdenum [Mo6Br14]2 clu...
Photoreduction of CO2 to methanol with hexanuclear molybdenum [Mo6Br14]2 clu...Pawan Kumar
 
Final MS Presentation.pptx
Final MS Presentation.pptxFinal MS Presentation.pptx
Final MS Presentation.pptxZahidMasood17
 
Efficient buchwald hartwig reaction catalyzed by spions-bis(NHC)-Pd(II)
Efficient buchwald hartwig reaction catalyzed by spions-bis(NHC)-Pd(II)Efficient buchwald hartwig reaction catalyzed by spions-bis(NHC)-Pd(II)
Efficient buchwald hartwig reaction catalyzed by spions-bis(NHC)-Pd(II)Iranian Chemical Society
 
Electrochemical Synthesis of MoO3 Nanoparticles Effect of Temperature Convert...
Electrochemical Synthesis of MoO3 Nanoparticles Effect of Temperature Convert...Electrochemical Synthesis of MoO3 Nanoparticles Effect of Temperature Convert...
Electrochemical Synthesis of MoO3 Nanoparticles Effect of Temperature Convert...IJERA Editor
 
Experimental and theoretical investigations of some pyrazolo-pyrimidine deriv...
Experimental and theoretical investigations of some pyrazolo-pyrimidine deriv...Experimental and theoretical investigations of some pyrazolo-pyrimidine deriv...
Experimental and theoretical investigations of some pyrazolo-pyrimidine deriv...Al Baha University
 
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...ijtsrd
 

Similar to New J. Chem., 2015,39, 931-937 (20)

Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
 
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
Metal-organic hybrid: Photoreduction of CO2 using graphitic carbon nitride su...
 
Spatial charge separation on the (110)/(102) facets of cocatalyst-free ZnIn2S...
Spatial charge separation on the (110)/(102) facets of cocatalyst-free ZnIn2S...Spatial charge separation on the (110)/(102) facets of cocatalyst-free ZnIn2S...
Spatial charge separation on the (110)/(102) facets of cocatalyst-free ZnIn2S...
 
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
Optical Control of Selectivity of High Rate CO2 Photoreduction Via Interband-...
 
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
 
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
Reduced graphene oxide–CuO nanocomposites for photocatalyticconversion of CO2...
 
Mixed Ligand, Palladium(II) and Platinum(II) Complexes of Tertiary Diphosphin...
Mixed Ligand, Palladium(II) and Platinum(II) Complexes of Tertiary Diphosphin...Mixed Ligand, Palladium(II) and Platinum(II) Complexes of Tertiary Diphosphin...
Mixed Ligand, Palladium(II) and Platinum(II) Complexes of Tertiary Diphosphin...
 
Octahedral rhenium k4[re6 s8(cn)6] and
Octahedral rhenium k4[re6 s8(cn)6] andOctahedral rhenium k4[re6 s8(cn)6] and
Octahedral rhenium k4[re6 s8(cn)6] and
 
Octahedral rhenium K4[Re6S8(CN)6] and Cu(OH)2cluster modifiedTiO2for the phot...
Octahedral rhenium K4[Re6S8(CN)6] and Cu(OH)2cluster modifiedTiO2for the phot...Octahedral rhenium K4[Re6S8(CN)6] and Cu(OH)2cluster modifiedTiO2for the phot...
Octahedral rhenium K4[Re6S8(CN)6] and Cu(OH)2cluster modifiedTiO2for the phot...
 
8100 10718-1-pb adsorption study
8100 10718-1-pb adsorption study8100 10718-1-pb adsorption study
8100 10718-1-pb adsorption study
 
Photoreduction of CO2 to methanol with hexanuclear molybdenum [Mo6Br14]2 clu...
Photoreduction of CO2 to methanol with hexanuclear molybdenum [Mo6Br14]2 clu...Photoreduction of CO2 to methanol with hexanuclear molybdenum [Mo6Br14]2 clu...
Photoreduction of CO2 to methanol with hexanuclear molybdenum [Mo6Br14]2 clu...
 
Maryam Bachelor thesis
Maryam Bachelor thesisMaryam Bachelor thesis
Maryam Bachelor thesis
 
Final MS Presentation.pptx
Final MS Presentation.pptxFinal MS Presentation.pptx
Final MS Presentation.pptx
 
Efficient buchwald hartwig reaction catalyzed by spions-bis(NHC)-Pd(II)
Efficient buchwald hartwig reaction catalyzed by spions-bis(NHC)-Pd(II)Efficient buchwald hartwig reaction catalyzed by spions-bis(NHC)-Pd(II)
Efficient buchwald hartwig reaction catalyzed by spions-bis(NHC)-Pd(II)
 
Electrochemical Synthesis of MoO3 Nanoparticles Effect of Temperature Convert...
Electrochemical Synthesis of MoO3 Nanoparticles Effect of Temperature Convert...Electrochemical Synthesis of MoO3 Nanoparticles Effect of Temperature Convert...
Electrochemical Synthesis of MoO3 Nanoparticles Effect of Temperature Convert...
 
Research Paper
Research PaperResearch Paper
Research Paper
 
Applied surface science
Applied surface scienceApplied surface science
Applied surface science
 
Experimental and theoretical investigations of some pyrazolo-pyrimidine deriv...
Experimental and theoretical investigations of some pyrazolo-pyrimidine deriv...Experimental and theoretical investigations of some pyrazolo-pyrimidine deriv...
Experimental and theoretical investigations of some pyrazolo-pyrimidine deriv...
 
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
Sonocatalytic Performance of Fe3O4 Cluster Microspheres Gratiphic Carbon Comp...
 

New J. Chem., 2015,39, 931-937

  • 1. This journal is ©The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 New J. Chem., 2015, 39, 931--937 | 931 Cite this: New J. Chem., 2015, 39, 931 Improved photocatalytic activity in a surfactant- assisted synthesized Ti-containing MOF photocatalyst under blue LED irradiation† Sedigheh Abedi and Ali Morsali* The Ti-incorporated metal–organic framework structure, MIL-125, its amine-functionalized form, MIL-NH2 and its surfactant-assisted version, SMIL-NH2 were synthesized via a solvothermal route. The samples were characterized by XRD, SEM, EDS, UV-vis spectroscopy, and N2 adsorption–desorption measurements. In the presence of a nonionic surfactant, cylindrical SMIL-NH2, with 50–250 nm particle size, was obtained. The photocatalytic properties of the materials were evaluated by deoximation reaction. Compared with MIL-NH2, the nanostructured SMIL-NH2 MOF has more than double the surface area and is able to catalyze the regeneration of carbonyl compounds from various oximes at twice the rate in an energy- efficient photocatalytic system. Furthermore, it has been successfully shown that this new nanostructured photocatalyst is significantly durable under the reaction conditions and can be easily recycled after the reaction, with significant retention of its photocatalytic activity. Introduction Today, the rapid everyday rise of waste associated with fossil fuels is one of the major concerns of ecologists. On the other hand, the rapid reduction in fossil sources of energy has led the attention of many enthusiasts to developing ways of utilizing carbon-neutral energy sources. Extensive efforts have been made to replace these kinds of nonrenewable sources with widely available, abundant solar energy. Photocatalytic systems are one the most promising routes to achieve this worthy global objective. Among them, heterogeneous photocatalytic systems, using the most widely used TiO2-based photocatalysts, have attracted much attention, due to their biological and chemical inertness as well nontoxic and inexpensive characteristics.1–3 The unique and tunable architectural properties of metal– organic frameworks (MOFs) have made them promising candidates for heterogeneous catalysis applications.4,5 The extraordinarily high surface areas, porosities and well-defined pore structures of these materials are the cause of recent demands for their use as catalysts.6–8 Recently, a variety of modified MOFs and their applications in light harvesting9–12 , in driving various photo- catalytic reactions such as H2 production,13–15 O2 production,16 and carbon dioxide reduction,17 in gas separation and storage and in the catalytic transformation of organic molecules, have been developed.18–20 The porous channels in MOF structures can be used as photocatalysis sites. Theoretical calculations show that MOFs are semiconductors or insulators with band gaps between 1.0 and 5.5 eV, which can be altered by changing the degree of conjugation in the ligands.21 Experimental results show that the band gaps of IRMOF-type samples can be tuned by varying the functionality of the linker.22 Among numerous scientific attempts to investigate the unexplored catalytic properties of crystalline MOFs,7,23–30 the development of robust MOFs with the potential for photo- catalytic conversion of light energy into chemical energy has had only limited success. A few investigations have been made into their application in the photocatalytic degradation of organic compounds.9,31 Garcia and coworkers revealed the semiconductor behaviour of MOF-5 and showed the potential performance of this MOF in the photocatalyzed degradation of phenol in aqueous solutions.32 The development of special photocatalytic systems for selective organic transformations is a further step in this field that has been more rarely investi- gated. Very recently, this barrier has been broken using MOFs containing other co-catalysis species, such as metal complex parti- cles. Lin’s group used functionalized dicarboxylic UiO-67 containing phosphorescent [IrIII (ppy)2(dcbpy)]Cl and [RuII (bpy)2(dcbpy)]Cl2 (where ppy is 2-phenylpyridine and bpy is 2,20 -bipyridine) in three photocatalytic organic transformations (the aza-Henry reaction, aerobic amine coupling, and aerobic oxidation of thioanisole).33 A visible-light responsive Ti-based MOF photo- catalyst (Ti-MOF-NH2) with 1.5 wt% Pt deposited into it was also used for the photocatalytic reduction of nitrobenzene Department of Chemistry, Faculty of Sciences, Tarbiat Modares University, P.O. Box 14115-4838, Tehran, Iran. E-mail: morsali_a@modares.ac.ir; Fax: +98 21 8009730; Tel: +98 21 82884416 † Electronic supplementary information (ESI) available. See DOI: 10.1039/ c4nj01536c Received (in Montpellier, France) 9th September 2014, Accepted 5th November 2014 DOI: 10.1039/c4nj01536c www.rsc.org/njc NJC PAPER Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16. View Article Online View Journal | View Issue
  • 2. 932 | New J. Chem., 2015, 39, 931--937 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 under visible-light irradiation. The co-catalytic role of the Pt particles was revealed by ESR measurement and, with optimum amounts of Pt deposited in Ti-MOF-NH2, the reduction reaction was carried out successfully within 70 h under visible light irradiation.34 To our knowledge, there are no reports of using visible-light responsive MOF materials without any additives in selective organic transformations. Since morphology-controlled synthesis of porous materials is very important for their adsorption, separation, and catalytic selectivity,35,36 in this study the changes in the structural properties of the Ti-based MOF photocatalyst, MIL-NH2, which is an amine-functionalized version of MIL-125, prepared by the conventional solvothermal method have been compared with those of the same MOF after surfactant-assisted solvothermal synthesis. Fortunately we could expose the visible light photocatalytic activity of the resultant Ti-based MOF photo- catalyst (Ti-MOF-NH2) in the regeneration of carbonyl com- pounds from oximes (deoximation reaction) in very mild photoreaction conditions under blue LED light irradiation. It will be revealed that the Ti-containing MOF structure synthe- sized in the presence of a nonionic amphiphilic surfactant, Pluronic P123, has a more uniform cylindrical morphology with 50–250 nm particle size and more than double the surface area of that synthesised by the conventional solvothermal method, and that it can consequently catalyze the regeneration of carbonyl compounds from various oximes in an energy-efficient photo- catalytic system. Furthermore, we will successfully show that this nanostructured catalyst is significantly stable under the reaction conditions and can be easily recycled after the reaction, retaining its photocatalytic reactivity. Experimental section Materials Poly(ethylene oxide)–poly(propylene oxide)–poly(ethylene oxide) tri- block copolymer template (EO20PO70EO20, Pluronic P123 from Aldrich), 2-amino-benzenedicarboxylic acid (H2BDC-NH2; C8H7NO4 from Aldrich), 1,4-benzenedicarboxylic acid (H2BDC; C8H6O4, from Merck), tetrapropyl orthotitanate (TPOT; Ti-(OC3H7)4, from Aldrich), N,N-dimethylformamide (DMF, from DAE JUNG Reagents Chemicals), and methanol (CH3OH) were used with- out additional purification. CH3CN (from Merck) was distilled and dried by refluxing in the presence of sodium hydride under an argon atmosphere and kept on 4 Å molecular sieves. All oximes were synthesized according to known procedures37 from the corresponding carbonyl compounds (from Merck), and were characterized by nuclear magnetic resonance spectroscopy (1 H-NMR and 13 C-NMR spectra; see the ESI†). Instrumentation Powder X-ray diffraction (PXRD) data were acquired on a Philips X-Pert diffractometer using Cu Ka radiation. SEM images were captured on a Philips XL-300 instrument. The nitrogen sorption isotherms were collected on a Belsorp (BELMAX, Japan). The GC analyses were performed on an Echrom A90 using a flame ionization detector (FID). NMR spectra were recorded on a Bruker 250 MHz spectrometer. Synthesis of MOF structures Synthesis of MIL-125 and MIL-NH2. MIL-125, or Ti8O8(OH)4- [O2C–C6H4–CO2]6, and MIL-NH2 were prepared according to the solvothermal methods reported by Serre et al.38 and Matsuoka et al. respectively.39 Typically, a mixture of terephthalic acid, or H2BDC, (3 mmol, 498 mg), 2 mmol of TPOT (0.6 ml), DMF (9 ml), and methanol (1.0 ml) was transferred to a Teflon-lined stainless-steel autoclave and heated for 48 h at 423 K under autogenous pressure. Back at room temperature, the white solid was recovered by filtration, washed twice with DMF and dried under air at 80 1C. The free solvent was removed by calcination at 200 1C overnight for 12 hours. MIL-NH2 was synthesized in a similar way with H2BDC-NH2 instead of H2BDC. Synthesis of SMIL-NH2. This structure was synthesized according to a known procedure39 with a little modification. In a typical experiment, 2.0 g of Pluronic P123 (Aldrich, average Mw D 5800), was dissolved in 40 ml of DMF at room temperature. The mixture was stirred magnetically, dissolving the surfactant completely and was subsequently sonicated for 15 minutes. Then, 2 mmol (0.6 ml) of TPOT and 3 mmol (0.54 g) of H2BDC-NH2 were added rapidly and the mixture was stirred at 35 1C for 5 h under an Ar atmosphere. Then, for the aging process, the reaction mixture was kept at 140 1C without stirring for 48 hours. The mixture was cooled to room temperature and the resultant yellow solid was filtered off, washed thoroughly with DMF and dried in air at 80 1C. Both the free solvent and the surfactant molecules were removed by calcination at 200 1C overnight for 12 hours. Photocatalytic reaction using the photocatalytic system A mixture of oxime (20 mg) and catalyst (30 mg) in dry CH3CN (10 ml) was prepared in a two-necked flask. One neck was used for the oxygen inlet and outlet and the other was equipped with a condenser with a bladder at the end. The flask was filled with pure oxygen, and sonicated for 5 minutes. Subsequently, it was irradiated under blue LED lamps (4 Â 3 W). Then the catalyst was filtered off through a 0.4 mm filter and the reaction mixture was analyzed using gas chromatography. After the reactions of different oximes, the catalyst was collected by centrifugation, stirred in 1 : 1 CH3CN : CH3OH for 24 hours and then dried in an oven at 75 1C and denoted as SMIL-NH2R. Results and discussion After the synthesis of MIL-125 (or Ti8O8(OH)4[O2C–C6H4–CO2]6) and MIL-NH2 according to the known procedures,38,39 the surfactant-assisted synthesized MIL-NH2, denoted as SMIL-NH2, was prepared using a solvothermal method in the presence of a nonionic structure-directing agent, Pluronic P123. The powder XRD pattern recorded for the synthesized MIL-125 control sample matches with the literature and the pattern simulated Paper NJC Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16. View Article Online
  • 3. This journal is ©The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 New J. Chem., 2015, 39, 931--937 | 933 based on the X-ray data of the MIL-125 single crystal (Fig. 1a and d). The major diffraction pattern is assigned to the rhombohedral structure. In addition, a comparison of the simulated MIL-125 single crystal and experimental powder X-ray diffraction patterns of both the synthesized MIL-NH2 and the surfactant-assisted synthe- sized structure, which is named SMIL-NH2, confirmed the parent structure estimated from the basic crystallographic data (Fig. 1a–c). As Fig. 2 shows, SEM images of MIL-NH2 and SMIL-NH2 indicate that in the presence of the nonionic block copolymer, Pluronic P123, the angular plate structures of MIL-NH2, with nonuniform crystal sizes between 400 and 1000 nm, clearly convert into the soft plate structures of SMIL-NH2, with smaller crystals of about 50–250 nm. Moreover, more uniform cylindrical structures have been constructed from the surfactant-assisted synthesis system. Recently, the effect of the surfactant-assisted medium on the morphology and particle size of [Cu3(btc)2], known as HKUST-1 has been studied.40 As has been well demonstrated, in the presence of the cationic surfactant CTAB in aqueous solution, the triethylammonium salt of the BTC ligand, btc3À , is surrounded by CTA+ species. Therefore, in the presence of higher amounts of CTAB as an additive, nucleation is retarded significantly, which causes a larger particle size for the final crystals. In contrast, it has been shown that via common routes of MOF synthesis in organic solvents, increased concentration of additives leads to smaller crystal size, due to suppression of crystal growth.40,41 It seems that in the presence of the nonionic surfactant P123, in DMF as a solvent, controlled particle formation of carboxylic acid-based MOFs in solution generally occurs and this affects the ultimate crystal size.41 However, the particle size decrease and morphology alteration observed in the presence of P123 is a result of competition between the surfactant and the ligand for coordination to Ti6+ sites in this medium, instead of the surfactant being assembled to make micelles and play its usual templating role in the solution. Very recently, we showed that in the presence of the same concen- tration of P123 in DMF solvent, but with copper salt and BDC ligand, the precursors of the HKUST-1 structure, the template effect of the nonionic surfactant directed the final structure.42 The typical octahedral morphology of the crystals remained intact while some new ordered domains were constructed within the original framework. Here, the existence of the oxophilic Ti6+ metal centers in the relatively harsh conditions of the solvothermal reactor makes a completely different path for dictating the final MOF structure. A comparison of N2 adsorption–desorption analysis of MIL-NH2 and SMIL-NH2 clearly verified this statement (Fig. 3). As shown in Fig. 3a, the hysteresis loop in the N2 adsorption–desorption isotherm of MIL-NH2 indicates a type IV mesoporous material, and this has completely disappeared in the isotherm of SMIL-NH2, which shows a type I isotherm, characteristic of microporous materials. This observation that is quite the opposite of what was seen in the case of HKUST-1, evidently proving the discrepancy of the surfactant role in the two cases. To investigate the photocatalytic activity of MIL-NH2 and its surfactant-assisted synthesized analogue, SMIL-NH2, regenera- tion of carbonyl compounds from oximes, or the deoximation reaction, was chosen as a probe. This is a useful organic transformation because of its use in the purification and characterization of carbonyl compounds and also its importance in organic synthesis.43–47 This reaction has proceeded in crystal- line TiO2-based photocatalytic systems under UV and also solar light.48,49 Recently, a new binary photocatalyst system including a periodic mesoporous organosilicate (PMO) decorated with amorphous TiO2 was developed for the deoximation reaction of various types of aldoximes and ketoximes under sunlight irradiation.50 Nowadays, the quest for the use of high photon Fig. 1 Powder XRD patterns: (a) simulated from X-ray data of MIL-125 single crystal, (b) MIL-NH2, (c) SMIL-NH2, (d) MIL-125, (e) SMIL-NH2R. Fig. 2 SEM images of solvothermally synthesized MIL-NH2, (a) and (b), and surfactant-assisted synthesized SMIL-NH2, (c) and (d). NJC Paper Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16. View Article Online
  • 4. 934 | New J. Chem., 2015, 39, 931--937 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 efficiency, low voltage of electricity, and high power stability LED lamps51 as a light source has attracted wide attention, owing to their simplified practical aspects as well from the standpoint of environmental and economic concerns for cleaner and cheaper organic transformations.52–55 Fortunately, both MIL-NH2 and SMIL-NH2 could successfully catalyze the deoximation reaction of acetophenone oxime in the presence of blue LED lamps (4 Â 3 W) under an oxygen atmosphere. At the beginning of the reaction, when the oxime concentration is high, in the presence of either MIL-NH2 or SMIL-NH2 the rate of acetophenone production is the same (Fig. 4). Interestingly, after 4 h, the concentration curves for the product formation with respect to time begin to separate and the rate of product formation with SMIL-NH2 is about twice as fast as that with MIL-NH2. Lattice parameters of the two structures have been provided in Table 1. As these clearly show, and as is confirmed by the N2 adsorption– desorption analysis, MIL-NH2 and SMIL-NH2 are entirely dis- similar in pore size distribution and eventually in their surface areas. While MIL-NH2 has a 3.3 nm average mesopore size and a 566 m2 gÀ1 surface area, SMIL-NH2 has a smaller pore size of 2.4 nm and a larger surface area of 1132 m2 gÀ1 , which is more than two times that of the former. Although the fact that both reactions have the same induction period needs more precise investigation, the improved lattice parameters may be one of the reasons for the higher catalytic activity of SMIL-NH2. In confirmation of the SEM results, these data evidently are in contrast with the previous results reported for HKUST-1.42 Here the nonionic surfactant, P123, acts as a capping agent instead of being assembled, making micelles and acting as scaffold for construction of the MOF structure. The discrepancy may be due to the different affinities of the two metal centers, copper and titanium ions, to the hydrophilic head groups of the surfactant in the organic solution. No change was observed in the absence of irradiation, catalyst or oxygen for runs carried out under the same experi- mental conditions (Table 2, entry 2). Since no reaction took place in the presence of MIL-125 as a catalyst (Table 2, entry 3) even after 48 h irradiation, the role of the amine-functionalized versions of the MIL can be described as an antenna effect. UV-vis spectra of MIL-NH2 and SMIL-NH2 are shown in Fig. 5. For both structures, the major absorption edge is similar and in the visible light region extends to around 500 nm, relative to the 350 nm absorption edge reported for MIL-125. The observed distinction is known to be induced by O–Ti ligand-to-metal charge transfer in TiO5(OH) inorganic clusters, apparently influenced by H2BDC being replaced with H2BDC-NH2.6,8 Therefore, the antenna role of the amine-functionalized MILs in the photocatalytic deoximation reaction of oximes in an oxygen atmosphere can be described as in Scheme 1. In this known manner, an electron and a hole are generated from light excitation, and transfer to the oxygen and oxime molecules, respectively.7,56–58 Moreover, a Ti-leaching test has been per- formed after a 10 h reaction of acetophenone oxime under blue Fig. 3 (a) N2 adsorption–desorption analysis and (b) BJH analysis of MIL-NH2, SMIL-NH2 and SMIL-NH2. Fig. 4 Yield-versus-time profiles of the deoximation reaction in the presence of (a) MIL-NH2, (b) SMIL-NH2 and (c) Degussa P25; y-axis: yield of acetophenone in %, x-axis: time in hours. Table 1 Lattice parameters of MIL-NH2 structures prepared in the absence or presence of P123 Sample SBET a (m2 gÀ1 ) Dmicro b (nm) Dmeso c (nm) Vmeso c (cm3 gÀ1 ) MIL-NH2 566 0.7 3.3 0.179 SMIL-NH2 1132 0.6 2.4 0.151 SMIL-NH2Rd 1066 0.6 2.4 0.151 a The Brunauer–Emmett–Teller (BET) surface area. b The micropore size distributions were calculated using an MP plot. c The mesopore volume and the mesopore size distributions were calculated using the Barrett–Joyner–Halenda (BJH) method. d The recycled catalyst. Paper NJC Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16. View Article Online
  • 5. This journal is ©The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 New J. Chem., 2015, 39, 931--937 | 935 LED light irradiation. The ICP analysis of the reaction solution showed that no Ti species has been leached, which validates the stability of the MOF structure as well as the heterogeneous character of our catalytic system. The remarkable photocatalytic activity of SMIL-NH2 in such mild reaction conditions encouraged us to examine its photo- catalytic ability for other structurally diverse oximes. Since photo- catalytic deoximation has been known to proceed via a cationic intermediate formed via the reaction of the oxime substrate with photogenerated holes,49 activated ketoximes with electron- donating substituents such as methyl and methoxy were equally well converted into the corresponding ketones in relatively short times under mild reaction conditions (Table 2, entries 4 and 5). Very recently, a deoximation reaction of 4-methoxyacetophenone oxime to the corresponding ketone in the presence of C–N codoped sonofabricated crystalline titania (anatase) as well as a sample of bicrystalline framework titania, (anatase/brookite, 78/22) was examined under blue LED irradiation.59 In similar reaction conditions to our photocatalytic system, with the former catalyst the reaction proceeded quantitatively within 2 h, Table 2 Regeneration of carbonyl compounds from oximes in blue LED light irradiation using SMIL-NH2 Entry Substrate Product Time (h) Yielda (%) 1 10 100 2 48b — 12c 100 3d 48 — 4 6 100 5 4.5 100 4.5e 86 6 48 57 7 48 74 8 24 100 9 24 100 10 48 33 11 48 38 a GC yield, reaction conditions: oxime (20 mg), catalyst (30 mg), 10 ml dry CH3CN, O2 atmosphere (1 atm), blue LED light (4 Â 3 W lamps). b In the dark, in the absence of O2 and/or catalyst the same results were obtained. c In the presence of the recycled catalyst, SMIL-NH2R. d In the presence of MIL-125 as a catalyst. e In the presence of P25 as a catalyst. Fig. 5 UV-vis spectra of (a) MIL-NH2 and (b) SMIL-NH2. Scheme 1 The antenna role of MIL-NH2 in transferring the generated electron and hole to the oxygen and oxime molecules during light excitation. NJC Paper Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16. View Article Online
  • 6. 936 | New J. Chem., 2015, 39, 931--937 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 whereas just 31 percent of the corresponding carbonyl compound was obtained with the latter catalyst. With the most reactive commercial crystalline TiO2, Degussa P25 (anatase/rutile, 75/25), 86 percent of 4-methoxyacetophenone was obtained in the same reaction conditions mentioned in Table 2, in comparison with a quantitative amount of the product formed in the presence of SMIL-NH2 within 4.5 h irradiation (Table 2, entry 5). It is worth mentioning that there is no report of the deoximation reaction with MOFs in such mild reaction conditions. Moreover, the crystalline structure of these materials is completely different and there are no diffraction peaks of any TiO2 crystalline phase (anatase, rutile or brookite) in their structure. Meanwhile, after 48 h irradiation, deactivated oximes containing electron withdrawing groups such as 4-chloroacetophenone oxime, 3-bromoacetophenone oxime and the more passive substrate, methyl 2-pyridyl ketone oxime gave weak to moderate yields of the corresponding carbonyl products (Table 2, entries 6, 7 and 11). Furthermore, the reaction of sterically hindered oxime substrates proceeded over a long time and was not completed in the case of deactivated substrates such as (4-chlorophenyl)(phenyl)methanone oxime (Table 2, entries 9 and 10). This observation shows the preference of the substrate to pass into the channels of the photoactive MOF framework rather than to react on the surface. In other words, the reaction proceeds within the channels of the MOF framework. For practical purposes, recycling of the catalyst is an important problem to solve. Considering the potential application of SMIL-NH2, we collected the catalyst used in the deoximation reactions of diverse substrates, thoroughly washed it with 1 :1 of CH3CN and MeOH, and subsequently dried it in an oven at 80 1C for 12 h. A powder XRD pattern of this recycled catalyst, denoted as SMIL-NH2R, is shown in Fig. 1e, and obviously indicates that the crystallinity of the MOF structure has been reduced slightly, while the original XRD pattern of the structure remains intact and matches with the X-ray data of the MIL-125 single crystal. The deoximation reaction of acetophenone oxime in the presence of SMIL-NH2R in the same optimum conditions mentioned in Table 2 gave a quantitative yield of acetophenone after 12 h irradiation (Table 2, entry 2). SEM images of SMIL-NH2R before and after washing the catalyst, shown in Fig. S1 (ESI†), not only verify the importance and effectiveness of the washing stage to disconnect the agglomerated particles of the MOF structure formed during the reaction, but also confirm the basic morpho- logy remains nearly intact after the deoximation reaction. In addition, some residual nitrogen content (about 7%) seen in the energy dispersive X-ray analysis (EDS) of the unwashed catalyst, which has been eliminated in the EDS analysis of the washed catalyst, clearly validates the SEM results (Fig. S1 and S2, ESI†). Accordingly, partial agglomeration of the cylindrical structures of SMIL-NH2 and trapping of a few organic molecules in the MOF structure (existence of carbon content in the EDS of SMIL-NH2R, Fig. S3, ESI†), can occupy active Ti centers and reduce the photocatalytic activity of the catalyst. It is also worth mentioning that despite the structural changes observed, SMIL-NH2R is more reactive than MIL-NH2. A comparison of the porosity parameters shown in Table 1 shows that just a little decrease in the surface area of SMIL-NH2R relative to SMIL-NH2 does not affect its activity, which is still greater than that of MIL-NH2. Conclusion We succeeded in the preparation of a new morphology of MIL-NH2 in the presence of Pluronic P123, as a structure- directing agent and in solvothermal conditions. Adsorption– desorption isotherm, BET, and SEM analyses clearly established the effectiveness of the procedure for the synthesis of uniform cylindrical SMIL-NH2 nanoparticles possessing double the surface area of MIL-NH2, with 50–250 nm particle sizes. Furthermore, regeneration of carbonyl compounds from various oximes in an energy-efficient photocatalytic system was successfully catalyzed by this stable nanostructured photocatalyst in an O2 atmosphere and under blue LED light irradiation. In addition, it was shown that this nanostructured MOF catalyst is stable in the reaction conditions and can be easily recycled after the reaction, retaining its photocatalytic reactivity. Acknowledgements This work was supported by Tarbiat Modares University. Notes and references 1 A. Mills and S. Le Hunte, J. Photochem. Photobiol., A, 1997, 108, 1–35. 2 N. J. Peill, L. Bourne and M. R. Hoffmann, J. Photochem. Photobiol., A, 1997, 108, 221–228. 3 M. R. Hoffmann, S. T. Martin, W. Choi and D. W. Bahnemann, Chem. Rev., 1995, 95, 69–96. 4 R. J. Kuppler, D. J. Timmons, Q.-R. Fang, J.-R. Li, T. A. Makal, M. D. Young, D. Yuan, D. Zhao, W. Zhuang and H.-C. Zhou, Coord. Chem. Rev., 2009, 253, 3042–3066. 5 I. Fechete, Y. Wang and J. C. Ve´drine, Catal. Today, 2012, 189, 2–27. 6 S.-N. Kim, J. Kim, H.-Y. Kim, H.-Y. Cho and W.-S. Ahn, Catal. Today, 2013, 204, 85–93. 7 J. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. T. Nguyen and J. T. Hupp, Chem. Soc. Rev., 2009, 38, 1450–1459. 8 N. D. McNamara, G. T. Neumann, E. T. Masko, J. A. Urban and J. C. Hicks, J. Catal., 2013, 305, 217–226. 9 J.-L. Wang, C. Wang and W. Lin, ACS Catal., 2012, 2, 2630–2640. 10 C. A. Kent, D. Liu, L. Ma, J. M. Papanikolas, T. J. Meyer and W. Lin, J. Am. Chem. Soc., 2011, 133, 12940–12943. 11 C. A. Kent, D. Liu, T. J. Meyer and W. Lin, J. Am. Chem. Soc., 2012, 134, 3991–3994. 12 K. C. Stylianou, R. Heck, S. Y. Chong, J. Bacsa, J. T. A. Jones, Y. Z. Khimyak, D. Bradshaw and M. J. Rosseinsky, J. Am. Chem. Soc., 2010, 132, 4119–4130. 13 C. Gomes Silva, I. Luz, F. X. Llabre´s i Xamena, A. Corma and H. Garcı´a, Chem. – Eur. J., 2010, 16, 11133–11138. Paper NJC Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16. View Article Online
  • 7. This journal is ©The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015 New J. Chem., 2015, 39, 931--937 | 937 14 A. Fateeva, P. A. Chater, C. P. Ireland, A. A. Tahir, Y. Z. Khimyak, P. V. Wiper, J. R. Darwent and M. J. Rosseinsky, Angew. Chem., 2012, 124, 7558–7562. 15 C. Wang, K. E. deKrafft and W. Lin, J. Am. Chem. Soc., 2012, 134, 7211–7214. 16 C. Wang, J.-L. Wang and W. Lin, J. Am. Chem. Soc., 2012, 134, 19895–19908. 17 Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu and Z. Li, Angew. Chem., Int. Ed., 2012, 51, 3364–3367. 18 S. Horike and S. Kitagawa, Metal-Organic Frameworks, Wiley-VCH Verlag GmbH & Co. KGaA, 2011, pp. 1–21. 19 Y.-X. Tan, Y.-P. He and J. Zhang, Chem. Mater., 2012, 24, 4711–4716. 20 M. C. Das, S. Xiang, Z. Zhang and B. Chen, Angew. Chem., Int. Ed., 2011, 50, 10510–10520. 21 M. Fuentes-Cabrera, D. M. Nicholson, B. G. Sumpter and M. Widom, J. Chem. Phys., 2005, 123, 124713. 22 J. Gascon, M. D. Herna´ndez-Alonso, A. R. Almeida, G. P. M. van Klink, F. Kapteijn and G. Mul, ChemSusChem, 2008, 1, 981–983. 23 A. Corma, H. Garcı´a and F. X. Llabre´s i Xamena, Chem. Rev., 2010, 110, 4606–4655. 24 S. Horike, M. Dinca˘, K. Tamaki and J. R. Long, J. Am. Chem. Soc., 2008, 130, 5854–5855. 25 F. Song, C. Wang, J. M. Falkowski, L. Ma and W. Lin, J. Am. Chem. Soc., 2010, 132, 15390–15398. 26 M. Yoon, R. Srirambalaji and K. Kim, Chem. Rev., 2012, 112, 1196–1231. 27 K. Leus, G. Vanhaelewyn, T. Bogaerts, Y.-Y. Liu, D. Esquivel, F. Callens, G. B. Marin, V. Van Speybroeck, H. Vrielinck and P. Van Der Voort, Catal. Today, 2013, 208, 97–105. 28 G. Calleja, R. Sanz, G. Orcajo, D. Briones, P. Leo and F. Martı´nez, Catal. Today, 2014, 227, 130–137. 29 L. Chen, Z. Gao and Y. Li, Catal. Today, DOI: 10.1016/ j.cattod.2014.03.074. 30 E. E. Macias, P. Ratnasamy and M. A. Carreon, Catal. Today, 2012, 198, 215–218. 31 L. Wen, L. Zhou, B. Zhang, X. Meng, H. Qu and D. Li, J. Mater. Chem., 2012, 22, 22603–22609. 32 M. Alvaro, E. Carbonell, B. Ferrer, F. X. Llabre´s i Xamena and H. Garcia, Chem. – Eur. J., 2007, 13, 5106–5112. 33 C. Wang, Z. Xie, K. E. deKrafft and W. Lin, J. Am. Chem. Soc., 2011, 133, 13445–13454. 34 T. Toyao, M. Saito, Y. Horiuchi, K. Mochizuki, M. Iwata, H. Higashimura and M. Matsuoka, Catal. Sci. Technol., 2013, 3, 2092–2097. 35 Y. Wan, Y. Shi and D. Zhao, Chem. Mater., 2008, 20, 932–945. 36 S. G. Rudisill, Z. Wang and A. Stein, Langmuir, 2012, 28, 7310–7324. 37 B. S. Furniss, Vogel’s textbook of practical organic chemistry, Pearson Education India, 1989. 38 M. Dan-Hardi, C. Serre, T. Frot, L. Rozes, G. Maurin, C. Sanchez and G. Fe´rey, J. Am. Chem. Soc., 2009, 131, 10857–10859. 39 Y. Horiuchi, T. Toyao, M. Saito, K. Mochizuki, M. Iwata, H. Higashimura, M. Anpo and M. Matsuoka, J. Phys. Chem. C, 2012, 116, 20848–20853. 40 T. Uemura and S. Kitagawa, Chem. Lett., 2005, 34, 132–137. 41 S. Hermes, T. Witte, T. Hikov, D. Zacher, S. Bahnmu¨ller, G. Langstein, K. Huber and R. A. Fischer, J. Am. Chem. Soc., 2007, 129, 5324–5325. 42 S. Abedi and A. Morsali, ACS Catal., 2014, 4, 1398–1403. 43 D. H. R. Barton, J. M. Beaton, L. E. Geller and M. M. Pechet, J. Am. Chem. Soc., 1961, 83, 4076–4083. 44 D. H. R. Barton and J. M. Beaton, J. Am. Chem. Soc., 1961, 83, 4083–4089. 45 R. S. Varma, Green Chem., 1999, 1, 43–55. 46 W. Chrisman, M. J. Blankinship, B. Taylor and C. E. Harris, Tetrahedron Lett., 2001, 42, 4775–4777. 47 A. Khazaei, R. Ghorbani Vaghei and M. Tajbakhsh, Tetra- hedron Lett., 2001, 42, 5099–5100. 48 C. Saravanselvi, N. Somasundaram, S. Vijaikumar and C. Srinivasan, Photochem. Photobiol. Sci., 2002, 1, 607–608. 49 K. Muthu, K. Selvam, B. Krishnakumar and M. Swaminathan, Appl. Catal., A, 2009, 358, 259–263. 50 S. Abedi, B. Karimi, F. Kazemi, M. Bostina and H. Vali, Org. Biomol. Chem., 2013, 11, 416–419. 51 N. G. Yeh, C.-H. Wu and T. C. Cheng, Renewable Sustainable Energy Rev., 2010, 14, 2161–2166. 52 S. Fu¨ldner, R. Mild, H. I. Siegmund, J. A. Schroeder, M. Gruber and B. Ko¨nig, Green Chem., 2010, 12, 400–406. 53 S. Fu¨ldner, T. Mitkina, T. Trottmann, A. Frimberger, M. Gruber and B. Ko¨nig, Photochem. Photobiol. Sci., 2011, 10, 623–625. 54 S. Fu¨ldner, P. Pohla, H. Bartling, S. Dankesreiter, R. Stadler, M. Gruber, A. Pfitzner and B. Ko¨nig, Green Chem., 2011, 13, 640–643. 55 Z. Zand, F. Kazemi and S. Hosseini, Tetrahedron Lett., 2014, 55, 338–341. 56 D. Farrusseng, S. Aguado and C. Pinel, Angew. Chem., Int. Ed., 2009, 48, 7502–7513. 57 U. Ravon, M. E. Domine, C. Gaudillere, A. Desmartin- Chomel and D. Farrusseng, New J. Chem., 2008, 32, 937–940. 58 L. Alaerts, E. Se´guin, H. Poelman, F. Thibault-Starzyk, P. A. Jacobs and D. E. De Vos, Chem. – Eur. J., 2006, 12, 7353–7363. 59 M. Abdollahi Kakroudi, F. Kazemi and B. Kaboudin, J. Mol. Catal. A: Chem., 2014, 392, 112–119. NJC Paper Publishedon20November2014.DownloadedbyTarbiatModarresUniversityon20/10/201513:56:16. View Article Online