You are on page 1of 11

Dalton

Transactions
View Article Online
PAPER View Journal | View Issue

Amine-functionalized Zn(II) MOF as an efficient


Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

Cite this: Dalton Trans., 2018, 47,


multifunctional catalyst for CO2 utilization and
8041 sulfoxidation reaction†
Parth Patel, a,c Bhavesh Parmar, b,d Rukhsana I. Kureshy, a,d

Noor-ul H. Khan *a,c,d and Eringathodi Suresh *b,d

Herein, a zinc(II)-based 3D mixed ligand metal organic framework (MOF) was synthesized via versatile
routes including green mechanochemical synthesis. The MOF {[Zn(ATA)(L)·H2O]}n (ZnMOF-1-NH2) has
been characterized by various physico-chemical techniques, including SCXRD, and composed of the
bipyridyl-based Schiff base (E)-N’-( pyridin-4-ylmethylene)isonicotinohydrazide (L) and 2-amino-
terephthalic acid (H2ATA) ligands as linkers. The MOF material has been explored as a multifunctional
heterogeneous catalyst for the cycloaddition of alkyl and aryl epoxides with CO2 and sulfoxidation reac-
tions of aryl sulfides. The influence of various reaction parameters is examined to optimize the perform-
ance of the catalytic reactions. It is found that solvent-free catalytic reaction conditions offer good cata-
lytic conversion in the case of cyclic carbonates, and for sulfoxide, good conversion and selectivity are
Received 3rd April 2018, achieved in the presence of DCM as a solvent medium under ambient reaction conditions. The chemical
Accepted 19th May 2018
and thermal stability of the catalyst are excellent and it is active for up to four catalytic cycles without sig-
DOI: 10.1039/c8dt01297k nificant loss in activity. Furthermore, based on the catalytic activity and structural evidence, a plausible
rsc.li/dalton mechanism for both catalytic reactions is proposed.

Introduction the utilization of CO2 as a chemical feedstock to produce valu-


able chemicals has been proposed as a sustainable approach.
The quest for green technologies that utilize the CO2 emission Therefore, significant efforts are currently being devoted to the
from industries and power plants and facilitate sulfoxidation development of appropriate materials that can selectively
of organic compounds to value-added chemical intermediates capture CO2 and facilitate catalytic conversion to useful pro-
by appropriate catalytic reactions is a prime area of research ducts through chemical fixation.10–14 The coupling reaction of
with respect to the greenhouse effect and human healthcare, CO2 with epoxides to provide cyclic carbonates in the presence
respectively.1–3 Recently, the development of adaptable cata- of different catalysts is considered as one of the best strategies
lysts that promote the transformation of CO2 to cyclic carbon- for CO2 conversion due to its excellent reactivity and high
ates has increased, which are necessary for a green and sus- atom efficiency. Cyclic carbonates are well-known and have sig-
tainable environment and the oxidation of organic sulfides to nificant and extensive applications as aprotic polar solvents,
value-added drug intermediates in the pharmaceutical precursors affording polycarbonates, and chemical intermedi-
industry.4–9 To negate the adverse effects of greenhouse gases, ates in organic synthesis.15–26 Many homogeneous catalysts
have been developed for the cycloaddition reaction of CO2 and
epoxides and sulfoxidation reactions with good activity and
a
Inorganic Materials and Catalysis Division, CSIR-Central Salt and Marine efficiency.27–39 However, their complex separation, purification,
Chemicals Research Institute, G. B. Marg, Bhavnagar-364 002, Gujarat, India. and recycling in homogenous catalysis severely limit their
E-mail: nhkhan@csmcri.res.in
b application in industry. Thus, robust, recyclable multifunc-
Analytical and Environmental Science Division and Centralized Instrument Facility,
CSIR-Central Salt and Marine Chemicals Research Institute, G. B. Marg, Bhavnagar- tional heterogenous catalysts with good chemical and thermal
364 002, Gujarat, India. E-mail: esuresh@csmcri.res.in, sureshe123@rediffmail.com stability that can function efficiently under mild reaction con-
c
Charotar University of Science & Technology, Changa-388 421, Anand, Gujarat, ditions are necessary.40 Materials with features such as high
India surface area and porosity, good adsorption capacity, adequate
d
Academy of Scientific and Innovative Research (AcSIR), CSIR-Central Salt and
Lewis/Brønsted acidic or basic sites, and high chemical and
Marine Chemicals Research Institute, G. B. Marg, Bhavnagar-364 002, Gujarat, India
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ thermal stability upon exposure to the reaction conditions
c8dt01297k including substrates favor efficient heterogenous catalytic reac-

This journal is © The Royal Society of Chemistry 2018 Dalton Trans., 2018, 47, 8041–8051 | 8041
View Article Online

Paper Dalton Transactions

tions. Several types of heterogeneous catalysts based on porous tions, conventional refluxing, and green mechanochemical
materials such as zeolites, silica-supported salts, and micro- (grinding) reactions using bipyridyl-based Schiff base, (E)-N′-
porous organic polymers (MOPs) have been explored for the ( pyridin-4-ylmethylene)isonicotinohydrazide (L) and amino
synthesis of cyclic carbonates and sulfoxidation reactions.41–48 functionalized 2-aminoterephthalic acid (H2ATA) as linkers,
As a new type of crystalline porous material, metal organic and its catalytic performance. The uniformly distributed Zn(II)
frameworks (MOFs) are constructed from multidentate organic Lewis acidic sites and amino/amide functional groups in the
building blocks linking metal or metal clusters, which display ligand moiety on the pore walls render ZnMOF-1-NH2 an
periodic architectures and inherent porosity and have a wide efficient heterogeneous catalyst for the chemical fixation of
range of applications. The captivating uniqueness of MOFs is CO2 to cyclic carbonates and sulfoxidation of organic sulfides
Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

their tunability and tailored structures, which are obtained under mild reaction conditions. Interestingly, the ZnMOF-1-
through a delicate design of molecular building units in the NH2 catalyst can be synthesized in good yield, crystallinity and
form of organic linkers and the coordination geometry around phase purity via the green mechano-chemical method.
the metal nodes. MOFs have emerged as porous frameworks
for various applications due to their robust/high thermal stabi-
lity, large surface area and tunable composition for the adsorp- Results and discussion
tion and separation of gas molecules, including the green-
house gas CO2.49–54 Recently, MOFs have also been revealed as Characterization and structural features of ZnMOF-1-NH2
efficient catalysts for heterogeneous catalysis reactions.55–60 The physical characterization of ZnMOF-1-NH2 synthesized by
Nevertheless, the use of MOFs as heterogeneous catalysts for adaptable routes to elucidate its bulk phase purity and
CO2 conversion and sulfoxidation reactions in particular has detailed structural analysis by SCXRD have been reported by
remained unexploited. Park et al. and Ma et al. conducted pio- us elsewhere.97 The conventional synthetic approach for MOFs
neering work in the area of MOF-catalyzed CO2 fixation reac- is based on reflux and solvothermal methods, which often
tions at ambient reaction conditions to yield value-added lead to the issue of solvent molecules remaining in the
intermediates.61–72 Also, Ma et al. and Jiang’s group reported network and low yield. Mechanochemistry (milling or grind-
CO2 cycloaddition under mild conditions (at room tempera- ing) offers an alternative green approach for the preparation of
ture and under 1 bar CO2 pressure) using an MOF with the nbo MOFs using metal salts and ligand precursors in an appropri-
topology and Al-based MOF USTC-253-TFA, respectively.73,74 ate stoichiometry.98,99 In addition, mechanochemical (milling
One of the most important catalytic reactions is the oxidation or grinding) reactions have advantages such as low energy
of sulfides to sulfoxides and sulfones, which has potential in requirement, less time consuming, minimal solubility issues
the chemical industry, biology, and medical chemistry.75–79 and environmental friendliness for the synthesis MOF
There are many reports on the H2O2-based oxidation of sul- materials with good yield and phase purity. Our synthetic strat-
fides by homogeneous organo-catalysts and heterogeneous egy for the bulk catalytic material (ZnMOF-1-NH2) involves
polyoxometalates.80–84 However, the heterogeneous oxidative adaptable approaches, which includes conventional reflux and
desulfurization of sulfides by MOF catalysts is scarcely a green mechanochemical method. The details of the MOF
reported in the literature, and most of these reactions involve synthesis via different routes and their characterization using
polyoxometalates (POM) encapsulated metal organic frame- powder X-ray diffraction (PXRD), FTIR and TGA are provided in
works (MOF), although enantioselective sulfoxidation has the ESI.†
been studied using chiral MOFs as catalysts.85–92 Therefore, The phase purity/bulk homogeneity and chemical/thermal
the development of efficient oxidation catalysts for various stability of the pristine as-synthesized catalyst and that recov-
types of sulfides with high yield and selectivity to sulfoxide is ered after 4 catalytic cycles were characterized via PXRD, FTIR,
still necessary. FE-SEM and TGA analysis. The simulated PXRD data of the
MOF ligands with functional moieties such as amino/ CIF file from the CCDC database matches well with the experi-
amide groups aligned on their pore walls can act as active mental PXRD pattern of the MOFs synthesized via different
sites, which can increase their affinity to the substrate/CO2 in routes, suggesting the bulk phase purity of the catalysts
catalytic reactions via supramolecular interactions. (Fig. S1†). As depicted in Fig. S2,† the FTIR data of both the
Additionally, the presence of active Lewis acidic sites from the pristine and recovered catalyst exhibits good agreement, which
metal ions uniformly distributed in MOFs can also enhance demonstrates the structural and chemical stability of
their catalytic efficiency.93–96 The robust nature of MOFs and ZnMOF-1-NH2 after the catalytic reaction, and its TGA profile
their chemical and thermal stability impart excellent recycl- indicates thermal stability up to ∼280–310 °C (Fig. S3†). The
ability to these materials. From the above background and per- FE-SEM images depicted in Fig. 1 clearly reveal a well-defined
spective, we envision that rationally designed MOF materials morphology in the crystalline material for both the pristine
can serve as potential heterogeneous catalysts for the chemical MOF and that recovered after 4 catalytic cycles, which retained
conversion of CO2 to cyclic carbonates and sulfoxidation reac- the textural features of the as-synthesized catalyst.
tion. Herein, we report the synthesis of a mixed-ligand three- The gas sorption analysis of the activated ZnMOF-1-NH2 in
dimensional MOF {[Zn(ATA)(L)]·H2O}n (ZnMOF-1-NH2) via ver- vacuum at 120 °C reveals its selectively towards CO2 over N2
satile synthetic routes, including diffusion of precursor solu- (Fig. S4†). Due to their structural flexibility, 2D Zn(II) MOFs

8042 | Dalton Trans., 2018, 47, 8041–8051 This journal is © The Royal Society of Chemistry 2018
View Article Online

Dalton Transactions Paper


Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

Fig. 1 FE-SEM images of ZnMOF-1-NH2 (a) conventionally synthesized,


(b) mechanochemically synthesized, (c) recovered after 4 catalytic
cycles of cycloaddition, and (d) recovered after 4 catalytic cycles of
sulfoxidation reaction.

belonging to the same family but having different substituents


in the dicarboxylic acid and the same N-donor Schiff base
ligand have been reported to show very interesting sorption
properties and exhibit selective adsorption of CO2 over N2. The
Pore size, polarity and kinetic issues can influence the sorp-
Fig. 2 (a) Coordination environment around the Zn metal centre in
tion properties of MOFs for the preference of CO2 over N2
ZnMOF-1-NH2, (b) 2D [Zn2(ATA)2]n sql sheets of ZnMOF-1-NH2 showing
adsorption. The polar amide functional group of the acyl the amino decorated dicarboxylate moiety coordinated with the Zn(II)
hydrazone Schiff base ligand/–NH2 from the ATA ligand of the metal centre and (c) packing along the b-axis showing the 2-fold inter-
framework can interact with the high quadrupole moment of penetrated 3D network in ZnMOF-1-NH2.
CO2.100,101 The adsorption analysis reveals that the activated
ZnMOF-1-NH2 adsorbs CO2 since it has a smaller kinetic dia-
meter (3.30 Å) than N2 (3.64 Å), and no significant uptake of N-donor Schiff base ligand (Fig. 2a). As depicted in Fig. 2b,
N2 is observed. As depicted in (Fig. S4†), the CO2 adsorption dimeric clusters of [Zn2(COO)2] secondary building units are
curve with hysteresis in the present case can be attributed to linked by the carboxylate oxygens of symmetrically disposed
the guest-induced framework response. The BET surface area ATA ligands, creating two-dimensional [Zn2(ATA)2]n sql sheets.
and Langmuir surface area of ZnMOF-1-NH2 are 1.1554 m2 g−1 Double-pillaring of the [Zn2(ATA)2]n sql net by the terminal
and 0.4612 ± 0.1069 m2 g−1, respectively. nitrogen atoms of the bipyridyl-based N-donor Schiff base
The 3D frameworks with interpenetration can self-orient ligand result in a three-dimensional network with the pcu
themselves upon supramolecular interaction between the topology for ZnMOF-1-NH2 (Fig. 2c). The effective voids of the
polar CO2 groups with the arylhydrazone moiety of the Schiff 3D framework are reduced due to the two-fold interpenetration
base ligand/NH2 group of the dicarboxylate decorating the of the 3D nets via N–H⋯O hydrogen bonding between the
framework pores, which favors better affinity for CO2 than N2 amide hydrogen of L and carboxylate oxygen of ATA, and prob-
and the hysteresis observed in the adsorption profile. The ably the lattice water occupy the voids by supramolecular inter-
measured CO2 uptake at 273 K is 1.4181 mmol of CO2 per actions. Penta coordination with the coordinatively unsatu-
gram of MOF at the relative pressure of 0.99 (P/P0), which rated Zn(II) metal center gives an active catalytic site and the
favors the cycloaddition reaction in the present investigation. amide/amine functional groups of the ligand moiety may also
The acidity (amount and strength) of the catalyst was deter- favour the catalytic performance of the MOF material via
mined by NH3-TPD measurements (Fig. S29†). The NH3-TPD efficient supramolecular interactions with the substrate
experiment shows that NH3 is desorbed in the temperature (Fig. 2a).
range of 50–400 °C with a major peak at 315 °C (desorbed
amount of NH3 is 27.081 mmol g−1). The crystal structure ana-
lysis reveals that ZnMOF-1-NH2 crystalized in a triclinic system Cycloaddition of CO2 with epoxides
with the P1ˉ space group and formed a 3D framework, in which The cycloaddition of CO2 with styrene oxide (SO) was selected
the Zn(II) metal nodes are linked via the dicarboxylate and as a model reaction to optimize the reaction conditions. Our

This journal is © The Royal Society of Chemistry 2018 Dalton Trans., 2018, 47, 8041–8051 | 8043
View Article Online

Paper Dalton Transactions

preliminary catalytic experiments on the cycloaddition of chemically prepared ZnMOF-1-NH2 for the cycloaddition reac-
carbon dioxide and epoxides by ZnMOF-1-NH2 were focused tion of CO2 with SO was conducted, which showed no signifi-
on mainly establishing the reaction parameters in the presence cant deviation in the product yield (Table 1, entry 11).
of a co-catalyst, including tert-butyl ammonium bromide To explore the optimal reaction conditions, a series of reac-
(TBAB), tert-butyl ammonium iodide (TBAI) and potassium tions were designed with an incremental variation in CO2
iodide (KI), using styrene oxide (SO) as a model substrate pressure, temperature, reaction time and catalyst amount to
under solvent-free conditions. The reactions were performed at obtain the best possible combination. The effect of an increase
a CO2 pressure of 0.8/1.0 MPa and temperature of 60/80 °C, in CO2 on the ZnMOF-1-NH2-catalyzed SO-CO2 cycloaddition is
with the reaction mixture containing SO (20 mmol), catalyst depicted in Fig. 3. The conversion of SO steadily increased up
Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

(0.21 mmol) and the respective co-catalyst (0.5 mmol), in a to 32% at a CO2 pressure of 0.4 MPa and faster styrene carbon-
50 mL stainless-steel autoclave with magnetic stirring at 600 ate formation was observed in the pressure range of 0.4–0.8
rpm. Details of the reaction conditions and parameters are MPa with a maximum of 95% conversion, with no further sig-
given in Table 1. In the controlled experiments with the nificant conversion above 0.8 MPa. The improved SO conver-
absence of catalyst/co-catalyst and presence of only TBAB, SO sion can be attributed to the enhancement in the solubility of
conversion was negligible (Table 1, entries 1 and 2, respect- CO2 in the liquid phase in the presence of the heterogeneous
ively), while the catalytic reaction with pristine MOF yielded catalyst with pressure, which is in agreement with the previous
38% styrene carbonate (SC) within 8 h reaction time (Table 1, report by Han et al.104
entry 3). Keeping the CO2 pressure at 0.8 MPa, varying reaction
times and temperatures were also monitored for the ZnMOF-1-
NH2/TBAB system, and a maximum of 95% SO conversion was
achieved at 80 °C with 8 h reaction time (Table 1, entries 4–7).
When the pressure was increased up to 1.0 MPa while keeping
the other reaction conditions constant (Table 1, entry 10), a
slight reduction in product yield was observed. Notably, the
conversion of SO is comparatively lower when TBAB is replaced
by TABI or KI as the co-catalyst (Table 1, entries 9 and 10,
respectively). Under the same reaction conditions using
2.5 mol% of the respective co-catalyst, the product yield
follows the order of TBAB > KI > TBAI, where the % yield of
product (SO to SC) is 88%, 85% and 81% (Table 1, entries 7, 9
and 10), respectively.
Higher activity was observed with the bromide salts com-
pared to the iodine salts in contrast to the literature reports,
where iodide was found to be the best promoter (in accordance
with the increase in nucleophilicity). Our results may be
because the bigger size of the iodide hindered diffusion in the
Fig. 3 Effect of pressure (MPa) on styrene carbonate formation
reaction process.102,103 Thus, we can deduce that the co-cata-
(Reaction conditions: SO = 20 mmol (2.28 mL at 25 °C), catalyst mol%:
lyst, TBAB, has an important role in the catalytic reaction, ZnMOF-1-NH2 = 1.0 mol% (0.21 mmol, 0.1 g); tetrabutylammonium
which may be attributed to the presence of the labile Br− bromide (TBAB) = 2.5 mol% (0.5 mmol, 0.161 g), 80 °C, 8 h, and 600
leaving group. A representative reaction for the mechano- rpm).

Table 1 Cycloaddition of styrene oxide and CO2 for styrene carbonatea

Entry Catalyst/co-catalyst PCO2 (MPa) Temp. (°C) Time (h) SO conversion (%) SC selectivity (%)

1 None 0.8 80 8 11 99
2 TBAB 0.8 80 8 14 99
3 ZnMOF-1-NH2 0.8 80 8 38 99
4 ZnMOF-1-NH2/TBAB 0.8 60 4 72 98
5 ZnMOF-1-NH2/TBAB 0.8 60 6 78 98
6 ZnMOF-1-NH2/TBAB 0.8 80 6 88 97
7 ZnMOF-1-NH2/TBAB 0.8 80 8 95 96
8 ZnMOF-1-NH2/TBAB 1.0 80 8 93 97
9 ZnMOF-1-NH2/TBAI 0.8 80 8 86 97
10 ZnMOF-1-NH2/KI 0.8 80 8 90 98
11 ZnMOF-1-NH2G/TBAB 0.8 80 8 94 96
a
Reaction conditions: SO = 20 mmol (2.28 mL at 25 °C), 600 rpm. Catalyst mol%: ZnMOF-1-NH2 = 1.0 mol% (0.21 mmol, 0.1 g); tetrabutyl-
ammonium bromide (TBAB) = 2.5 mol% (0.5 mmol, 0.161 g).

8044 | Dalton Trans., 2018, 47, 8041–8051 This journal is © The Royal Society of Chemistry 2018
View Article Online

Dalton Transactions Paper


Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

Fig. 4 Effect of temperature (°C) on styrene carbonate formation


(Reaction conditions: SO = 20 mmol (2.28 mL at 25 °C), catalyst mol%:
Fig. 6 Effect of catalyst loading (mol%) on styrene carbonate formation
ZnMOF-1-NH2 = 1.0 mol% (0.21 mmol, 0.1 g); tetrabutylammonium
(Reaction conditions: SO = 20 mmol (2.28 mL at 25 °C), tetrabutyl-
bromide (TBAB) = 2.5 mol% (0.5 mmol, 0.161 g), PCO2 = 0.8 MPa, 8 h,
ammonium bromide (TBAB) = 2.5 mol% (0.5 mmol, 0.161 g), PCO2 = 0.8
and 600 rpm).
MPa, 80 °C, 8 h, and 600 rpm).

The effect of temperature on the conversion of styrene car-


of 95–96% at a 1–1.2 mol% catalyst loading. Thus, 1 mol%
bonate is presented in Fig. 4. The efficiency of substrate con-
catalyst was retained as the optimised loading in further cata-
version is improved with temperature, keeping the other para-
lytic experiments.
meters constant. Upon an increase in temperature from 25 °C
The scope and generality of the mentioned series of experi-
to 80 °C, the conversion of SO increased from 25% to finally
ments clearly validate the optimized parameters for the cata-
95% in 8 h of catalytic reaction. As shown in Fig. 5, the conver-
lyst/co-catalyst, temperature and pressure as ZnMOF-1-NH2/
sion of styrene oxide increased with an increase in reaction
TBAB, 8 h, and 0.8 MPa, respectively, in cycloaddition reac-
time and finally a total conversion of 95% was achieved with
tions for different aromatic and aliphatic substrates to yield
in 8 h reaction time, and no further improvement was
the respective cyclic carbonate. Consequently, the coupling
observed upon a further increase in reaction time. The effect
reactions of CO2 with different epoxides catalyzed by
of catalyst loading was also optimized, which is shown in
ZnMOF-1-NH2/TBAB were investigated, and the results are
Fig. 6. Keeping all the other parameters constant, an increase
summarized in Table 2. The catalytic system could convert all
in the amount of catalyst resulted in the maximum conversion

Table 2 Substrate screening using the ZnMOF-1-NH2 catalyst in the


cycloaddition of epoxide with CO2a

Entry Epoxide Conversion (%) Yield (%)

1 93 88

2 96 93

3 91 89

4 92 89

5 88 84
6 73 72

7 18 18

Fig. 5 Effect of reaction time (h) on styrene carbonate formation a


Reaction conditions: Epoxide = 20 mmol, catalyst mol%: ZnMOF-1-
(Reaction conditions: SO = 20 mmol (2.28 mL at 25 °C), tetrabutyl- NH2 = 1.0 mol% (0.21 mmol, 0.1 g); tetrabutylammonium bromide
ammonium bromide (TBAB) = 2.5 mol% (0.5 mmol, 0.161 g), PCO2 = 0.8 (TBAB) = 2.5 mol% (0.5 mmol, 0.161 g), PCO2 = 0.8 MPa, 8 h, 80 °C, and
MPa, 80 °C, 8 h, and 600 rpm). 600 rpm.

This journal is © The Royal Society of Chemistry 2018 Dalton Trans., 2018, 47, 8041–8051 | 8045
View Article Online

Paper Dalton Transactions

the studied epoxides to the corresponding cyclic carbonates posed coupling reaction mechanism is initiated by the attack
effectively under the optimized reaction conditions with good of the oxygen atom from the epoxide with the vacant coordina-
yield and selectivity (Table 2, entries 1 to 6) except in the case tively unsaturated Zn(II) Lewis acid site of ZnMOF-1-NH2,
of cyclohexene oxide (Table 2, entry 7). In the case of functio- which can activate the epoxy ring. In the second step, Br− gen-
nalized aliphatic propylene oxide, an increase in substituents/ erated from TBAB attacks the less hindered carbon atom of the
chain length has a decreasing effect in product yield (Table 2, coordinated epoxides, followed by a ring-opening step.
entries 2, 3, and 5). Subsequently, CO2 interacts with the oxygen anion of the
For aromatic epoxide, bulky substitution such as Br on the opened epoxy ring, forming an alkyl carbonate anion, which is
aromatic ring resulted in a significant decrease in the corres- converted into the corresponding cyclic carbonate through
Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

ponding cyclic carbonate product, which may be because the ring closure. The good catalytic performance of ZnMOF-1-NH2
steric factor delayed the approach of the substrate towards the is believed to be due to the synergistic effect of the catalytic
Lewis acidic metal nodes (Table 2, entries 1, 4 and 6). system in the presence of TBAB. The presence of Lewis basic
Compared to terminal epoxide, the internal epoxide 1,2-epoxy- sites such as the amine functionality in the ligand moiety can
cyclohexane does not undergo any significant cycloaddition also assist the activation of CO2 by the MOF, which is ben-
reaction (18% product yield), which can be attributed to the eficial for its catalytic performance.
sterically hindered cyclohexene moiety (Table 2, entry 6).105 In
all the catalytic reactions, the product analysis was performed
Catalytic oxidation of sulfides to sulfoxide
using both GC and NMR spectroscopy (Fig. S6–S17†).
Based on previous reports61,65,106,107 and the crystal struc- Considering the importance of sulfoxides as efficient synthetic
ture of ZnMOF-1-NH2, the probable mechanism for the cata- auxiliaries and valuable pharmaceuticals precursors, we
lytic reaction is proposed and illustrated in Fig. 7. The pro- explored the catalytic oxidation of aromatic sulfides using
ZnMOF-1-NH2 since it exhibits good thermal and chemical
stability as a heterogeneous catalyst. The oxidation of thioani-
sole (0.5 mmol) was selected as a model reaction to evaluate
the catalytic performance of ZnMOF-1-NH2 in the presence of
tert-butyl hydroperoxide (TBHP, 1.0 mmol) as an oxidant and
dichloromethane (DCM, 4.0 mL) as a solvent in a 10 mL glass
tube at 200 rpm to optimise the reaction conditions (Table 3).
Initially, the influence of oxidants (H2O2/TBHP/UHP) was
studied for the oxidation of thioanisole in the optimization
experiments. Among the three oxidants, TBHP proved to be
the best, which gave the product with satisfactory conversion
when 1 mmol was used. Subsequently oxidation reactions were
performed at 25 °C with varying reaction times of 1, 2 and 3 h
and the corresponding sulfide conversion was 77%, 91% and
99% with the selectivity of 78%, 88% and 95%, respectively
(Table 3, entries 5–7). Indeed, TBHP and H2O2 alone produced
23% and trace amount of product at 3 h and 12 h reaction
time, but the combination of catalyst and H2O2 at 25 °C and
Fig. 7 Proposed catalytic mechanism for the cycloaddition of CO2 to 40 °C offered 22% and 50% sulfide conversion upon 12 h of
epoxide using ZnMOF-1-NH2 as the catalyst and TBAB as a co-catalyst. reaction time (Table 3, entries 1–4), respectively. The

Table 3 Oxidation of thioanisole into methyl phenyl sulfoxidea

Entry Catalyst/oxidant Temp. (°C) Time (h) Sulfide conversion (%) Sulfoxide selectivity (%)

1 H2O2 25 12 Trace —
2 ZnMOF-1-NH2/H2O2 25 12 22 4
3 ZnMOF-1-NH2/H2O2 40 12 52 38
4 TBHP 25 3 23 72
5 ZnMOF-1-NH2/TBHP 25 1 77 78
6 ZnMOF-1-NH2/TBHP 25 2 91 88
7 ZnMOF-1-NH2/TBHP 25 3 99 95
8 ZnMOF-1-NH2/TBHP 40 1 99 93
9 ZnMOF-1-NH2/TBHP 40 2 99 86
10 ZnMOF-1-NH2/UHP 25 3 28 90
a
Reaction conditions: Thioanisole = 0.5 mmol (0.06 mL at 25 °C), catalyst mol%: ZnMOF-1-NH2 = 2.1 mol% (0.021 mmol, 0.010 g); H2O2/tert-
butyl hydroperoxide (TBHP)/UHP = 1.0 mmol, DCM = 4.0 mL, and 200 rpm.

8046 | Dalton Trans., 2018, 47, 8041–8051 This journal is © The Royal Society of Chemistry 2018
View Article Online

Dalton Transactions Paper

ZnMOF-1-NH2/UHP combination at 25 °C also furnished poor and selectivity under the optimized reaction conditions were
yield with 25% conversion of sulfide in 3 h reaction time obtained in the case of thioanisole, and an increase in chain
(Table 3, entry 10). length/substitution on the phenyl ring in the substrates
Reaction variables such as temperature and time were resulted in a decrease in product yield. In the case of ortho-
modulated using ZnMOF-1-NH2/TBHP and the sulfide conver- bromo thioanisole, an extension in the reaction time from 3 h
sion at 40 °C with 1 and 2 h reaction times displayed 99% con- to 6 h produced the desired sulfoxide product with lower con-
version; however, the selectivity was 93% and 86% (Table 3, version and selectivity (Table 5, entry 5). This may be attribu-
entries 8 and 9), respectively. The lower selectivity in entries 8 ted to the slow diffusion and steric hindrance sustained upon
and 9 may be due to the further oxidation of sulfoxide to substitution with bulky substrates, which hamper the easy
Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

sulfone under the extended settings. access to the catalytically active metal sites.
The effect of solvent on thioanisole conversion in the cata- In summary, it is reasonable to conclude that the selective
lytic reactions was also investigated using a variety of solvents oxidation of various sulfides was observed using ZnMOF-1-
including water, and the results are displayed in Table 4. With NH2 as an efficient heterogenous catalyst, where with an
DCM as the solvent medium, the product yield and selectivity increase in the size of the substrate, the conversion and selecti-
were maximum (99% conversion and 95% selectivity); thus, vity decreased slightly owing to the larger steric hindrance. In
DCM was chosen as the ideal solvent for further catalytic reac- all the catalytic reactions, the product analysis was performed
tions (Table 4, entry 1). Tracing the above experiments, the using both GC and NMR spectroscopy (Fig. S18–S27†). The
optimized reaction conditions are 25 °C with ZnMOF-1-NH2/ sulfoxide formation was further established unambiguously by
TBHP as the catalyst/oxidant combination in DCM with con- crystalizing the product obtained in the reaction (Table 5,
centration of the precursors as mentioned in Table 3 for the entry 3) and solving the crystal structure of phenyl methyl
further sulfoxidation reaction of different aromatic substrates benzyl sulfoxide (Fig. S28†). Furthermore, the plausible
(Table 5). As revealed in Table 5, the maximum product yield mechanism for sulfoxide formation is shown in Fig. 8. The
penta-coordinated Zn(II) dimeric cluster possessing a vacant
coordination site is accessible for coordination with the
Table 4 Oxidation of thioanisole into methyl phenyl sulfoxide in oxygen atom of TBHP for the formation of an oxygenated inter-
different solventsa mediate. The oxygenated intermediate undergoes nucleophilic
attack on the sulfide and the ensuing concerted oxygen trans-
Entry Solvent Conversion (%) Selectivity (%)
fer results in the formation of sulfoxide.91
1 DCM 99 95
2 DMSO 99 81 Chemical stability and catalytic recyclability of ZnMOF-1-NH2
3 Ethanol 83 95
4 Ethyl acetate 72 94 In the catalytic process, the chemical stability and recyclability
5 Diethyl ether 55 98 of a catalyst is highly important for its industrial application.
6 H2O 42 92
The catalytic performance remained high in both catalytic
a
Reaction conditions: Thioanisole = 0.5 mmol (0.06 mL at 25 °C), cata-
lyst mol%: ZnMOF-1-NH2 = 2.1 mol% (0.021 mmol, 0.010 g); tert-butyl
hydroperoxide (TBHP) = 1.0 mmol, solvent = 4.0 mL, and 200 rpm.

Table 5 Substrate screening using ZnMOF-1-NH2 catalyst in the oxi-


dation of sulfides into sulfoxidea

Entry Sulfide Time (h) Conversion (%) Yield (%)

1 3 99 95

2 3 95 92

3 3 89 83

4 3 90 88

5 6 73 71

a
Reaction conditions: Sulfide = 0.5 mmol, catalyst mol%: ZnMOF-1-
NH2 = 2.1 mol% (0.021 mmol, 0.010 g); tert-butyl hydroperoxide Fig. 8 Proposed catalytic mechanism for the sulfoxidation reaction
(TBHP) = 1.0 mmol, DCM = 4.0 mL, 200 rpm, 25 °C, and 3 h. using ZnMOF-1-NH2 and TBHP as an oxidant.

This journal is © The Royal Society of Chemistry 2018 Dalton Trans., 2018, 47, 8041–8051 | 8047
View Article Online

Paper Dalton Transactions

occurred (Fig. S5†). Therefore, the above evidence clearly


support that ZnMOF-1-NH2 is a robust and excellent multi-
functional heterogeneous catalyst for CO2 fixation and sulfoxi-
dation reactions, which can be synthesized via a green
mechanochemical method and exhibits good recyclability. The
good catalytic activity of ZnMOF-1-NH2 for the catalytic reac-
tions can be mainly attributed to the synergic effect of the
well-oriented high density of active metal sites/amino and
amide functional moieties of the ligand exposed for inter-
Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

actions with the substrates.

Conclusions
Judicious choice of catalytically active metal nodes and ligands
incorporating appropriate functional groups as linkers has a
great influence in the development of efficient and benign
catalytic systems. In summary, we synthesized and character-
ized an amine-functionalized three-dimensional metal organic
framework involving Zn(II) metal nodes and aromatic dicarbox-
ylate/N-donor Schiff base ligand as linkers via different routes.
The synthesized ZnMOF-1-NH2 possesses unsaturated coordi-
nation sites and functional groups in the ligand moiety, and
thus was exploited as a multifunctional heterogeneous catalyst
for the CO2-epoxide coupling reaction and oxidative desulfuri-
zation of sulfides with high conversion and selectivity. Its
Fig. 9 (a) Recycling of catalyst ZnMOF-1-NH2 up to 4 recycles for excellent chemical stability, easy separation and recyclability
cycloaddition and sulfoxidation reactions and (b) Comparison of the with no obvious decrease in catalytic activity demonstrate the
PXRD data of the as-synthesized ZnMOF-1-NH2 with the recovered potential this material as a heterogenous catalyst.
sample after 4 catalytic cycles of cycloaddition and sulfoxidation
reactions.

Conflicts of interest
There are no conflicts to declare.
reactions using ZnMOF-1-NH2 as the catalyst, and the recov-
ered catalyst was reused for four consecutive cycles without sig-
nificant loss in efficiency (Fig. 9a).
The catalyst was recovered after 4 cycles by centrifugation
Acknowledgements
and filtration followed by thorough washing and subjected to The registration number of this publication is CSIR-CSMCRI –
different physico-chemical analyses such as FTIR, PXRD and 061/2018. Financial support from DST Project HCP0009 (PP),
SEM to establish its chemical stability. As shown in Fig. 9b, CSIR for SRF (BP) and analytical support by AESD&CIF of
the PXRD patterns before and after the catalytic reaction are CSIR-CSMCRI is gratefully acknowledged. We thank Ms
identical and matches well with the simulated pattern from Ridhdhi Laiya for PXRD data, Mr Vinod Boricha for NMR data,
the SCXRD data, and the FTIR spectrum of the material after Mr Satyaveer Gothwal for TGA data, Mr Vinod Kumar Agrawal
the catalytic performance retains the characteristic peaks of for FTIR data, Mr Viral Vakani for CHN analysis, Mr Jayesh
the compound, indicating that its structural components are Chaudhari for FE-SEM analysis and Mr Harpalsinh Rathod for
stable (Fig. S2†). To further support this, the FE-SEM images N2/CO2-adsorption desorption analysis.
of the compound also retained the well-defined crystalline
morphology and texture of the material before and after the
catalytic reactions (Fig. 1). To estimate any leaching of the References
metal from the ZnMOF-1-NH2 catalyst during the reaction, the
solid catalyst was separated from both reaction mixtures 1 J. Artz, T. E. Müller, K. Thenert, J. Kleinekorte, R. Meys,
halfway through the reaction. The reaction was further contin- A. Sternberg, A. Bardow and W. Leitner, Chem. Rev., 2018,
ued without catalyst in the presence of only the filtrate. No 118, 434–504.
further increase in the product yield for both reactions was 2 T. Sakakura, J.-C. Choi and H. Yasuda, Chem. Rev., 2007,
observed, which clearly confirms that the catalytically active 107, 2365–2387.
sites reside on the catalyst and no leaching of the metal 3 R. Bentley, Chem. Soc. Rev., 2005, 34, 609–624.

8048 | Dalton Trans., 2018, 47, 8041–8051 This journal is © The Royal Society of Chemistry 2018
View Article Online

Dalton Transactions Paper

4 J. W. Comerford, I. D. V. Ingram, M. North and X. Wu, 30 F. D. Bobbink, W. Gruszka, M. Hulla, S. Das and
Green Chem., 2015, 17, 1966–1987. P. J. Dyson, Chem. Commun., 2016, 52, 10787–10790.
5 I. Omae, Coord. Chem. Rev., 2012, 256, 1384–1405. 31 H. Büttner, J. Steinbauer and T. Werner, ChemSusChem,
6 S. Huang, B. Yan, S. Wang and X. Ma, Chem. Soc. Rev., 2015, 8, 2655–2669.
2015, 44, 3079–3116. 32 B. Bousquet, A. Martinez and V. Dufaud, ChemCatChem,
7 J. Legros, J. R. Dehli and C. Bolm, Adv. Synth. Catal., 2005, 2018, 10, 843–848.
347, 19–31. 33 L. Wang, G. Zhang, K. Kodama and T. Hirose, Green
8 D. Zhang, J.-P. Dutasta, V. Dufaud, L. Guy and Chem., 2016, 18, 1229–1233.
A. Martinez, ACS Catal., 2017, 7, 7340–7345. 34 Z. Xue, X. Zhao, J. Wang and T. Mu, Chem. – Asian J.,
Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

9 I. Fernández and N. Khiar, Chem. Rev., 2003, 103, 3651– 2017, 12, 2271–2277.
3706. 35 J. Gaoa, W. Maa, L. Yuana, Y. Dai and C. Li, Appl. Catal.,
10 J. Yu, L.-H. Xie, J.-R. Li, Y. Ma, J. M. Seminario and A, 2013, 467, 187–195.
P. B. Balbuena, Chem. Rev., 2017, 117, 9674–9754. 36 R.-H. Wu, J. Wu, M.-X. Yu and L.-G. Zhu, RSC Adv., 2017,
11 M. Aresta, A. Dibenedetto and A. Angelini, Chem. Rev., 7, 44259–44264.
2014, 114, 1709–1742. 37 C. Ren, R. Fang, X. Yu and S. Wang, Tetrahedron Lett.,
12 P. Lanzafame, G. Centi and S. Perathoner, Chem. Soc. Rev., 2018, 59, 982–986.
2014, 43, 7562–7580. 38 D. Julião, A. C. Gomes, M. Pillinger, R. Valença,
13 D. W. Keith, Science, 2009, 325, 1654–1655. J. C. Ribeiro, I. S. Gonçalves and S. S. Balula, Appl. Catal.,
14 P. Markewitz, W. Kuckshinrichs, W. Leitner, J. Linssen, B, 2018, 230, 177–183.
P. Zapp, R. Bongartz, A. Schreiber and T. E. Müller, Energy 39 S. M. G. Pires, M. M. Q. Simões, I. C. M. S. Santos,
Environ. Sci., 2012, 5, 7281–7305. S. L. H. Rebelo, F. A. Almeida Paz, M. G. P. M. S. Neves
15 T. Ema, Y. Miyazaki, J. Shimonishi, C. Maeda and and J. A. S. Cavaleiro, Appl. Catal., B, 2014, 160–161, 80–
J. Hasegawa, J. Am. Chem. Soc., 2014, 136, 15270–15279. 88.
16 X. Wang, Y. Zhou, Z. Guo, G. Chen, J. Li, Y. Shi, Y. Liu and 40 C. Copÿret, M. Chabanas, R. P. Saint-Arroman and
J. Wang, Chem. Sci., 2015, 6, 6916–6924. J.-M. Basset, Angew. Chem., Int. Ed., 2003, 42, 156–181.
17 K. R. Roshan, B. M. Kim, A. C. Kathalikkattil, J. Tharun, 41 M. Taherimehr, J. P. C. C. Sertã, A. W. Kleij,
Y. S. Won and D. W. Park, Chem. Commun., 2014, 50, C. J. Whiteoak and P. P. Pescarmon, ChemSusChem, 2015,
13664–13667. 8, 1034–1042.
18 S. Subramanian, J. Park, J. Byun, Y. Jung and C. T. Yavuz, 42 S. Verma, R. I. Kureshy, T. Roy, M. Kumar, A. Das,
ACS Appl. Mater. Interfaces, 2018, 10, 9478–9484. N. H. Khan, S. H. R. Abdi and H. C. Bajaj, Catal.
19 M. Liu, K. Gao, L. Liang, J. Sun, L. Sheng and M. Arai, Commun., 2015, 61, 78–82.
Catal. Sci. Technol., 2016, 6, 6406–6416. 43 R. Ma, L.-N. He and Y.-B. Zhou, Green Chem., 2016, 18,
20 C.-S. Cao, Y. Shi, H. Xu and B. Zhao, Dalton Trans., 2018, 226–231.
47, 4545–4553. 44 C. Maeda, J. Shimonishi, R. Miyazaki, J. Hasegawa and
21 Y. Kumatabara, M. Okada and S. Shirakawa, ACS T. Ema, Chem. – Eur. J., 2016, 22, 6556–6563.
Sustainable Chem. Eng., 2017, 5, 7295–7301. 45 S. He, F. Wang, W.-L. Tong, S.-M. Yiu and M. C. W. Chan,
22 J. Bayardon, J. Holz, B. Schäffner, V. Andrushko, Chem. Commun., 2016, 52, 1017–1020.
S. Verevkin, A. Preetz and A. Börner, Angew. Chem., Int. 46 M. H. Kim, T. Song, U. R. Seo, J. E. Park, K. Cho,
Ed., 2007, 46, 5971–5974. S. M. Lee, H. J. Kim, Y.-J. Ko, Y. K. Chung and S. U. Son,
23 H. Liu, Z. Huang, Z. Han, K. Ding, H. Liu, C. Xia and J. Mater. Chem. A, 2017, 5, 23612–23619.
J. Chen, Green Chem., 2015, 17, 4281–4290. 47 J. A. Castro-Osma, K. J. Lamb and M. North, ACS Catal.,
24 P. Unnikrishnana and D. Srinivas, J. Mol. Catal. A: Chem., 2016, 6, 5012–5025.
2015, 398, 42–49. 48 T.-T. Liu, R. Xu, J.-D. Yi, J. Liang, X.-S. Wang, P.-C. Shi,
25 S. Fujita, B. M. Bhanage, H. Kanamaru and M. Arai, Y.-B. Huang and R. Cao, ChemCatChem, 2018, 10, 2036–
J. Mol. Catal. A: Chem., 2005, 230, 43–48. 2040.
26 S. H. Kim and S. H. Hong, ACS Catal., 2014, 4, 3630–3636. 49 H. C. Zhou, J. R. Long and O. M. Yaghi, Chem. Rev., 2012,
27 M. Cokoja, M. E. Wilhelm, M. H. Anthofer, 112, 673–674.
W. A. Herrmann and F. E. Kühn, ChemSusChem, 2015, 8, 50 B. Li, H.-M. Wen, Y. Cui, W. Zhou, G. Qian and B. Chen,
2436–2454. Adv. Mater., 2016, 28, 8819–8860.
28 J. Tharun, A. C. Kathalikkattil, R. Roshan, D.-H. Kang, 51 S. Qiu, M. Xue and G. Zhu, Chem. Soc. Rev., 2014, 43,
H.-C. Woo and D.-W. Park, Catal. Commun., 2014, 54, 31– 6116–6140.
34. 52 E. Barea, C. Montoro and J. A. R. Navarro, Chem. Soc. Rev.,
29 S. Arayachukiat, C. Kongtes, A. Barthel, 2014, 43, 5419–5430.
S. V. C. Vummaleti, A. Poater, S. Wannakao, L. Cavallo 53 L. E. Kreno, K. Leong, O. K. Farha, M. Allendorf, R. P. Van
and V. D’Elia, ACS Sustainable Chem. Eng., 2017, 5, 6392– Duyne and J. T. Hupp, Chem. Rev., 2012, 112, 1105–1125.
6397. 54 M. Kurmoo, Chem. Soc. Rev., 2009, 38, 1353–1379.

This journal is © The Royal Society of Chemistry 2018 Dalton Trans., 2018, 47, 8041–8051 | 8049
View Article Online

Paper Dalton Transactions

55 J. Liu, L. Chen, H. Cui, J. Zhang, L. Zhang and C.-Y. Su, S.-L. Pan, M.-S. Hung and S.-H. Ueng, J. Med. Chem., 2016,
Chem. Soc. Rev., 2014, 43, 6011–6061. 59, 419–430.
56 L. Zhu, X.-Q. Liu, H.-L. Jiang and L.-B. Sun, Chem. Rev., 76 C. F. Pereira, F. Figueira, R. F. Mendes, J. Rocha,
2017, 117, 8129–8176. J. T. Hupp, O. K. Farha, M. M. Q. Simões, J. P. C. Tomé
57 A. H. Chughtai, N. Ahmad, H. A. Younus, A. Laypkov and and F. A. Almeida Paz, Inorg. Chem., 2018, 57, 3855–3864.
F. Verpoort, Chem. Soc. Rev., 2015, 44, 6804–6849. 77 C. Yang, Q. Jin, H. Zhang, J. Liao, J. Zhu, B. Yu and
58 A. Dhakshinamoorthy, M. Alvaro and H. Garcia, Chem. J. Deng, Green Chem., 2009, 11, 1401–1405.
Commun., 2012, 48, 11275–11288. 78 J. M. Shin, Y. M. Cho and G. Sachs, J. Am. Chem. Soc.,
59 A. Dhakshinamoorthy, M. Opanasenko, J. Čejka and 2004, 126, 7800–7811.
Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

H. Garcia, Catal. Sci. Technol., 2013, 3, 2509– 79 P. K. Bera, D. Ghosh, S. H. R. Abdi, N. H. Khan,
2540. R. I. Kureshy and H. C. Bajaj, J. Mol. Catal. A: Chem., 2012,
60 D. Farrusseng, S. Aguado and C. Pinel, Angew. Chem., Int. 361–362, 36–44.
Ed., 2009, 48, 7502–7513. 80 G. Licini, M. Mba and C. Zonta, Dalton Trans., 2009,
61 R. Babu, A. C. Kathalikkattil, R. Roshan, J. Tharun, 5265–5277.
D.-W. Kim and D.-W. Park, Green Chem., 2016, 18, 232– 81 R. Frenzel, Á. G. Sathicq, M. N. Blanco, G. P. Romanelli
242. and L. R. Pizzio, J. Mol. Catal. A: Chem., 2015, 403, 27–36.
62 A. C. Kathalikkattil, D.-W. Kim, J. Tharun, H.-G. Soek, 82 J. Wang, Y. Niu, M. Zhang, P. Ma, C. Zhang, J. Niu and
R. Roshan and D.-W. Park, Green Chem., 2014, 16, 1607– J. Wang, Inorg. Chem., 2018, 57, 1796–1805.
1616. 83 J.-K. Li, J. Dong, C.-P. Wei, S. Yang, Y.-N. Chi, Y.-Q. Xu and
63 R. Babu, R. Roshan, A. C. Kathalikkattil, D. W. Kim and C.-W. Hu, Inorg. Chem., 2017, 56, 5748–5756.
D.-W. Park, ACS Appl. Mater. Interfaces, 2016, 8, 33723– 84 F. Jalilian, B. Yadollahi, M. Riahi Farsani,
33731. S. Tangestaninejad, H. A. Rudbari and R. Habibi, Catal.
64 A. C. Kathalikkattil, R. Roshan, J. Tharun, R. Babu, Commun., 2015, 66, 107–110.
G.-S. Jeong, D.-W. Kim, S. J. Cho and D.-W. Park, Chem. 85 H. An, Y. Hou, L. Wang, Y. Zhang, W. Yang and S. Chang,
Commun., 2016, 52, 280–283. Inorg. Chem., 2017, 56, 11619–11632.
65 A. C. Kathalikkattil, R. Roshan, J. Tharun, H.-G. Soek, 86 X. Zhao, Y. Duan, F. Yang, W. Wei, Y. Xu and C. Hu, Inorg.
H.-S. Ryu and D.-W. Park, ChemCatChem, 2014, 6, 284– Chem., 2017, 56, 14506–14512.
292. 87 H. Haddadi, S. M. Hafshejani and M. R. Farsani, Catal.
66 R. Babu, R. Roshan, Y. Gim, Y. H. Jang, J. F. Kurisingal, Lett., 2015, 145, 1984–1990.
D. W. Kim and D.-W. Park, J. Mater. Chem. A, 2017, 5, 88 B.-B. Lu, J. Yang, Y.-Y. Liu and J.-F. Ma, Inorg. Chem., 2017,
15961–15969. 56, 11710–11720.
67 A. C. Kathalikkattil, R. Babu, R. K. Roshan, H. Lee, 89 D. N. Dybtsev, A. L. Nuzhdin, H. Chun, K. P. Bryliakov,
H. Kim, J. Tharun, E. Suresh and D.-W. Park, J. Mater. E. P. Talsi, V. P. Fedin and K. Kim, Angew. Chem., Int. Ed.,
Chem. A, 2015, 3, 22636–22647. 2006, 45, 916–920.
68 J. F. Kurisingal, R. Babu, S.-H. Kim, Y. X. Li, J. S. Chang, 90 W. Xuan, C. Ye, M. Zhang, Z. Chen and Y. Cui, Chem. Sci.,
S.-J. Cho and D.-W. Park, Catal. Sci. Technol., 2018, 8, 591– 2013, 4, 3154–3159.
600. 91 K. Tanaka, K. Kubo, K. Iida, K. Otani, T. Murase,
69 H. He, J. A. Perman, G. Zhu and S. Ma, Small, 2016, 12, D. Yanamoto and M. Shiro, Asian J. Org. Chem., 2013, 2,
6309–6324. 1055–1060.
70 W.-Y. Gao, H. Wu, K. Leng, Y. Sun and S. Ma, Angew. 92 Z. Yang, C. Zhu, Z. Li, Y. Liu, G. Liu and Y. Cui, Chem.
Chem., Int. Ed., 2016, 55, 5472–5476. Commun., 2014, 50, 8775–8778.
71 X. Wang, W.-Y. Gao, Z. Niu, L. Wojtas, J. A. Perman, 93 Y.-B. Zhang, H. Furukawa, N. Ko, W. Nie, H. J. Park,
Y.-S. Chen, Z. Li, B. Aguila and S. Ma, Chem. Commun., S. Okajima, K. E. Cordova, H. Deng, J. Kim and
2018, 54, 1170–1173. O. M. Yaghi, J. Am. Chem. Soc., 2015, 137, 2641–2650.
72 H. He, Q. Sun, W. Gao, J. A. Perman, F. Sun, G. Zhu, 94 Z. Lu, Y. Xing, L. Du, H. He, J. Zhang and C. Hang, RSC
B. Aguila, K. Forrest, B. Space and S. Ma, Angew. Chem., Adv., 2017, 7, 47219–47224.
Int. Ed., 2018, 57, 4657–4662. 95 H. Liu, F.-G. Xi, W. Sun, N.-N. Yang and E.-Q. Gao, Inorg.
73 W.-Y. Gao, Y. Chen, Y. Niu, K. Williams, L. Cash, Chem., 2016, 55, 5753–5755.
P. J. Perez, L. Wojtas, J. Cai, Y.-S. Chen and S. Ma, Angew. 96 M. Taherimehr, B. Van de Voorde, L. H. Wee,
Chem., Int. Ed., 2014, 53, 2615–2619. J. A. Martens, D. E. De Vos and P. P. Pescarmon,
74 Z.-R. Jiang, H. Wang, Y. Hu, J. Lu and H.-L. Jiang, ChemSusChem, 2017, 10, 1283–1291.
ChemSusChem, 2015, 8, 878–885. 97 B. Parmar, Y. Rachuri, K. K. Bisht and E. Suresh, Inorg.
75 S.-Y. Lin, T.-K. Yeh, C.-C. Kuo, J.-S. Song, M.-F. Cheng, Chem., 2017, 56, 10939–10949.
F.-Y. Liao, M.-W. Chao, H.-L. Huang, Y.-L. Chen, 98 T. Friščić, J. Mater. Chem., 2010, 20, 7599–7605.
C.-Y. Yang, M.-H. Wu, C.-L. Hsieh, W. Hsiao, Y.-H. Peng, 99 M. Klimakow, P. Klobes, A. F. Thünemann, K. Rademann
J.-S. Wu, L.-M. Lin, M. Sun, Y.-S. Chao, C. Shih, S.-Y. Wu, and F. Emmerling, Chem. Mater., 2010, 22, 5216–5221.

8050 | Dalton Trans., 2018, 47, 8041–8051 This journal is © The Royal Society of Chemistry 2018
View Article Online

Dalton Transactions Paper

100 K. Roztocki, D. Jędrzejowski, M. Hodorowicz, 104 J. Song, Z. Zhang, S. Hu, T. Wu, T. Jiang and B. Han,
I. Senkovska, S. Kaskel and D. Matoga, Inorg. Chem., 2016, Green Chem., 2009, 11, 1031–1036.
55, 9663–9670. 105 A. Zhu, T. Jiang, B. Han, J. Zhang, Y. Xie and X. Ma, Green
101 R. W. Flaig, T. M. Osborn Popp, A. M. Fracaroli, Chem., 2007, 9, 169–172.
E. A. Kapustin, M. J. Kalmutzki, R. M. Altamimi, 106 R. Babu, S.-H. Kim, A. C. Kathalikkattil,
F. Fathieh, J. A. Reimer and O. M. Yaghi, J. Am. Chem. R. R. Kuruppathparambil, D. W. Kim, S. J. Cho and
Soc., 2017, 139, 12125–12128. D.-W. Park, Appl. Catal., A, 2017, 544,
102 J. Sun, S. Fujita, F. Zhao and M. Arai, Green Chem., 2004, 126–136.
6, 613–616. 107 P. Patel, B. Parmar, R. I. Kureshy, N. H. Khan and
Published on 21 May 2018. Downloaded by University of South Dakota on 6/21/2018 3:00:32 PM.

103 F. Jutz, J.-D. Grunwaldt and A. Baiker, J. Mol. Catal. A: E. Suresh, ChemCatChem, 2018, DOI: 10.1002/
Chem., 2008, 279, 94–103. cctc.201800137.

This journal is © The Royal Society of Chemistry 2018 Dalton Trans., 2018, 47, 8041–8051 | 8051

You might also like