You are on page 1of 13

Article

Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX pubs.acs.org/IC

Cycloaddition of CO2 with an Epoxide-Bearing Oxindole Scaffold by


a Metal−Organic Framework-Based Heterogeneous Catalyst under
Ambient Conditions
Bhavesh Parmar,†,‡,# Parth Patel,§,∥,# Renjith S. Pillai,⊥ Raj Kumar Tak,‡,§
Rukhsana I. Kureshy,‡,§,∥ Noor-ul H. Khan,*,‡,§,∥ and Eringathodi Suresh*,†,‡

Analytical and Environmental Science Division and Centralized Instrument Facility and §Inorganic Materials and Catalysis Division,
CSIR-Central Salt and Marine Chemicals Research Institute, G. B. Marg, Bhavnagar 364 002, Gujarat, India

Academy of Scientific and Innovative Research (AcSIR), Ghaziabad 201 002, India
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.


Charotar University of Science & Technology, Changa, 388 421 Anand, Gujarat, India

Downloaded via UNIV OF SOUTHERN INDIANA on July 19, 2019 at 13:00:02 (UTC).

Department of Chemistry, Faculty of Engineering and Technology, SRM Institute of Science and Technology, Kattankulathur 603
203, Chennai, India
*
S Supporting Information

ABSTRACT: The synthesis and characterization of a mixed


ligand metal−organic framework (MOF) with good thermal
and chemical stability, {[Co(BDC)(L)·2H 2 O]·xG} n
(CoMOF-2), involving an aromatic dicarboxylate (H2BDC
= 1,4-benzenedicarboxylic acid) and an acyl-decorated N-
donor linker [L = (E)-N′-(pyridin-4-ylmethylene) isonicoti-
nohydrazide] by various physicochemical techniques, includ-
ing Single crystal X-Ray Diffraction (SXRD), are reported.
The MOF showed a good affinity for CO2 capture, and Grand
Canonical Monte Carlo simulation studies exposed strong
interactions of CO2 with the functionalized N-donor ligand of
the framework. CoMOF-2 and KI act as an efficient binary
catalyst for the sustainable utilization of CO2 with spiro-epoxy
oxindole to spirocyclic carbonate under ambient conditions. Notably, herein we report MOF-based catalysis for the
cycloaddition of oxindole-based epoxides with CO2 for harvesting new spirocyclic carbonates. Interestingly, we could isolate and
crystallize six of the spirocyclic carbonate products, and the structure of the newly synthesized molecules has been established
by SXRD analysis. We present a plausible proposed catalytic mechanism through activation of the epoxide ring by the Lewis
acidic/basic sites present on the framework surface that is validated by molecular modeling.

■ INTRODUCTION
Interest in metal−organic frameworks (MOFs), with a porous
abundance, nontoxicity, and environmental issues. One of the
strategies for utilizing CO2 is cycloaddition with epoxides by
nature and good chemical and thermal stability, is rapidly developing suitable catalysts for producing value-added
expanding, and their potential is being exploited in various chemicals that have a wide range of applications.18−26
applications in the areas of gas storage/separation, sensing, MOFs have attracted interest as a potential heterogeneous
drug delivery, air purification, etc.1−8 MOFs are attractive catalyst for the CO2−epoxide cycloaddition reaction based on
catalytic materials because of their inherent topological and their large surface area, tunable pore size, and chemical
structural properties, which could be rationally designed by the stability.27−30 However, MOF-based sustainable catalysis for
sensible choice of the metal nodes and organic linkers.9−12 CO2 capture and efficient conversion in cycloaddition with
Nonetheless, the chemical and thermal stability of the pristine epoxides under ambient conditions into useful organic
MOF in different solvents, water, and acidic/basic media is a pharmaceutical intermediates are limited. The conventional
major concern in practical applications, including cataly- CO2−epoxide cycloaddition reaction mainly focused on
sis.13−15 A prudent choice of the functionalized ligands as aromatic/aliphatic epoxide substrates, and heterocyclic epox-
linkers and metal nodes with a versatile coordination geometry ides have been infrequently explored.31−46 Spirocyclic
can govern the porosity, surface area, and stability of the MOF compounds with a privileged heterocyclic oxindole motif are
material with respect to the catalytic application.16,17 Carbon
dioxide (CO2) has attracted more interest recently as a Received: April 26, 2019
renewable C1 building block for organic synthesis due to its

© XXXX American Chemical Society A DOI: 10.1021/acs.inorgchem.9b01234


Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

widely available in natural alkaloids and recognized as report.65 CHNS analyses were performed using an elementar vario
pharmacologically interesting compounds.47 For example, MICRO CUBE analyzer. Infrared (IR) spectra were recorded using
oxindole-fused tetrahydropyrane or chroman scaffolds des- the KBr pellet method on a Perkin-Elmer GX FTIR spectrometer. For
ignated as oxaspirooxindole are notable for their pharmaceut- each IR spectrum, 10 scans were recorded at 4 cm−1 resolution. 1H
and 13C nuclear magnetic resonance (NMR) spectra for the ligand,
ical activities.48−52 Spirocyclic alkaloids spirotryprostatins, epoxides, and cyclic carbonate derivatives were recorded on a JEOL
horsfiline and gelsemine, with an oxindole scaffold have JNM-ECZ 600R (600 MHz) spectrometer at 25 °C that was
attracted a significant amount of interest of medicinal chemists calibrated with respect to internal reference tetramethylsilane.
because of the presence of the spirotryprostatins in bioactive Thermogravimetric analysis (TGA) was carried out using a Mettler
natural products.53−56 Due to their biological activity and Toledo Star SW 8.10 instrument. TGA was performed in a nitrogen
physical properties, oxindole or spiro-oxindole core structures environment while the temperature was increased from room
represent an interesting synthetic challenge.57−59 The strategy temperature to 700 °C at a rate of 10 °C/min. Powder X-ray
for CO2−oxindole epoxide cycloaddition by developing a diffraction (PXRD) data were collected using a PANalytical
suitable catalyst is an innovative way to access spirocyclic Empyrean (PIXcel 3D detector) system with Cu Kα radiation.
Single-crystal structures were determined using a BRUKER SMART
oxindoles. So far, the only synthetic procedure for spirocyclic
APEX (CCD) or BRUKER D8 QUEST (PHOTON) diffractometer.
oxindole carbonate by CO2−epoxide cycloaddition is reported The N2 and CO2 adsorption−desorption isotherm and BET surface
on the basis of a deep eutectic solvent.60 Syntheses of organic area were measured on a Micromeritics 3 Flex instrument. Field
molecular architectures with the privileged spiro skeletons emission scanning electron microscopy (FE-SEM) micrographs were
represent a challenge, and these architectures have fascinated recorded using a JEOL JSM-7100F instrument employing a 15 kV
researchers as a class of biologically and pharmaceutically accelerating voltage.
important compounds.47 The spiro architecture is significantly Synthesis (diffusion method) of {[Co(BDC)(L)·2H2O]·xG}n
biologically important and may render the oxindole carbonate (CoMOF-2). H2BDC (benzene-1,4-dicarboxylic acid) (2.5 mmol)
as an interesting molecule. and NaOH (5 mmol) were dissolved in 20 mL to form a clear
solution. N donor linker L (1.25 mmol) was separately solubilized in
The methodology for the assembly of the oxindole core 20 mL of MeOH. A stock ligand solution was prepared by mixing the
fused with a cyclic carbonate by CO2 cycloaddition remains individual linker solution with constant stirring for 15 min and further
elusive using a MOF-based catalyst. We use the MOF- diluted to 70 mL with a MeOH/H2O (1:1) mixture. A stock ligand
catalyzed protocol for the synthesis of spirocyclic oxindole solution (3 mL) was layered above the 3 mL of the Co(ClO4)2·6H2O
carbonates facilitating cycloaddition of CO2−oxindole epox- (1 mmol in 15 mL of H2O) solution with 7−8 mL of the buffer
ides under ambient conditions, significantly improving the solution (1:1 MeOH:H2O) between them. The test tube was sealed
transformation, carbon footprint, and sustainability. with parafilm and allowed to stand at room temperature for slow
As a continuation of our ongoing research in the area of CO2 diffusion for crystallization. X-ray diffraction quality crystals (red in
utilization,61−64 a microporous three-dimensional (3D) MOF- color) appeared within 1 week. Yield: ∼79%. Elemental analysis (%).
based catalytic material, {[Co(BDC)(L)]·2H 2 O·xG} n Calcd for {[Co(BDC)(L)·2H2O]·xG}n: C, 45.58; H, 4.74; N, 10.12.
Found: C, 45.54; H, 4.66; N, 10.07. IR (cm−1, KBr): 3385 (w), 3146
(CoMOF-2; G = guest), involving benzene-1,4-dicarboxylic (br), 2899 (w), 2842 (w), 1680 (m), 1601 (s), 1554 (s), 1406 (s),
acid (H2BDC) and N-donor Schiff base ligand L [(E)-N′- 1286 (m), 1139 (w), 1068 (w), 1016 (w), 951 (w), 826 (w), 755 (w),
(pyridin-4-ylmethylene) isonicotinohydrazide] with accessible 696 (m), 543 (w).
porosity and densely populated metal sites has been Bulk Synthesis of CoMOF-2. The bulk powder material of
synthesized by two different approaches and characterized CoMOF-2 was prepared by stirring a mixture of Co(NO3)2·6H2O
well by physicochemical methods. Crystals of CoMOF-2 have (10 mmol), H2BDC (10 mmol), NaOH (20 mmol), and L (10
been harvested by diffusion, and the phase pure bulk material mmol) in 100 mL of a mixed solvent [5:4:1 (v:v)
has been synthesized by room temperature (RT) stirring of the H2O:MeOH:EtOH] for 8 h at room temeperature in a 200 mL
constituent ligands with a metal salt. CoMOF-2 possessing RBF. The resulting precipitate was filtered and washed with a MeOH/
H2O mixture [1:1 (v:v)] and acetone and then air-dried. Yield: ∼92%.
functionally decorated amide group acts as an efficient binary Elemental analysis (%). Calcd for CoMOF-2: C, 45.58; H, 4.74; N,
catalyst for CO2 sequestration/utilization of an oxindole-based 10.12. Found: C, 45.15; H, 4.83; N, 10.01. FTIR (cm−1, KBr): 3382
epoxide into a useful oxindole carbonate with mild reaction (w), 3162 (br), 2899 (w), 2851 (w), 1686 (w), 1625 (m), 1580 (s),
parameters. The efficient catalytic activity of CoMOF-2 in 1402 (s), 1290 (m), 1145 (w), 1065 (w), 1017 (w), 946 (w), 833
cycloaddition is governed by active Lewis acidic metal sites, the (w), 750 (w), 694 (m), 533 (w).
basic amide functional group on L, and the porosity. The Synthetic Procedure for Spiro-Epoxy Oxindole Derivatives.
chemical stability in the solid state or solution, acidic/basic Spiro-epoxy oxindole derivatives (S1−S8) were synthesized according
medium, and the features mentioned above for CoMOF-2 to a previous report.66,67 Trimethylsulfoxonium iodide (1.0 mmol)
support the excellent catalytic efficiency of this MOF-based and CsCO3 (2.0 mmol) were dissolved in dry CH3CN at 50 °C for 1
material as a heterogeneous catalyst in oxindole epoxide−CO2 h under a N2 atmosphere. The N-substituted isatin derivative (1.0
mmol) was dissolved in dry CH3CN (5 mL) and added dropwise over
cycloaddition with good yield and recyclability. Preferable a 10 min period to the solution described above. Reaction progress
adsorption sites for the CO2 with CoMOF-2 determined by and completion were monitored by thin layer chromatography. The
Grand Canonical Monte Carlo (GCMC) molecular simulation reaction crude was filtered through a Celite bed and evaporated to
studies and the possible reaction mechanism for the catalytic dryness. Column chromatography [silica gel (60−120 mesh), EtOAc/
activity have been corroborated by computational methods. hexane mixture (20:80) as the eluent] was used to purify the final


product, and it was identified by 1H and 13C NMR spectroscopy.
Crystal structures of S1, S2, S4, S5, and S8 have also been elucidated
EXPERIMENTAL SECTION using Single crystal X-Ray Diffraction (SXRD).
Materials and General Methods. All reagents and solvents Synthetic Procedure for Spiro-Epoxy Oxindole−CO2 Cyclo-
purchased from commercial sources were used without further addition. Cycloaddition of oxindole epoxide substrates and CO2 was
purification, and synthetic manipulations were performed using carried out in a glass tube (5 mL capacity) with constant stirring. For
distilled water. The N donor (E)-N′-(pyridin-4-ylmethylene) each reaction, spiro-epoxy oxindole (0.15 mmol), CoMOF-2 (5.0 mol
isonicotinohydrazide (L) was synthesized according to our previous %), and co-catalyst KI (or TBAI or TBAB; 0.2 mol %), wherever

B DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

Figure 1. (a) PXRD data of CoMOF-2 synthesized by diffusion and RT stirring with simulated SXRD. (b) PXRD data of CoMOF-2 dispersed for
1 week in various organic solvents. (c) PXRD data of CoMOF-2 dispersed in an aqueous acidic/basic solution (0.01 M for 12h). (d) PXRD data of
CoMOF-2 dispersed in boiling water at different temperatures for 24 h.

applicable, were introduced into the glass tube with 1 mL of CH3CN good agreement, endorsing the phase purity of the material
as the solvent. The glass tube was then pressurized with 1 atm of CO2 (Figure 1a).
at RT. The reaction mixture was heated to the desired temperature Symmetric and antisymmetric νCO bands of BDC appeared
(RT or 40 °C) while being stirred (600 rpm). Upon completion of
the reaction, cycloaddition was stopped and the mixture cooled to
at ∼1406 and ∼1554 cm−1, respectively, with a ΔνCO of
RT. The reaction mixture was then centrifuged to separate the catalyst ∼150 cm−1 indicating chelating/bridging carboxylate coordi-
and washed with ethyl acetate twice, and the products were isolated nation in CoMOF-2. A broad band appeared at 3400−3150
(in ethyl acetate) and identified by gas chromatography using cm−1 in the IR data, indicating vO−H and vN−H vibrations from
dodecane as an internal standard [Bruker 450-GC instrument water molecules and L (Figure S1). TGA data of CoMOF-2
equipped with an HP-5 capillary column (30 m × 0.25 mm), using disclosed good thermal stability, preserving the crystallinity up
a flame ionization detector]. The cyclic carbonate products were also to ∼330 °C, and the degradation of the organic moiety
identified by 1H and 13C NMR spectroscopy. Crystal structures of
P1−P4, P6, and P8 have also been elucidated using SXRD.
originates after ∼330 °C. TGA of the as-synthesized sample
Catalyst Recyclability. Catalyst CoMOF-2 was recovered by revealed a weight loss before 120 °C, corresponding to the loss
centrifugation after each catalytic reaction, washed with a H2O.MeOH of lattice water as well as methanol guest molecules, and
mixture [1:1 (v:v)] and acetone, air-dried, and used for the next structural decomposition commences around 330 °C (Figure
catalytic reaction under similar conditions for up to six cycles. During S2). The chemical stability of CoMOF-2 in polar/nonpolar
the recyclability study, fresh KI was added during each experiment. organic solvents was investigated by soaking the material in the
The chemical stability of the recovered catalyst was analyzed by different solvents for 1 week. PXRD data of the recovered
PXRD, FTIR, and FE-SEM. For recycled CoMOF-2. FTIR (cm−1,
KBr): 3375 (w), 3157 (br), 2899 (w), 2850 (w), 1687 (w), 1624 (m), material from dimethylformamide and dimethyl sulfoxide after
1581 (s), 1403 (s), 1290 (s), 1144 (w), 1065 (w), 1017 (w), 946 (w), 7 days showed a slight variation from the pristine sample,
833 (w), 750 (w), 698 (m), 538 (w). indicating that the crystallinity and chemical stability

■ RESULTS AND DISCUSSION


Characterization and Chemical Stability Study of
deteriorate with time in these solvents. However, the chemical
stability of the CoMOF-2 recovered from the rest of the
solvents is validated by PXRD data that are in agreement with
CoMOF-2. CoMOF-2 synthesized by different routes those of the pristine sample as shown in Figure 1b.
(diffusion and RT stirring) has been characterized with respect Furthermore, the stability of CoMOF-2 was systematically
to its thermal and chemical stability and phase purity by assessed in acidic/basic media by immersion in respective
different physicochemical methods such as FTIR, PXRD, aqueous solutions (0.01 M for 12 h) at RT. Recovered material
TGA, and FE-SEM. The calculated and measured PXRD data tested by PXRD and FTIR revealed a good match with the
of the synthesized MOF formed by flexible routes showed pristine compound representing excellent stability and
C DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

Figure 2. (a) [M2(BDC)2]n 2D sql sheets. (b and c) Double pillaring of 2D sql sheets by L and 2-fold interpenetrated 3D framework in CoMOF-2.
(d) Simplified pcu network topology in CoMOF-2.

retaining the crystallinity and/or structural integrity of the link the dimeric units by two BDC moieties and the third one
material in acidic/alkaline media (Figure 1c and Figure S3). yields the chelating mode of coordination.
The water stability of CoMOF-2 was evaluated by soaking the The [M2(COO)2] secondary building unit (SBU) of the
material in water or boiling water at different temperatures and dimeric cluster is governed by Co2+ nodes linked by syn−syn
verified by PXRD and FTIR indicating excellent water stability coordination from symmetric BDC moieties, which is further
(Figure 1d and Figure S3). FE-SEM analysis of synthesized coupled by the chelated carboxylate oxygen from different
CoMOF-2 shows a block-shaped morphology for the material BDC ligands creating [M2(BDC)2]n 2D sql sheets (Figure 2a).
obtained by the diffusion method and an agglomerated Each symmetrically disposed BDC ligand forms a μ3-η1η1η1η1-
spherical morphology for the bulk material synthesized by chelated bridging mode of coordination with the Co(II),
RT stirring and the recovered bulk material after six catalytic generating 2D {Co2(BDC)2}n sheets, and the Co(II)···Co(II)
cycles (Figure S4). The concomitant effect of the coordina- distance across the dimeric unit is 4.11 Å. The edge lengths
tively saturated Co2 clusters and effective coordination between the centers of Co(II) pairs of the square grid window
between the metal centers and the ligand moiety probably are 12.69 Å, and the diagonal distance of the {Co2(BDC)2}n
4,4′-net is 17.10 Å. The 2D {Co2(BDC)2}n sheets have an
impart excellent thermal and chemical stability on CoMOF-2.
offset ABAB mode of orientation along the bc plane in which
Crystal Structure of CoMOF-2. SXRD data revealed that
discrete alternate AAAA as well as BBBB sheets are doubly
CoMOF-2 crystallizes in monoclinic space group C2/c and
pillared via axial coordination by the terminal nitrogen of L
features a 3D framework composed of two-dimensional generating an interpenetrated 3D framework (Figure 2b,c).
[M2(BDC)2]n sheets based on dimeric Co(II) carboxylate The 3D framework can be simplified to 6-connected uninodal
clusters, pillared by L. Half of the dimer building block nets with pcu topology having a point symbol of {412.63}
[Co2(BDC)2(L)2] constitutes the asymmetric unit holding (Figure 2d). Hydrogen bonding interaction in CoMOF-2
center of symmetry, which comprises one molecule each of indicates that the encapsulated water molecules are involved in
completely deprotonated BDC and L, crystallographically strong O−H···O interaction in bridging the interpenetrated 3D
independent Co2+, and two water molecules as the solvent of framework (Figure 3 and Table S7).
crystallization. Gas Adsorption Experimental and Computational
A distorted octahedral N2O4 geometry around Co(II) is Study. Activated CoMOF-2 (CoMOF-2′) was prepared via
provided by four carboxylate oxygens from three BDC ligands exchange of the reaction solvent with DCM for 24 h, followed
creating a distorted square base and pyridyl nitrogen from two by heating at 120 °C for 12 h in an oven. The permanent
L moieties that completes the axial coordination. Among three porosity and pore properties of the activated sample of
different BDC ligands coordinated to each Co(II), the CoMOF-2′ have been elucidated by gas adsorption isotherms.
carboxylate oxygens are involved in the bridging mode to N2 (77 K) sorption measurements showed negligible uptake,
D DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

Figure 3. (a) Close-up of H-bonding interaction between two


different nets in CoMOF-2. (b) H-Bonding interaction between two
interpenetrated networks in CoMOF-2. Figure 4. (a) CO2 adsorption isotherm for CoMOF-2′ measured at
273 and 298 K. (b) Local views of the snapshots extracted from the
GCMC simulations for CO2 in CoMOF-2 at 1 bar and 298 K.
and the observed BET surface area for CoMOF-2′ was 6.8 m2/
g (Figure S5). The small surface area of CoMOF-2 can be (GCMC) simulations fairly reproduce the shape of CO2
attributed to the interpenetration of the 3D framework as adsorption isotherms at 273 and 298 K (Figure S7). To
depicted in Figure 2c, which in turn can also decrease the elucidate the isosteric heat (Qst) for CO2 adsorption in
effective porosity. However, adsorption analysis of CO2 CoMOF-2′, the Clausius−Clapeyron equation was used from
revealed promising results, due to the smaller kinetic diameter their corresponding experimental isotherm at 273 and 298 K.
of CO2 (3.30 Å) compared to that of N2 (3.64 Å) (Figure 4a). The isosteric heat (Qst) for CO2 reaches ∼35.0 kJ mol−1,
Pillared linker L from the framework with a polar amide indicating amide functional groups of L may be involved in
functional group can interact with CO2 having a high strong interactions with the adsorbed CO2 molecules (Figure
quadrupole moment. Preferential adsorption of CO2 over N2 S8). In addition, the microscopic mechanism revealed that
can be influenced by the pore size, polarity, or kinetic issue of CO2 molecules are interacting with the carbonyl group of L,
the 3D framework. with mean characteristic O(CO2)−Ocarbonyl distances in the
CO2 adsorption measurements at 273 and 298 K reached range of 2.8−3.5 Å (Figure S9) backing the adsorption of CO2
∼51 cm3/g (2.26 mmol/g) and ∼45 cm3/g (2.04 mmol/g), inside the pore (Figure 4b and Figure S10).
respectively, at 1 atm, which clearly authenticates the Catalytic Performance of CoMOF-2 in Cycloaddition
adsorption capacity of CoMOF-2′ (Figure 4a). Almost of Spiro-Epoxy Oxindoles with CO2. The good chemical
identical CO2 adsorption at 273 and 298 K, despite thermal and thermal stability, strong CO2 adsorption, functionally
agitation at higher temperatures, may be due to the ideal pore decorated L, and the presence of high-density metal sites are
size to fit CO2 and its supramolecular interaction with the features required for an excellent catalytic material for CO2−
framework. The theoretically predicted small accessible surface epoxide cycloaddition. By virtue of these attributes and the
area revealed CoMOF-2′ excludes the N2 due to its kinetic presence of Co(II) with a weakly chelated carboxylate oxygen,
diameter being larger than the pore size, which is further the amide functionality of L in CoMOF-2 can act as a Lewis
endorsed by the experimental surface area (6.8 m2/g) obtained acid/basic site in activating the epoxide ring. Unlike a
from the N2 isotherm at 77 K. The adsorption by CoMOF-2′ conventional aromatic/aliphatic epoxide, we have performed
at 1 atm (2.26 and 2.04 mmol/g at 273 and 298 K, the cycloaddition of CO2 with a series of heterocyclic epoxides
respectively) is comparable to those of functionally decorated (Scheme 1). A series of spiro-epoxy oxindole substrates have
acylamide/amide MOFs.68−71 Grand Canonical Monte Carlo been synthesized by a reported procedure using trimethylsul-
E DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

Scheme 1. Cycloaddition of Spiro-Epoxy Oxindole no promising results (Table 1, entries 1−3 and 11−13). For
Derivatives and Carbon Dioxide the best choice of a suitable co-catalyst, reactions were

Table 1. Cycloaddition of 1-Benzylspiro[indoline-3,2′-


oxiran]-2-one (S1) and CO2 for 1-Benzylspiro[indoline-
3,4′-[1,3]dioxolane]-2,2′-dione (P1) Formationa
entry catalyst/co-catalyst temp (°C) time (h) S1 conversion (%)b
1 none RT 24 1
2 CoMOF-2 RT 24 8
foxonium iodide and cesium carbonate at 50 °C in acetonitrile 3 KI RT 24 14
with the respective substituted isatin precursor in quantitative 4 CoMOF-2/TBAB RT 24 52
yield. All of the synthesized spiro-epoxy oxindoles were 5 CoMOF-2/TBAI RT 24 46
characterized by 1H and 13C NMR, and the crystal structures 6 CoMOF-2/KI RT 24 82
of five epoxide derivatives have been established unambigu- 7 CoMOF-2/KI RT 48 85
ously (Figures S11 and S14−S29). Oxindole−epoxide has 8 CoMOF-2/KI 40 24 99
been studied for ring-opening reaction using various 9 CoMOF-2/KI 40 8 99
nucleophiles, and reports of ring expansion are rare.72,73 Ring 10 CoMOF-2/KI 40 6 93
expansion products of oxindole−epoxides by catalytic cyclo- 11 none 40 8 6
addition of CO2 can govern a series of innovative molecules 12 CoMOF-2 40 8 13
that have been infrequently explored, and the spirocyclic 13 KI 40 8 23
oxindole product itself or their derivatives may find significance 14c CoMOF-2D/KI 40 8 97
as intermediates in medicinal chemistry or in the pharmaceut- 15 CoMOF-2/KCl 40 8 82
ical industry. A series of eight spiro-epoxy oxindole substrates a
Reaction conditions: 0.15 mmol of 1S, 600 rpm. Catalyst mole
(S1−S8) with different substitutions have been synthesized percents: 5.0 mol % CoMOF-2 and 0.2 mol % TBAB, TBAI, or KI.
and subjected to cycloaddition with CO2 under a pressure of 1 CH3CN as the solvent. PCO2 = 1 atm. bSelectivity of 99%. cCoMOF-
atm in acetonitrile using CoMOF-2 and a co-catalyst. All 2D indicates the compound was synthesized by the diffusion method.
spirocyclic products were extracted and characterized by gas
chromatography and 1H and 13C NMR spectroscopy, and the
crystal structures of six new spirocyclic products have been performed in the presence of TBAB, TBAI, and KI by retaining
established (Figure 5 and Figures S30−S45). Initially, the initially set up reaction parameters with S1 and co-catalyst
experiments with different parameters were conducted at 1 concentrations 0.15 mmol and 0.2 mol %.
atm of CO2 at RT to optimize the reaction conditions using 1- Excellent results under ambient conditions with 82% S1
benzylspiro[indoline-3,2′-oxiran]-2-one (S1) as the model conversion and 99% product selectivity were achieved for
substrate. A blank reaction and only MOF/KI with the CoMOF-2/KI in a 24 h reaction time, and an only marginal
model substrate revealed minor S1 conversion and upon increase is noticed with an increase in the reaction time to 48 h
variation of the reaction time and/or temperature also yielded (Table 1, entries 6 and 7). With the same reaction parameters,

Figure 5. Thermal ellipsoid plot depicting the crystal structure of spirocyclic carbonates (P1−P4, P6, and P8; 50% probability factor for the
thermal ellipsoids).

F DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

Table 2. Substrate Screening Using Catalyst CoMOF-2 in the Cycloaddition of Spiro-Epoxy Oxindole CO2a

a
Reaction conditions: 0.15 mmol of epoxide, 5.0 mol % CoMOF-2/0.2 mol % KI as the catalyst, CH3CN as the solvent, PCO2 = 1 atm, 8 h, 40 °C,
600 rpm.

the MOF−co-catalyst (TBAB and TBAI) combinations could the best solvent system for catalytic reaction was acetonitrile
attain only 52% and 46% S1 conversions, respectively (Table 1, with 99% S1 conversion and selectivity. On the basis of
entries 4 and 5, respectively), indicating KI is a more preliminary systematic screening, the CoMOF-2/KI combina-
competent co-catalyst. The initial experiment for screening tion showed optimized parameters (catalyst/co-catalyst and S1
the co-catalyst clearly demonstrates the reactivity of KI > concentrations of 5.0/0.2 mol % and 0.15 mmol at PCO2 = 1
TBAB > TBAI (Table 1, entries 4−6). Due to the bulkiness of atm, 40 °C, 8 h, in acetonitrile) for the best catalytic
TBAB and TBAI, the diffusion of these co-catalysts toward the
performance and was used for further catalytic studies. The
active sites is slower than that of KI, making it the best co-
catalytic activity of the material synthesized by the diffusion
catalyst in this case. Upon loading KCl as a co-catalyst under
method under optimized reaction conditions was also
the optimized reaction condition, the epoxide conversion was
82%, comparatively lower than that with KI as a co-catalyst performed to evaluate and compare with the bulk catalytic
(Table 1, entry 15). Thus, S1 conversion was only 85% at RT material, which showed almost identical results [97% substrate
in 48 h, but when the temperature was increased to 40 °C, 99% conversion (Table 1, entry 14)].
conversion and selectivity were achieved within 24 h (Table 1, We further probed systematically the set of optimal
entry 8). If the temperature is held at 40 °C, S1 conversions at parameters for cycloaddition of CO2−epoxyindoles (S1−S8)
8 and 6 h were 99% and 93%, respectively (Table 1, entries 9 using binary catalyst CoMOF-2/KI (Table 2). As shown in
and 10, respectively) and the best solvent system for catalytic Table 2, all substrates undergo catalytic conversion to the
reaction was acetonitrile on the basis of S1 conversion and corresponding oxindole spirocyclic carbonates in good yields
selectivity. Reaction parameters for time and temperature were (ranging from 75% to 99%).
monitored systematically for CoMOF-2/KI, and the corre- In the case of N-benzyl spiro-epoxy indole derivatives (S1−
sponding plots are provided in Figure S12. Catalytic S5), substitution on the aromatic ring (F and CH3) has an only
optimization conducted in various solvents (toluene, methanol, marginal effect on product yield (Table 2, entries 1−3).
dichloromethane, acetonitrile, and acetone) is provided in However, substitution with Cl on the aromatic ring decreased
Table S5. The reactions in solvents proceed with good S1 the product yield by 24%, which can be attributed to the bulky
conversions ranging from 76% to 99% with different product Cl group hindering the accessibility of the reactant and product
selectivities. On the basis of solvent optimization experiments, toward the active sites by slow diffusion, occurring on the
G DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

Figure 6. Proposed mechanism for CoMOF-2/KI-catalyzed cycloaddition of CO2−spiro-epoxy oxindoles.

surface of the MOF. For the second series (S6−S8), methyl/ basic catalytic site (−NH− functional group) residing in the
allyl substitution on the aromatic/indole nitrogen showed an ligand moiety.17 Ligands decorated by Lewis basic sites such as
only marginal decrease (10% and 8%) for the product an amino/acylamide moiety acting as an efficient MOF-based
compared to the pristine N-methyl spiro-epoxy oxindole catalyst for CO2−epoxide cycloaddition are also well
(Table 2, entries 6−8), and the flexibility of the allyl group documented in the literature.7,74−76 The basic site favors
in S8 probably supports the marginal change in the product CO2 capture in the narrow pores as mentioned above, which is
yield. supported by CO2 adsorption studies and further supple-
Catalytic Mechanism, Density Functional Theory mented by molecular modeling experiments in enhancing the
(DFT) Calculation, and Recyclability Studies. On the product yield by CO2 conversion.
basis of the literature33,64,74,75 and SXRD analysis, a tentative The plausible mechanism for CoMOF-2/KI-catalyzed
reaction mechanism for the catalysis as depicted in Figure 6 cycloaddition of CO2−spiro-epoxy oxindoles is reinforced by
has been proposed. Due to the bulkiness of the oxindole the DFT calculations (Supporting Information). The energy
substrate and the reduced effective porosity of the inter- profile and mechanistic pathway disclose the synergistic
penetrated framework, the catalytic reaction probably occurs participation of both the Co(II) metal site and L with the
on the surface of the MOF. The densely populated metal acyl hydrazine moiety interacting with the epoxides via the
clusters with a weakly chelated carboxylate group can open and lone pair of oxygen facilitating the ring opening as depicted in
act as a Lewis acidic site, and a Lewis basic site on ligand L Figure 7. In the catalytic reaction, activation and epoxide ring
(−NH of the amide group) can synergistically activate the opening occur upon coordination of the oxygen from the
epoxide ring (Figure 6). Ma et al. exemplified CO2−epoxide epoxide with a Lewis acidic metal. Simultaneously, nucleophilic
cycloaddition by MOF-based catalysts bearing both Lewis attack on the less hindered carbon of epoxides by the iodide
acidic and basic sites in the activation of epoxide by a Lewis ion (I−) from KI takes place, and the surrounding CO2
H DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

Figure 7. Energy profile diagram for different intermediates and transition states involved in the cycloaddition of spiro-epoxy oxindole and CO2
using the cluster model structure of the CoMOF-2 catalyst with the I− ion.

involved in the reaction generating the alkyl carbonate anion iodide ion on the α carbon [αC(epoxide)] of S1 generates an
upon reaction of CO2 with the opened epoxy ring and alkoxide species (IC) with a relative energy (50.9 kcal/mol)
subsequent ring closure step with regeneration of the catalyst leading to the first transition state (TS-1). The stronger Co−
engenders the resultant cyclic carbonate. It is worth O(C6) interaction with an interatomic distance of 2.34 Å is
mentioning here the mechanism in the cycloaddition of achieved by ring opening of αC(epoxide) followed by the
spiro-epoxy oxindoles with CO2 using DES as an efficient formation of a new αC(epoxide)−I bond with a distance of 2.57
process reported by Khan et al.60 The hydrogen bond donor Å, and the observed relative energy difference between TS-1
(HBD) and hydrogen bond acceptor (HBA) as a combination and IC is 17.0 kcal/mol.
in a 1:2 molar ratio involving urea and choline chloride as a The rate-determining step is identified as the high-energy
deep eutectic solvent have shown an excellent efficiency. The barrier ring opening in CO2 fixation with S1, which is on par
ring opening of epoxide is favored by the synergistic effect of with the previous DFT energy profile. The alkoxide species
DES by polarization of the C−O bond of epoxide via H- with iodide substitution (TS-1) adopts an intermediate state
bonding followed by nucleophilic attack by Cl. Successively, (Int-2) in the later stage with a total energy value of 30.72
the ring-opened intermediate reacts with CO2 and forms the kcal/mol. As indicated in Int-2, the approach of the CO2
catalytic product regenerating the DES. For CoMOF-2, both molecule at a distance of 2.58 Å and subsequent attack with S1
the Lewis acidic metal site and the basic site from L (acyl having a negative charge on oxygen resulted in the formation
amide functionalized) support activation, opening of the of a transition state (TS-2). In the TS-2 step, physical binding
epoxide ring, and CO2 addition.
of CO2 between Co and alkoxide takes place via the formation
The energy profile diagram and the experimental details of
the catalytic reaction mechanism determined by computational of new Co−O(CO2) and O(CO2)−αC(epoxide) bonds having
methods are provided in the Supporting Information. As distances of 3.05 and 2.08 Å, respectively, with a relative
depicted in Figure 7, the interaction of a cobalt metal center energy of 84.5 kcal/mol. In addition, ring closure leads to the
with the oxygen of the 1-benzylspiro[indoline-3,2′-oxiran]-2- formation of a new intermediate (Int-3) consuming 26.05
one (S1) results in an intermediate (Int-1), which is kcal/mol. The realization of a new O(epoxide)−C(CO2) bond
considered as the initial step in the catalytic pathway, and its leading to the formation of the product 1-benzylspiro[indoline-
energy is very similar to the sum of the energies of cluster 3,4′-[1,3]dioxolane]-2,2′-dione (P1) happens in step Int-3 by
model of CoMOF-2 (catalyst), a nucleophile (I−), CO2, and the interaction of the carbocation of CO2 with an oxygen atom
S1 (isolated reactants). At the initial stage, the distance of S1. Interestingly, a decrease in the O(CO2)−αC(epoxide) bond
between the α carbon [αC(epoxide)] of S1 and the I− ion was
>5.5 Å. The approach of the I− ion toward S1 (Int-1) distances is observed (from 1.98 to 1.42 Å), whereas Co−
decreases the relative energy (−2.09 kcal/mol) and the O(CO2) and C−I distances increase from 3.04 to 4.41 Å and
distance between the cobalt and O atom from S1 (4.06 Å) from 3.14 to 4.07 Å, respectively. The product (P1)
as well as that between the α carbon [αC(epoxide)] of S1 and the detachment followed by regeneration of the catalyst occurs
I− ion (4.0 Å). During ring opening, nucleophilic attack by the with a relative energy of −3.1 kcal/mol in the final step (FC).
I DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

The isolation of the product, separation, and the recyclability chemical fixation of CO2 to yield pharmaceutically important
of the catalyst are important protocols targeting energy organic spirocyclic oxindoles under ambient conditions. The
minimization and environmental issues. The catalytic activity mechanism of catalytic reaction is proposed and supported by
was sustained for up to six repeated cycles, and the chemical DFT calculations. The findings also provide insight for the
stability and structural integrity of the recycled catalyst were synthesis of robust new MOFs utilizing a tailor-made
intact (Figure 8a). The PXRD and FTIR spectra of recycled functionalized ligand with improved catalytic properties in
cycloaddition with sustainable utilization of CO2 and
heterocyclic epoxides producing value-added pharmaceutical
intermediates.


*
ASSOCIATED CONTENT
S Supporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.inorg-
chem.9b01234.
Crystallography, computational results, FTIR, TGA, FE-
SEM, N2 isotherm, and NMR spectra (PDF)
Accession Codes
CCDC 1889952−1889962 contain the supplementary crys-
tallographic data for this paper. These data can be obtained
free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by
emailing data_request@ccdc.cam.ac.uk, or by contacting The
Cambridge Crystallographic Data Centre, 12 Union Road,
Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

■ AUTHOR INFORMATION
Corresponding Authors
*E-mail: nhkhan@csmcri.res.in.
*E-mail: esuresh@csmcri.res.in or sureshe123@rediffmail.com.
ORCID
Bhavesh Parmar: 0000-0003-4263-7635
Parth Patel: 0000-0001-9742-2784
Renjith S. Pillai: 0000-0003-0165-3305
Raj Kumar Tak: 0000-0003-3726-8687
Rukhsana I. Kureshy: 0000-0001-6611-7605
Figure 8. (a) Recyclability of CoMOF-2 for up to six catalytic cycles Noor-ul H. Khan: 0000-0002-5204-5943
for S1 conversion. (b) PXRD data of recovered CoMOF-2 after the Eringathodi Suresh: 0000-0002-1934-6832
sixth cycle for cycloaddition reaction of S1 compared with PXRD data
Author Contributions
of the as-synthesized compound. #
B.P. and P.P. contributed equally to this work.
Notes
material showed a pattern identical to that of the pristine The authors declare no competing financial interest.
compound (Figure 8b and S1). No enhancement of the
product upon removal of the catalyst clearly demonstrates no
metal leaching and authenticates the active sites reside on the
■ ACKNOWLEDGMENTS
The registration number of this publication is CSIR-CSMCRI-
catalyst (Figure S46).


021/2019. Financial support from CSIR-SRF (BP), DST
Project HCP0009 (PP), and analytical support by AESD&CIF
CONCLUSIONS of CSIR-CSMCRI are gratefully acknowledged. The authors
In summary, a mixed ligand phase pure 3D MOF (CoMOF-2) thank Ms. Pooja Goswami for PXRD, Dr. Vinod Boricha for
with good chemical stability has been synthesized and/or NMR data, Mr. Jayesh Chaudhari for FE-SEM analysis, Mr.
crystallized by a RT stirring/diffusion method of the precursors Vinod Kumar Agrawal for FTIR data, Mr. Satyaveer Gothwal
and characterized by different analytical methods. CoMOF-2 for TGA data, Mr. Viral Vakani for CHN analysis, and Mr.
has been employed as a recyclable heterogeneous catalyst of Harpalsinh Rathod for N2/CO2 adsorption−desorption
high utility for conversion of a heterocyclic epoxide under mild analysis. R.S.P. thanks the SRM Supercomputer Center, SRM
reaction condition and utilizing a minimum energy to yield the IST, for providing the computational facility.
corresponding spirocyclic oxindoles efficiently. Spirocyclic
oxindole carbonates are a new series of molecules, and crystal
structures of some of these compounds and substrates have
■ REFERENCES
(1) Li, J.-R.; Sculley, J.; Zhou, H.-C. Metal-Organic Frameworks for
been determined. The microporosity of CoMOF-2 for CO2 Separations. Chem. Rev. 2012, 112, 869−932.
adsorption and Lewis acidic metal and the basic −NH sites in (2) He, Y.; Zhou, W.; Qian, G.; Chen, B. Methane storage in metal-
L facilitate epoxide−CO2 activation and interactions toward organic frameworks. Chem. Soc. Rev. 2014, 43, 5657−5678.

J DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

(3) Yu, J.; Xie, L.-H.; Li, J.-R.; Ma, Y.; Seminario, J. M.; Balbuena, P. Amines Featuring Tetrasubstituted Tertiary Carbons. J. Am. Chem.
B. CO2 Capture and Separations Using MOFs: Computational and Soc. 2016, 138, 14194−14197.
Experimental Studies. Chem. Rev. 2017, 117, 9674−9754. (23) Ninokata, R.; Yamahira, T.; Onodera, G.; Kimura, M. Nickel-
(4) Adil, K.; Belmabkhout, Y.; Pillai, R. S.; Cadiau, A.; Bhatt, P. M.; Catalyzed CO2 Rearrangement of Enol Metal Carbonates for the
Assen, A. H.; Maurin, G.; Eddaoudi, M. Gas/vapour separation using Efficient Synthesis of β-Ketocarboxylic Acids. Angew. Chem., Int. Ed.
ultra-microporous metal-organic frameworks: insights into the 2017, 56, 208−211.
structure/separation relationship. Chem. Soc. Rev. 2017, 46, 3402− (24) Guo, W.; Gónzalez-Fabra, J.; Bandeira, N. A. G.; Bo, C.; Kleij,
3430. A. W. A Metal-Free Synthesis of N-Aryl Carbamates under Ambient
(5) Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van Duyne, Conditions. Angew. Chem., Int. Ed. 2015, 54, 11686−11690.
R. P.; Hupp, J. T. Metal-Organic Framework Materials as Chemical (25) Schäffner, B.; Schäffner, F.; Verevkin, S. P.; Börner, A. Organic
Sensors. Chem. Rev. 2012, 112, 1105−1125. Carbonates as Solvents in Synthesis and Catalysis. Chem. Rev. 2010,
(6) Lustig, W. P.; Mukherjee, S.; Rudd, N. D.; Desai, A. V.; Li, J.; 110, 4554−4581.
Ghosh, S. K. Metal-organic frameworks: functional luminescent and (26) Vivek, J. P.; Berry, N.; Papageorgiou, G.; Nichols, R. J.;
photonic materials for sensing applications. Chem. Soc. Rev. 2017, 46, Hardwick, L. J. Mechanistic Insight into the Superoxide Induced Ring
3242−3285. Opening in Propylene Carbonate Based Electrolytes using in Situ
(7) Horcajada, P.; Gref, R.; Baati, T.; Allan, P. K.; Maurin, G.; Surface-Enhanced Infrared Spectroscopy. J. Am. Chem. Soc. 2016, 138,
Couvreur, P.; Férey, G.; Morris, R. E.; Serre, C. Metal-Organic 3745−3751.
Frameworks in Biomedicine. Chem. Rev. 2012, 112, 1232−1268. (27) He, H.; Perman, J. A.; Zhu, G.; Ma, S. Metal-Organic
(8) DeCoste, J. B.; Peterson, G. W. Metal-Organic Frameworks for Frameworks for CO2 Chemical Transformations. Small 2016, 12,
Air Purification of Toxic Chemicals. Chem. Rev. 2014, 114, 5695− 6309−6324.
5727. (28) Maina, J. W.; Pozo-Gonzalo, C.; Kong, L.; Schütz, J.; Hill, M.;
(9) Dhakshinamoorthy, A.; Alvaro, M.; Garcia, H. Commercial Dumée, L. F. Metal organic framework based catalysts for CO2
metal-organic frameworks as heterogeneous catalysts. Chem. Commun. conversion. Mater. Horiz. 2017, 4, 345−361.
2012, 48, 11275−11288. (29) Trickett, C. A.; Helal, A.; Al-Maythalony, B. A.; Yamani, Z. H.;
(10) Huang, Y.-B.; Liang, J.; Wang, X.-S.; Cao, R. Multifunctional Cordova, K. E.; Yaghi, O. M. The chemistry of metal-organic
metal-organic framework catalysts: synergistic catalysis and tandem frameworks for CO2 capture, regeneration and conversion. Nat. Rev.
reactions. Chem. Soc. Rev. 2017, 46, 126−157. Mater. 2017, 2, 17045.
(11) Dhakshinamoorthy, A.; Asiri, A. M.; Garcia, H. Metal-organic (30) Sharma, N.; Dhankhar, S. S.; Nagaraja, C. M. A Mn(II)-
frameworks catalyzed C-C and C-heteroatom coupling reactions. porphyrin based metal-organic framework (MOF) for visible-light-
Chem. Soc. Rev. 2015, 44, 1922−1947. assisted cycloaddition of carbon dioxide with epoxides. Microporous
(12) Liu, J.; Chen, L.; Cui, H.; Zhang, J.; Zhang, L.; Su, C.-Y. Mesoporous Mater. 2019, 280, 372−378.
Applications of metal-organic frameworks in heterogeneous supra- (31) Guillerm, V.; Weseliński, L. J.; Belmabkhout, Y.; Cairns, A. J.;
molecular catalysis. Chem. Soc. Rev. 2014, 43, 6011−6061. D’Elia, V.; Wojtas, L.; Adil, K.; Eddaoudi, M. Discovery and
(13) Burtch, N. C.; Jasuja, H.; Walton, K. S. Water Stability and
introduction of a (3,18)-connected net as an ideal blueprint for the
Adsorption in Metal-Organic Frameworks. Chem. Rev. 2014, 114,
design of metal-organic frameworks. Nat. Chem. 2014, 6, 673−680.
10575−10612.
(32) Feng, D.; Chung, W.-C.; Wei, Z.; Gu, Z.-Y.; Jiang, H.-L.; Chen,
(14) Emerson, A. J.; Hawes, C. S.; Marshall, M.; Knowles, G. P.;
Y.-P.; Darensbourg, D. J.; Zhou, H.-C. Construction of Ultrastable
Chaffee, A. L.; Batten, S. R.; Turner, D. R. High-Connectivity
Porphyrin Zr Metal-Organic Frameworks through Linker Elimination.
Approach to a Hydrolytically Stable Metal-Organic Framework for
J. Am. Chem. Soc. 2013, 135, 17105−17110.
CO2 Capture from Flue Gas. Chem. Mater. 2018, 30, 6614−6618.
(33) Gao, W.-Y.; Chen, Y.; Niu, Y.; Williams, K.; Cash, L.; Perez, P.
(15) Howarth, A. J.; Liu, Y.; Li, P.; Li, Z.; Wang, T. C.; Hupp, J. T.;
Farha, O. K. Chemical, thermal and mechanical stabilities of metal- J.; Wojtas, L.; Cai, J.; Chen, Y.-S.; Ma, S. Crystal Engineering of an
organic frameworks. Nat. Rev. Mater. 2016, 1, 15018. nbo Topology Metal-Organic Framework for Chemical Fixation of
(16) Yuan, S.; Feng, L.; Wang, K.; Pang, J.; Bosch, M.; Lollar, C.; CO2 under Ambient Conditions. Angew. Chem., Int. Ed. 2014, 53,
Sun, Y.; Qin, J.; Yang, X.; Zhang, P.; Wang, Q.; Zou, L.; Zhang, Y.; 2615−2619.
Zhang, L.; Fang, Y.; Li, J.; Zhou, H.-C. Stable Metal-Organic (34) Beyzavi, M. H.; Klet, R. C.; Tussupbayev, S.; Borycz, J.;
Frameworks: Design, Synthesis, and Applications. Adv. Mater. 2018, Vermeulen, N. A.; Cramer, C. J.; Stoddart, J. F.; Hupp, J. T.; Farha, O.
30, 1704303. K. A Hafnium-Based Metal-Organic Framework as an Efficient and
(17) He, H.; Sun, Q.; Gao, W.; Perman, J. A.; Sun, F.; Zhu, G.; Multifunctional Catalyst for Facile CO2 Fixation and Regioselective
Aguila, B.; Forrest, K.; Space, B.; Ma, S. A Stable Metal-Organic and Enantioretentive Epoxide Activation. J. Am. Chem. Soc. 2014, 136,
Framework Featuring a Local Buffer Environment for Carbon Dioxide 15861−15864.
Fixation. Angew. Chem., Int. Ed. 2018, 57, 4657−4662. (35) Alkordi, M. H.; Weseliński, L. J.; D’Elia, V.; Barman, S.; Cadiau,
(18) Aresta, M.; Dibenedetto, A.; Angelini, A. Catalysis for the A.; Hedhili, M. N.; Cairns, A. J.; AbdulHalim, R. G.; Basset, J.-M.;
Valorization of Exhaust Carbon: from CO2 to Chemicals, Materials, Eddaoudi, M. CO2 conversion: the potential of porous-organic
and Fuels. Technological Use of CO2. Chem. Rev. 2014, 114, 1709− polymers (POPs) for catalytic CO2-epoxide insertion. J. Mater. Chem.
1742. A 2016, 4, 7453−7460.
(19) Markewitz, P.; Kuckshinrichs, W.; Leitner, W.; Linssen, J.; (36) Kathalikkattil, A. C.; Babu, R.; Roshan, R. K.; Lee, H.; Kim, H.;
Zapp, P.; Bongartz, R.; Schreiber, A.; Müller, T. E. Worldwide Tharun, J.; Suresh, E.; Park, D.-W. An lcy-topology amino acid MOF
innovations in the development of carbon capture technologies and as eco-friendly catalyst for cyclic carbonate synthesis from CO2:
the utilization of CO2. Energy Environ. Sci. 2012, 5, 7281−7305. Structure-DFT corroborated study. J. Mater. Chem. A 2015, 3,
(20) Rokicki, G.; Parzuchowski, P. G.; Mazurek, M. Non-isocyanate 22636−22647.
polyurethanes: synthesis, properties, and applications. Polym. Adv. (37) Aguila, B.; Sun, Q.; Wang, X.; O’Rourke, E.; Al-Enizi, A. M.;
Technol. 2015, 26, 707−761. Nafady, A.; Ma, S. Lower Activation Energy for Catalytic Reactions
(21) Yang, L.-C.; Rong, Z.-Q.; Wang, Y.-N.; Tan, Z. Y.; Wang, M.; through Host-Guest Cooperation within Metal-Organic Frameworks.
Zhao, Y. Construction of Nine-Membered Heterocycles through Angew. Chem., Int. Ed. 2018, 57, 10107−10111.
Palladium-Catalyzed Formal [5 + 4] Cycloaddition. Angew. Chem., Int. (38) Zhou, Z.; He, C.; Xiu, J.; Yang, L.; Duan, C. Metal-Organic
Ed. 2017, 56, 2927−2931. Polymers Containing Discrete Single-Walled Nanotube as a
(22) Cai, A.; Guo, W.; Martínez-Rodríguez, L.; Kleij, A. W. Heterogeneous Catalyst for the Cycloaddition of Carbon Dioxide to
Palladium-Catalyzed Regio-and Enantioselective Synthesis of Allylic Epoxides. J. Am. Chem. Soc. 2015, 137, 15066−15069.

K DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

(39) Kurisingal, J. F.; Rachuri, Y.; Pillai, R. S.; Gu, Y.; Choe, Y.; Park, (57) Antonchick, A. P.; Gerding-Reimers, C.; Catarinella, M.;
D.-W. Ionic liquid-functionalized UiO-66 framework: An experimen- Schürmann, M.; Preut, H.; Ziegler, S.; Rauh, D.; Waldmann, H.
tal and theoretical study on the cycloaddition of CO2 and epoxide. Highly enantioselective synthesis and cellular evaluation of spiroox-
ChemSusChem 2019, 12, 1033−1042. indoles inspired by natural products. Nat. Chem. 2010, 2, 735−740.
(40) Nguyen, P. T. K.; Nguyen, H. T. D.; Nguyen, H. N.; Trickett, (58) Vintonyak, V. V.; Warburg, K.; Kruse, H.; Grimme, S.; Hübel,
C. A.; Ton, Q. T.; Gutiérrez-Puebla, E.; Monge, M. A.; Cordova, K. K.; Rauh, D.; Waldmann, H. Identification of Thiazolidinones Spiro-
E.; Gándara, F. New Metal-Organic Frameworks for Chemical Fused to Indolin-2-ones as Potent and Selective Inhibitors of the
Fixation of CO2. ACS Appl. Mater. Interfaces 2018, 10, 733−744. Mycobacterium tuberculosis Protein Tyrosine Phosphatase B. Angew.
(41) Ding, M.; Jiang, H.-L. Incorporation of Imidazolium-Based Chem., Int. Ed. 2010, 49, 5902−5905.
Poly(ionic liquid)s into a Metal-Organic Framework for CO2 Capture (59) Tak, R. K.; Gupta, N.; Kumar, M.; Kureshy, R. I.; Khan, N. H.;
and Conversion. ACS Catal. 2018, 8, 3194−3201. Suresh, E. Regioselective Alcoholysis and Hydrochlorination Reac-
(42) De, D.; Pal, T. K.; Neogi, S.; Senthilkumar, S.; Das, D.; Gupta, tions of Spiro-Epoxy Oxindoles at the Spiro-Centre: Synthesis of 3,3-
S. S.; Bharadwaj, P. K. A Versatile CuII Metal-Organic Framework Disubstituted Oxindoles and Application for Anticancer Agents. Eur.
Exhibiting High Gas Storage Capacity with Selectivity for CO2: J. Org. Chem. 2018, 2018, 5678−5687.
Conversion of CO2 to Cyclic Carbonate and Other Catalytic Abilities. (60) Tak, R. K.; Patel, P.; Subramanian, S.; Kureshy, R. I.; Khan, N.
Chem. - Eur. J. 2016, 22, 3387−3396. H. Cycloaddition Reaction of Spiro-Epoxy Oxindole with CO2 at
(43) Subramanian, S.; Park, J.; Byun, J.; Jung, Y.; Yavuz, C. T. Highly Atmospheric Pressure Using Deep Eutectic Solvent. ACS Sustainable
Efficient Catalytic Cyclic Carbonate Formation by Pyridyl Salicyli- Chem. Eng. 2018, 6, 11200−11205.
mines. ACS Appl. Mater. Interfaces 2018, 10, 9478−9484. (61) Patel, P.; Parmar, B.; Kureshy, R. I.; Khan, N. H.; Suresh, E.
(44) Ugale, B.; Kumar, S.; Dhilip Kumar, T. J.; Nagaraja, C. M. Efficient Solvent-Free Carbon Dioxide Fixation Reactions with
Environmentally Friendly, Co-catalyst-Free Chemical Fixation of CO2 Epoxides Under Mild Conditions by Mixed-Ligand Zinc(II) Metal-
at Mild Conditions Using Dual-Walled Nitrogen-Rich Three-Dimen- Organic Frameworks. ChemCatChem 2018, 10, 2401−2408.
sional Porous Metal-Organic Frameworks. Inorg. Chem. 2019, 58, (62) Patel, P.; Parmar, B.; Kureshy, R. I.; Khan, N. H.; Suresh, E.
3925−3936. Amine-functionalized Zn(II) MOF as an efficient multifunctional
(45) Yue, S.; Wang, P.; Hao, X. Synthesis of cyclic carbonate from catalyst for CO2 utilization and sulfoxidation reaction. Dalton Trans.
CO2 and epoxide using bifunctional imidazolium ionic liquid under 2018, 47, 8041−8051.
mild conditions. Fuel 2019, 251, 233−241. (63) Parmar, B.; Patel, P.; Kureshy, R. I.; Khan, N. H.; Suresh, E.
(46) Sun, Y.; Huang, H.; Vardhan, H.; Aguila, B.; Zhong, C.; Sustainable Heterogeneous Catalysts for CO2 Utilization by Using
Perman, J. A.; Al-Enizi, A. M.; Nafady, A.; Ma, S. Facile Approach to Dual Ligand ZnII/CdII Metal-Organic Frameworks. Chem. - Eur. J.
Graft Ionic Liquid into MOF for Improving the Efficiency of CO2 2018, 24, 15831−15839.
(64) Parmar, B.; Patel, P.; Pillai, R. S.; Kureshy, R. I.; Khan, N. H.;
Chemical Fixation. ACS Appl. Mater. Interfaces 2018, 10, 27124−
Suresh, E. Efficient Catalytic Conversion of Terminal/ Internal
27130.
Epoxides to Cyclic Carbonates by Porous Co(II) MOF at Ambient
(47) Singh, G. S.; Desta, Z. Y. Isatins As Privileged Molecules in
Conditions: Structure Property Correlation and Computational
Design and Synthesis of Spiro-Fused Cyclic Frameworks. Chem. Rev.
Studies. J. Mater. Chem. A 2019, 7, 2884−2894.
2012, 112, 6104−6155.
(65) Parmar, B.; Rachuri, Y.; Bisht, K. K.; Suresh, E. Syntheses and
(48) Yu, B.; Yu, D.-Q.; Liu, H.-M. Spirooxindoles: Promising
Structural Analyses of New 3D Isostructural Zn(II) and Cd(II)
scaffolds for anticancer agents. Eur. J. Med. Chem. 2015, 97, 673−698.
Luminescent MOFs and their Application Towards Detection of
(49) Zhou, R.; Wu, Q.; Guo, M.; Huang, W.; He, X.; Yang, L.; Peng,
Nitroaromatics in Aqueous Media. Chemistry Select 2016, 1, 6308−
F.; He, G.; Han, B. Organocatalytic cascade reaction for the
6315.
asymmetric synthesis of novel chroman-fused spirooxindoles that (66) Chouhan, M.; Senwar, K. R.; Sharma, R.; Grover, V.; Nair, V.
potently inhibit cancer cell proliferation. Chem. Commun. 2015, 51, A. Regiospecific epoxide opening: a facile approach for the synthesis
13113−13116. of 3-hydroxy-3-aminomethylindolin-2-one derivatives. Green Chem.
(50) Zhang, M.; Fu, Q.-Y.; Gao, G.; He, H.-Y.; Zhang, Y.; Wu, Y.-S.; 2011, 13, 2553−2560.
Zhang, Z.-H. Catalyst-Free, Visible-Light Promoted One-Pot Syn- (67) Tak, R. K.; Gupta, N.; Kumar, M.; Kureshy, R. I.; Khan, N. H.;
thesis of Spirooxindole-Pyran Derivatives in Aqueous Ethyl Lactate. Suresh, E. Regioselective Alcoholysis and Hydrochlorination Reac-
ACS Sustainable Chem. Eng. 2017, 5, 6175−6182. tions of Spiro-Epoxy Oxindoles at the Spiro-Centre: Synthesis of 3,3-
(51) Mao, H.; Lin, A.; Tang, Y.; Shi, Y.; Hu, H.; Cheng, Y.; Zhu, C. Disubstituted Oxindoles and Application for Anticancer Agents. Eur.
Organocatalytic oxa/aza-Michael-Michael Cascade Strategy for the J. Org. Chem. 2018, 2018, 5678−5687.
Construction of Spiro [Chroman/Tetrahydroquinoline-3,3′-oxindole] (68) Zheng, B.; Bai, J.; Duan, J.; Wojtas, L.; Zaworotko, M. J.
Scaffolds. Org. Lett. 2013, 15, 4062−4065. Enhanced CO2 Binding Affinity of a High-Uptake rht-Type Metal-
(52) Kouznetsov, V. V.; Merchan Arenas, D. R.; Arvelo, F.; Forero, J. Organic Framework Decorated with Acylamide Groups. J. Am. Chem.
S. B.; Sojo, F.; Munoz, A. 4-Hydroxy-3-methoxyphenyl Substituted 3- Soc. 2011, 133, 748−751.
methyl-tetrahydroquinoline Derivatives Obtained Through Imino (69) Roztocki, K.; Jȩdrzejowski, D.; Hodorowicz, M.; Senkovska, I.;
Diels-Alder Reactions as Potential Antitumoral Agents. Lett. Drug Kaskel, S.; Matoga, D. Isophthalate-Hydrazone 2D Zinc-Organic
Des. Discovery 2010, 7, 632−639. Framework: Crystal Structure, Selective Adsorption, and Tuning of
(53) Ramakumar, K.; Maji, T.; Partridge, J. J.; Tunge, J. A. Synthesis Mechanochemical Synthetic Conditions. Inorg. Chem. 2016, 55,
of Spirooxindoles via the tert-Amino Effect. Org. Lett. 2017, 19, 9663−9670.
4014−4017. (70) Stylianou, K. C.; Warren, J. E.; Chong, S. Y.; Rabone, J.; Bacsa,
(54) Galliford, C. V.; Scheidt, K. A. Pyrrolidinyl-Spirooxindole J.; Bradshaw, D.; Rosseinsky, M. J. CO2 selectivity of a 1D
Natural Products as Inspirations for the Development of Potential microporous adenine-based metal-organic framework synthesised in
Therapeutic Agents. Angew. Chem., Int. Ed. 2007, 46, 8748−8758. water. Chem. Commun. 2011, 47, 3389−3391.
(55) Marti, C.; Carreira, E. M. Construction of Spiro[pyrrolidine- (71) Chen, M.; Chen, S.; Chen, W.; Lucier, B. E. G.; Zhang, Y.;
3,3′-oxindoles] - Recent Applications to the Synthesis of Oxindole Zheng, A.; Huang, Y. Analyzing Gas Adsorption in an Amide-
Alkaloids. Eur. J. Org. Chem. 2003, 2003, 2209−2219. Functionalized Metal Organic Framework: Are the Carbonyl or
(56) Zhu, G.; Bao, G.; Li, Y.; Sun, W.; Li, J.; Hong, L.; Wang, R. Amine Groups Responsible? Chem. Mater. 2018, 30, 3613−3617.
Efficient Catalytic Kinetic Resolution of Spiro-epoxyoxindoles with (72) Hajra, S.; Maity, S.; Roy, S. Regioselective Friedel-Crafts
Concomitant Asymmetric Friedel-Crafts Alkylation of Indoles. Angew. Reaction of Electron-Rich Benzenoid Arenes and Spiroepoxyoxindole
Chem., Int. Ed. 2017, 56, 5332−5335. at the Spiro-Centre: Efficient Synthesis of Benzofuroindolines and

L DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX
Inorganic Chemistry Article

2H- Spiro[benzofuran]-3,3′-oxindoles. Adv. Synth. Catal. 2016, 358,


2300−2306.
(73) Jiang, H.; Jie, H.; Li, J. Scandium triflate-catalyzed selective ring
opening and rearrangement reaction of spiro-epoxyoxindole and
carbonyl compounds. RSC Adv. 2016, 6, 100307−100311.
(74) Kim, J.; Kim, S.-N.; Jang, H.-G.; Seo, G.; Ahn, W.-S. CO2
cycloaddition of styrene oxide over MOF catalysts. Appl. Catal., A
2013, 453, 175−180.
(75) Li, P.-Z.; Wang, X.-J.; Liu, J.; Phang, H. S.; Li, Y.; Zhao, Y.
Highly Effective Carbon Fixation via Catalytic Conversion of CO2 by
an Acylamide-Containing Metal-Organic Framework. Chem. Mater.
2017, 29, 9256−9261.
(76) Guo, X.; Zhou, Z.; Chen, C.; Bai, J.; He, C.; Duan, C. New rht-
Type Metal-Organic Frameworks Decorated with Acylamide Groups
for Efficient Carbon Dioxide Capture and Chemical Fixation from
Raw Power Plant Flue Gas. ACS Appl. Mater. Interfaces 2016, 8,
31746−31756.

M DOI: 10.1021/acs.inorgchem.9b01234
Inorg. Chem. XXXX, XXX, XXX−XXX

You might also like