0% found this document useful (0 votes)
18 views11 pages

Giri Et Al 2019 Connecting The Dots Knitting C Phenylresorcin 4 Arenes With Aromatic Linkers For Task Specific Porous

The article discusses the development of task-specific porous organic polymers (POPs) by knitting C-phenylresorcin[4]arene with various aromatic linkers to enhance their properties for applications in molecular separation, catalysis, and environmental remediation. The study demonstrates that these interconnected porous structures exhibit improved catalytic activity, gas uptake, and pollutant removal efficiency compared to their zero-dimensional counterparts. The findings highlight the potential of integrating supramolecular chemistry with porous materials to address challenges in green energy and environmental issues.

Uploaded by

krs.sam1991
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
18 views11 pages

Giri Et Al 2019 Connecting The Dots Knitting C Phenylresorcin 4 Arenes With Aromatic Linkers For Task Specific Porous

The article discusses the development of task-specific porous organic polymers (POPs) by knitting C-phenylresorcin[4]arene with various aromatic linkers to enhance their properties for applications in molecular separation, catalysis, and environmental remediation. The study demonstrates that these interconnected porous structures exhibit improved catalytic activity, gas uptake, and pollutant removal efficiency compared to their zero-dimensional counterparts. The findings highlight the potential of integrating supramolecular chemistry with porous materials to address challenges in green energy and environmental issues.

Uploaded by

krs.sam1991
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 11

Article

Cite This: Chem. Mater. 2019, 31, 8440−8450 pubs.acs.org/cm

Connecting the Dots: Knitting C‑Phenylresorcin[4]arenes with


Aromatic Linkers for Task-Specific Porous Organic Polymers
Arkaprabha Giri, MD. Waseem Hussain, Bahadur Sk, and Abhijit Patra*
Indian Institute of Science Education and Research Bhopal, Bhopal By-Pass Road, Bhauri, Bhopal 462066, Madhya Pradesh, India
*
S Supporting Information

ABSTRACT: Macrocyclic cavitands having aesthetically


appealing architectures and excellent host−guest complex-
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ation abilities exhibit a broad range of applications from


molecular separation, catalysis, and sensing to drug delivery.
However, the close packing of the zero-dimensional (0D)
Downloaded via DARTMOUTH COLG on December 14, 2024 at 19:48:02 (UTC).

porous cavitands reduces their activity in the solid state.


Knitting the macrocyclic cavitands using suitable aromatic
linkers may lead to a new generation of porous organic
polymers (POPs) where the intrinsic properties of the
cavitands can be augmented through interconnected pores
in the solid state. Herein, we demonstrate the design strategy
of linking the 0D-discrete pores by connecting C-
phenylresorcin[4]arene (RN4) through three different
aromatic linkers. The flexible azo linkers generate highly dispersible hierarchically mesoporous POP (RN4-Az-OH) exhibiting
a remarkable catalytic activity toward metal-free cycloaddition of CO2 with epoxides under solvent-free reaction conditions.
Alkyne-based rigid linkers lead to microporous polymer (RN4-OH), which shows CO2 and H2 uptake at low pressure. The
fluorine-rich linkers produce ultramicroporous hydrophobic POP (RN4-F) exhibiting high efficiency toward the charge-specific
size-selective removal of organic micropollutants from water. Resorcin[4]arene-derived POPs show superior performance
compared to the pristine 0D-porous building units. Thus, “connecting the dots” (0D pores) gives rise to a new interface
between supramolecular chemistry and porous organic materials, which can be explored further to address the challenging
problems related to green energy and environmental remediation.

■ INTRODUCTION
Applications of porous materials depend primarily on (i)
assembly of the macrocycles, it is challenging to maintain the
permanent porosity upon guest removal.
On the other hand, zeolites, metal−organic frameworks
accessibility of the pores and (ii) permeability of the probe
(MOFs), covalent organic frameworks (COFs), polymers of
molecules through the pores. Based on the two parameters,
intrinsic microporosity (PIMs), and porous organic polymers
porous materials can be categorized as having (a) open
(POPs) possess open, continuous two-dimensional (2D), or
porosity and (b) closed porosity.1,2 Open porosity refers to the three-dimensional (3D) rigid networks.7−17 These materials
pores where the probe molecules can easily permeate from one are highly porous because of effective polymeric connection
pore to another, and closed porosity refers to zero-dimensional (covalent or coordinate bonds), leading to shape-persistent
(0D) pores where molecular occupants cannot flow. Macro- interconnected pores (Figure 1a). A possible avenue to retain
cyclic cavitands like cyclodextrins, calix[n]arenes, and pillar- the guest-responsive properties of 0D-macrocyclic cavitands in
[n]arenes possess well-defined cavities and excellent guest the solid state is to convert them into 2D/3D frameworks
recognition properties, which trigger an avalanche of through suitable linkers. The fruits of integrating macrocyclic
applications ranging from molecular separation, sensing, cavitands have been realized in the recent reports of cavitand-
catalysis, mimicking enzyme activity, and drug delivery to the based porous polymers (Figure 1b).18−22 Dichtel and co-
development of artificial molecular machines.3−6 These workers developed the well-known macrocyclic cavitand, β-
interesting features of the cavitands are mostly demonstrated cyclodextrin (β-CD)-based POP (SBET 263 m2 g−1) for the
in the solution state. However, in the solid state, the close rapid removal of organic micropollutants from water.18 Dai
packing of the macrocyclic cavitands mostly reduces the open and co-workers reported β-CD-based hypercross-linked POP
channels or voids spaces leading to the 0D-discrete pores.2 (SBET, 1225 m2 g−1) for adsorption of aromatic molecules from
Thus, the cavities are buried inside the solid matrix and are
largely inaccessible to the molecular species (Figure 1a). The Received: June 28, 2019
narrow channels restrict facile mass transfer. At the same time, Revised: September 1, 2019
due to the noncovalent interactions in the supramolecular Published: September 11, 2019

© 2019 American Chemical Society 8440 DOI: 10.1021/acs.chemmater.9b02563


Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

Figure 1. (a) Schematic illustration of connecting the zero-dimensional (0D) porous macrocyclic cavitands by aromatic linkers leading to the
porous organic polymers (POPs). The interconnected pores are accessible to the guests, and the unique properties of the cavitands are enhanced in
the solid state. (b) Common macrocyclic building units (e.g., cyclodextrin, pillar[n]arene, and calix[n]arene) used for the fabrication of porous
solids for varied applications. Our approach is to choose a unique macrocyclic building block having multiple reactive sites and connect it with
aromatic linkers (tuning of linker length and diverse functionality) for the fabrication of task-specific POPs. (c) Knitting the p-substituted C-
phenylresorcin[4]arene (RN4) core, exploiting three reactive sites (denoted by red, green, and blue spheres) through different linkers, generates
the POPs having enhanced porosity and superior performance.

water.19 Coskun and co-workers employed pillar[5]arene for literature, the environmentally benign applications exhibited
the development of conjugated microporous polymers (CMPs, by the RN4-based POPs as well as the detailed structure−
SBET 400 m2 g−1) for propane/methane separation.20 Trabolsi activity relationship established in the present study are unique
and co-workers developed calix[4]arene-based POPs (SBET 596 (Tables S5−S8) and would contribute to the development of
m2 g−1) for water purification.21 Yuan and co-workers recently the task-specific porous materials.
reported calix[4]resorcinarene-based POPs (SBET 303 m2 g−1)
for iodine adsorption.22 A similar approach of “macrocycle/
cage-to-framework” was demonstrated by Zhang and co-
■ RESULTS AND DISCUSSION
Resorcin[n]arenes (n = 4, 6, and 8) are structurally quite
workers.23−25 Such frameworks were shown to exhibit high gas distinct compared to highly investigated macrocyclic hosts like
uptake with excellent selectivity.25 crown ethers, cyclodextrins, cucurbit[n]urils, calix[n]arenes,
Resorcin[n]arenes are known for the cationic guest and pillar[n]arenes (Figure S1).5 Resorcin[n]arenes can easily
encapsulation and also as catalysts for various chemical be prepared from inexpensive starting materials such as
reactions owing to their intrinsic porosity.6 We employed C- resorcinol and corresponding aromatic/aliphatic aldehyde
phenylresorcin[4]arene (RN4, Figure S1) as a new 0D porous (Figure 2a). The substitution by phenyl groups in the lower
macrocyclic building block to augment the guest-responsive rim of resorcin[4]arene deepens the cavity and provides
properties in the form of a 3D porous solid (Figure 1). We additional functionality to facilitate the polycondensation.
connected RN4 with a series of diverse functional linkers Thus, C-phenylresorcin[4]arene (RN4) is a versatile, func-
(Figure 1c), and it could tune the specific surface area of the tional building block compared to pristine resorcin[4]arene. p-
POPs from 340 to 1230 m2 g−1, much higher than that of p- Hydroxyphenyl (HRN4) and p-bromophenylresorcin[4]arene
hydroxyphenylresorcin[4]arene (HRN4, 156 m2 g−1). We cores were synthesized through the Brønsted acid-catalyzed
observed that the linker length and functionality governed the condensation reactions of resorcinol with p-hydroxy and p-
physicochemical properties of the RN4-based POPs (Figure bromobenzaldehyde in a one-pot synthesis, respectively
S2). The flexible and long linkers containing CO2-philic diazo (Figure 2a, Scheme S1, and Figures S38−S43).27
groups provide hierarchically porous well-dispersible POP, Synthesis and Characterization of the POPs. The
RN4-Az-OH. It turned out to be a promising heterogeneous POPs were synthesized through functionalizing upper and
catalyst facilitating a remarkable metal-free conversion of CO2 lower rims and utilizing phenolic hydroxyl groups of RN4
and epoxide into cyclic organic carbonates under solvent-free employing three different fabrication procedures (Figure 2a,
reaction conditions (Table S7). The rigid and short alkyne Figure S2a). Liu and co-workers developed a facile route of
linkers impart microporosity in RN4-OH suitable for gas making azo-linked POPs through a simple diazo coupling
uptake at low pressure (Table S6). The fluorine-rich linkers reaction between aryl amines and phenols.28 Envisioning the
endow hydrophobicity in the resultant polymer,26 RN4-F, wider scope of the methodology and scalability, we chose p-
which exhibits charge-specific size-selective removal of organic hydroxyphenylresorcin[4]arene (HRN4) as the polyphenolic
micropollutants from water (Table S8). Thus, the objective of core and benzidine as an aromatic diamine linker for the
enhanced solid-state applications of macrocyclic cavitand- fabrication of azo-linked POP, RN4-Az-OH in an aqueous
based POPs compared to pristine cavitands was unequivocally medium (yield: 90−95%). Another scalable synthetic protocol,
demonstrated through selective gas uptake, CO2 conversion, the aromatic nucleophilic substitution reaction between p-
and micropollutant separation. In view of the existing hydroxyphenylresorcin[4]arene (HRN4) and tetrafluoroter-
8441 DOI: 10.1021/acs.chemmater.9b02563
Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

Figure 2. (a) Synthetic schemes of C-phenylresorcin[4]arene-based porous organic polymers (POPs). RN4-Az-OH: diazo coupling between p-
hydroxyphenylresorcin[4]arene and benzidine, RN4-OH: Sonogashira polycondensation between p-bromophenylresorcin[4]arene and 1,4-
diethynylbenzene, and RN4-F: aromatic nucleophilic substitution reaction between p-hydroxyphenylresorcin[4]arene and tetrafluorotereph-
thalonitrile (12 phenolic OH groups are likely to be equally reactive; a more generalized structure shown in Figure S2a). Digital photographs of the
respective POPs in the form of powder. (b) FTIR spectra, (c) solid-state 13C- (CP/MAS) NMR spectra, and (d) thermogravimetric analyses of the
POPs: RN4-Az-OH, RN4-OH, and RN4-F.

ephthalonitrile, was employed for the fabrication of fluorine- aromatic carbons present in the macrocyclic core as well as in
rich POP, RN4-F (yield: 60−70%). Alkyne-linked POP, RN4- the linkers (Figure 2c, Figures S3−S5). The peak at 40−45
OH, was developed through Pd(II)-catalyzed A4-B2 type ppm attributed to the aliphatic carbon (C−H) of the C-
Sonogashira cross-coupling polycondensation between p- phenylresorcin[4]arene core. The aromatic carbons linked
bromophenylresorcin[4]arene and 1,4-diethynylbenzene with the azo-groups in RN4-Az-OH appeared at 150 ppm
(yield ∼50%). All the POPs are insoluble in water and (Figure 2c, Figure S3). The presence of CC in RN4-OH was
common organic solvents and are stable upon exposure to confirmed by the peaks near 80 ppm (Figure 2c, Figure S4).
acids and bases providing a facile heterogeneous platform. The carbons with ether linkages (C−O−C) in RN4-F
The structural characterizations of POPs were carried out resonated at 140 ppm (Figure 2c, Figure S5). The X-ray
through detailed spectroscopic and microscopic analyses. The photoelectron spectroscopy (XPS) analysis further corrobo-
broad peaks around 3300−3400 cm−1 in the FTIR spectra of rated the formation of POPs (Figure S6). Thermogravimetric
all POPs were due to the aromatic-OH stretching of the analysis (TGA) profiles of the POPs revealed that the RN4-Az-
resorcin[4]arene core (Figure 2b). In the case of RN4-Az-OH, OH was stable up to 250 °C, whereas thermal stabilities of
the characteristic peaks at 1160 and 1400 cm−1 were attributed RN4-OH and RN4-F were found to be up to 350 °C (Figure
to the symmetric and asymmetric vibrations of the NN 2d). The broad powder X-ray diffraction patterns (PXRD)
linkages, respectively, indicating the formation of the azo-POP. suggested the amorphous nature of the POPs as expected due
In RN4-OH, the peak appeared at 2200 cm−1 was due to the to the kinetically driven irreversible polymerization conditions
CCC stretching confirming the C−C bond formation (Figure S7).29
(Figure 2b). RN4-F showed a peak around 2246 cm−1 due to The diverse functionalities present in the linkers and
the CN stretching (Figure 2b). The peaks at 1304 and 1052 different fabrication methods have a profound effect on the
cm−1 were attributed to the C−F and C−O−C stretching, morphology of the POPs as revealed by the field emission
respectively, which further confirmed the formation of RN4-F. scanning electron microscopy (FESEM) images (Figure 3a−c,
The solid-state 13C NMR spectra of POPs revealed the Figure S8). The fused sphere-like morphology was observed
peaks in the range of 100−165 ppm corresponding to the for RN4-Az-OH and RN4-OH, whereas large aggregated
8442 DOI: 10.1021/acs.chemmater.9b02563
Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

Figure 3. FESEM images of (a) RN4-Az-OH, (b) RN4-OH, and (c) RN4-F; inset: the respective HRTEM images. (d) Digital photographs of
POPs dispersed in water. (e) Nitrogen sorption isotherms of RN4-F, RN4-OH, RN4-Az-OH, and p-hydroxyphenylresorcin[4]arene (HRN4) at 77
K (filled circles: adsorption, hollow circles: desorption); (f) the corresponding pore size distributions of POPs using nonlocal density functional
theory (NLDFT) analysis. (g) CO2 adsorption isotherms of RN4-F, RN4-OH, RN4-Az-OH, and HRN4 at 273 K.

particles of RN4-F were noticeable. The high-resolution volumes of RN4-Az-OH, RN4-OH, and RN4-F at P/P0 = 0.95
transmission electron microscopy (HRTEM) images of the were found to be 0.55, 0.71, and 0.95 cm3 g−1, respectively,
POPs showed the irregular distribution of pores (Figure 3a−c, indicating the promising scope of the POPs as sorbents. We
inset, Figure S9) further suggesting the amorphous nature.29 demonstrated three mutually connected applications, namely,
The size and functionality of the linkers also influence the selective gas uptake, catalytic conversion of CO2, and removal
dispersibility of the POPs. RN4-OH and RN4-F are hydro- of organic micropollutants from water to substantiate the
phobic due to the presence of alkyne and fluorine-rich linkers, benefits of interconnected pores in macrocyclic cavitand-based
respectively.21,26 Owing to the hydrophobic nature, RN4-OH POPs.
and RN4-F particles float on the surface of water (Figure 3d). Carbon Dioxide and Hydrogen Uptake. The heter-
Nitrogen Sorption Isotherm. The surface area and oatom (O, N, and F)-rich networks along with the high surface
porosity of POPs were estimated by nitrogen adsorption− area, prompted us to measure the CO2 and H2 uptake capacity
desorption isotherms at 77 K (Figure 3e, Figure S11). p- of the POPs (Table 1, Figure 3g, and Figure S13). CO2 uptake
Hydroxyphenylresorcin[4]arene (HRN4) showed type III
isotherms indicating weak adsorbate−adsorbent interac- Table 1. Summary of Gas Sorption Properties of
tion.30,31 The Brunauer−Emmett−Teller (BET) specific sur- Resorcinarene-Based POPs and Macrocyclic Core
face area of the monomer HRN4 was found to be 156 m2 g−1 pore CO2 H2
arising due to the interparticle void space (Figures S11d and SBET size pore volume uptake uptake
S12).31 After knitting with aromatic linkers, the connectivity POP (m2 g−1) (nm) (cm3 g−1)a (wt %)b (wt %)c
among the pores was established, and a clear trend of an HRN4 156 2.8 0.28 6.1 0.3
increase in the surface area was observed. The BET specific RN4-Az-OH 340 2.8 0.55 9.0 1.1
surface areas of RN4-Az-OH, RN4-OH, and RN4-F were RN4-OH 720 1.4 0.71 9.9 2.0
found to be 340, 720, and 1230 m2 g−1, respectively. The high RN4-F 1230 0.7 0.95 11.4 1.4
surface area of RN4-F among the cavitand-based POPs is due a
Pore volumes calculated at P/P0 = 0.95. bCO2 uptake measured at
to the smaller length of the F-rich linker as well as a highly 273 K. cH2 uptake measured at 77 K.
cross-linkable polymerization condition (Section S3.9, Table
S5). by RN4-F, RN4-OH, and RN4-Az-OH was found to be 11.4
The pore size distributions of POPs were estimated by wt % (2.6 mmol g−1), 9.9 wt % (2.3 mmol g−1), and 9 wt %
nonlocal density functional theory (NLDFT) analysis (Figure (2.1 mmol g−1) at 273 K, 1 bar, respectively. The surface area
3f). The average pore width of RN4-Az-OH was in the of RN4-OH is almost double than that of RN4-Az-OH.
mesoporous region (2.8 nm) due to the larger azo linker. However, both the POPs exhibit almost similar CO2 uptake
Additionally, RN4-Az-OH exhibited hierarchical pore size capacity. This interesting observation led us to estimate the
distribution (Figure 3f). In contrast, RN4-OH possesses the isosteric heat of adsorption (Qst) for probing the interactions
pores in the microporous region (1.4 nm). In the case of RN4- between CO2 and POPs (Figure S13b). RN4-Az-OH possesses
F, the narrow bimodal pore size distribution was observed at a significantly high Qst of 30.8 kJ mol−1 at the onset of
0.7 nm (ultramicroporous) and 1.4 nm (Figure 3f). The peaks adsorption.33−35 The highly decorated pore walls with CO2-
observed for the POPs in the microporous regime are likely to philic functionalities like NN and −OH lead to
be contributed by the intrinsic pores of resorcinarene-based favorable interactions between RN4-Az-OH and CO2. RN4-F
building blocks (Figure 1c, Figure S12).32 The total pore and RN4-OH have an initial Qst of 29.5 and 28.5 kJ mol−1,
8443 DOI: 10.1021/acs.chemmater.9b02563
Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

respectively. Employing the ideal adsorbed solution theory The reverse trend of catalytic performance with respect to
(IAST) model,36,37 the selectivity of CO2 over N2 uptake for the specific surface area of the POPs can be explained
RN4-Az-OH, RN4-OH, and RN4-F was found to be 19, 20, considering the following aspects: (a) porosity, (b) availability
and 35, respectively (Figure S14). Being ultramicroporous, of phenolic −OH groups, (c) interaction of CO2 with the pore
RN4-F exhibits a high selectivity toward CO2 over N2.38,39 wall, and finally (d) dispersibility of the catalyst in the reaction
The H2 uptake measurement was carried out at 77 K and 1 medium. The hierarchical mesoporosity of RN4-Az-OH
atm. High uptake values of 1.4 wt % (6.9 mmol g−1), 2.0 wt % improves the mass transfer, while ultramicropores of RN4-F
(10.0 mmol g−1), and 1.1 wt % (5.5 mmol g−1) were observed pose a restriction for the easy permeation of the reagents and
for RN4-F, RN4-OH, and RN4-Az-OH, respectively (Figure products through the pores.51 The lesser number of phenolic
S13c). The uptake capacity of the POPs is quite significant −OH groups in RN4-OH (in the monomeric core, p-
compared to many of the reported porous materials (Table bromophenylresorcin[4]arene) and RN4-F (formation of
S6).17 The high H2 uptake capacity of RN4-OH can be ether linkages) compared to RN4-Az-OH further reduce the
attributed to the presence of residual palladium nanoparticles epoxide activation sites. Additionally, the high value of the heat
in the polymer,40 which is substantiated through the PXRD of adsorption of RN4-Az-OH (30.8 kJ mol−1) indicates the
[peak at 2θ = 40° corresponding to the Pd (111) plane],41 facile interaction of CO2 with the diazo and phenolic −OH-
EDS, and HRTEM data (Figures S7b and S15−S17). functionalized pore walls, which further facilitates the CO2
Cycloaddition of CO2 with Epoxides. The catalysis insertion. The high dispersibility of RN4-Az-OH in polar
inside the molecular container has several advantages over that solvents as well as in the reaction medium makes the
of the bulk solution.6,42 The restricted internal volume of the catalytically active centers more expose to the substrates
cavitand provides size-selective substrate encapsulation.43 In (Figure 4b).52,53 The high negative zeta potential of −32.5 mV
the case of resorcin[4]arenes, a large number of phenolic −OH in water at neutral pH reaffirms the stability of the dispersion
groups can activate various substrates, including epoxides (Table S2, Figures S18 and S19).
through H-bonding.6,44 However, often the degradation of the It has been reported that the π-electron-rich aromatic cavity
cavitand in the reaction condition prevents its reusability.44 As of resorcin[4]arene provides facile interaction with cationic
RN4-based POPs exhibit preferential adsorption of CO2, we guests (e.g., tetra-n-butylammonium bromide and tetraethy-
have chosen catalytic conversion of CO2 and epoxide into lammonium tetrafluoroborate).6 The cationic tetra-n-butylam-
cyclic carbonates as a model reaction to demonstrate the monium ion is effectively trapped through cation-π interaction
inside the π-electron-rich cavity of the macrocyclic core of the
advantage of the “cavitand-to-framework” design strategy for
POPs (Figure S20). As a result, the nucleophilic attack to the
high catalyst activity and recyclability. The above reaction has
epoxide carbon by the counter bromide anion becomes more
drawn significant attention due to the high atom economy as
facile leading to the increase in the catalytic efficiency.54 We
well as the versatile use of cyclic organic carbonates.44−50 We
protected the −OH groups through postsynthetic modification
optimized the catalytic conversion of CO2 to cyclic carbonates
of RN4-Az-OH with epichlorohydrin to obtain RN4-Az-OR
by taking styrene oxide as the model substrate, RN4-based (Figures S21, S23, and S24).55 It showed only 34% conversion
POPs as catalysts and tetra-n-butylammonium bromide in the same catalytic conditions (Table 2, entry 5). The
(TBAB) as a cocatalyst at 2.5 bar of CO2 pressure at 90 °C reduced efficiency of RN4-Az-OR compared to RN4-Az-OH
under solvent-free conditions. RN4-Az-OH showed excellent and other POPs suggests the importance of −OH groups in
conversion up to 95% within 12 h, while RN4-OH and RN4-F the reaction. RN4-Az-OH also showed excellent catalytic
exhibited the conversion up to 86 and 75% under the similar conversion (84%) for epichlorohydrin under ambient con-
reaction conditions, respectively (Table 2, Figure 4a). ditions like 35 °C and 1 bar of CO2 pressure in 24 h (Table 2,
entry 7).
Table 2. Catalytic Performance of POPs and Cocatalyst for The catalytic conversion of various epoxides to correspond-
the Cycloaddition of CO2 and Styrene Oxidea ing cyclic carbonates depicts the broad applicability of RN4-
Az-OH (Figure 4a). A faster conversion was observed for the
smaller-sized epoxides like propylene oxide and epichlorohy-
drin (Figure 4a). A clear trend based on the molecular size was
observed when the reactions were carried out at a large scale
(20 mmol) with a low cocatalyst amount (TBAB, 0.12 mmol)
entry catalyst conversion (%)b
for the fixed time (12 h). For the smallest substrate, propylene
1 TBAB 25 oxide, the conversion was 98% (Table S3, entry 1). However,
2c RN4-Az-OH 95 for the larger substrates like 1,2-epoxyhexane, the conversion
3c RN4-OH 86 was 65% under identical conditions (Table S3, entry 4). The
4c RN4-F 75 hindrance to diffusion of large-sized substrate molecules
5c RN4-Az-OR 34 through the pores may decrease the efficiency of the reaction
6d Zn/RN4-Az-OH 98
leading to the low conversion. The percentage of conversion of
7e RN4-Az-OH 84
epoxide to the corresponding cyclic carbonate was found to be
8e Zn/RN4-Az-OH 92
92% even in the 10th cycle employing RN4-Az-OH
a demonstrating the reusability of the catalyst (Figure 4c).
Reaction conditions: styrene oxide (10 mmol, 1.2 g), TBAB (0.25
mmol), CO2 (2.5 bar), at 90 °C, and for 12 h. b% of conversion FTIR spectrum of recycled RN4-Az-OH was found to be
calculated through 1H NMR analysis. cCatalyst (POP: 30 mg). dZn/ similar to that of the pristine one (Figure S26). The efficiency
RN4-Az-OH (30 mg), for 6 h. e Catalyst (POP: 30 mg), of RN4-Az-OH was further enhanced through the metalation
epichlorohydrin (2.5 mmol), TBAB (0.5 mmol), for 24 h, CO2 (1 with Zn(OAc)2, (Zn/RN4-Az-OH, Table 2, entries 6 and 8,
bar), at 35 °C. Table S3). The composite material Zn/RN4-Az-OH consisting
8444 DOI: 10.1021/acs.chemmater.9b02563
Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

Figure 4. (a) RN4-Az-OH-catalyzed CO2 fixation with various epoxide derivatives to cyclic organic carbonates. Reaction conditions: epoxide (10
mmol), TBAB (0.25 mmol), CO2 (2.5 bar), at 90 °C; and catalyst (RN4-Az-OH) and 30 mg % of conversion was calculated through 1H NMR
analysis. (b) The relative stability of the dispersion of POPs (1 mg mL−1) in propylene oxide after 12 h. (c) Recyclability of RN4-Az-OH up to ten
cycles. (d) A plausible mechanism of cycloaddition of CO2 with epoxide leading to cyclic carbonate catalyzed by cavitand-based POPs.

of submicron-sized spherical ZnO particles anchored to the example, methylene blue (MB), rhodamine B (RhB), rhod-
polymeric network boosts the activation of epoxides through amine 101 (Rh101), and cresyl violet (CV), and anionic dyes,
Lewis acidic sites along with the hydroxyl-functionalized for example, methyl blue (MeB), Congo red (CR), methyl
framework (Figures S21−S26). orange (MO), and rose bengal (RB), for the study. The
Analyzing the trend of CO2 fixation by RN4-based POPs, we decrease in the absorbance at λmax of the corresponding
have proposed a plausible reaction mechanism (Figure cationic/anionic dyes in the presence of POPs was monitored
4d).56,57 At the first step, the H-bonding between the −OH over time (Figure 5a,b). The RN4-based POPs adsorbed the
groups of the resorcin[4]arene core and the epoxide oxygen cationic dyes even without any mechanical agitation (Figure
takes place.44,56 Subsequently, the nucleophile, Br− of the S27). Considering methylene blue (MB) as the model
cocatalyst TBAB, selectively attacks the less sterically hindered micropollutant,59 we checked the percentage (%) removal
carbon of the epoxide, which leads to the ring opening. This efficiency for all the POPs in 30 min (Figure 5a,c). RN4-based
step is considered to be the rate-determining step of the POPs exhibited much higher dye adsorption capacity
reaction.56 The oxide anion of the open ring epoxide attacks compared to that of the pristine p-hydroxyphenylresorcin[4]-
the carbon atom of CO2 resulting in a carbonate intermediate. arene (Figure 5c). RN4-F was found to be the best adsorbent
The intermediate carbonate can effectively H-bonded with the among the POPs investigated due to the high pore volume, the
−OH groups, and as a consequence, it reduces the energy presence of a large number of electronegative fluorine atoms
barrier of the forward reaction. Finally, the intramolecular ring (facilitating fluorine−cation interaction), as well as the
closure leads to the cyclic carbonate as a product, and the hydrophobic nature of the polymer (Figure 5c, Figure
regeneration of the cocatalyst takes place. In the case of Zn/ S28).26 Albeit of a low surface area, RN4-Az-OH showed
RN4-Az-OH, both the Brønsted acidic phenolic −OH groups better cationic dye adsorption capacity than that of RN4-OH
of the C-phenylresorcin[4]arene core as well as the Lewis due to the presence of a greater number of phenolic −OH
acidic Zn(II) metal centers act synergistically lowering down groups and the high negative zeta potential (Figure 5c).
the activation energy barrier of the forward path. Subsequently, we considered RN4-F as the model adsorbent
Removal of Organic Micropollutants from Water. to demonstrate the selective dye uptake. RN4-F effectively
Resorcin[n]arenes-based cavitands are well-known hosts for a adsorbed unipositive dyes, for example, cresyl violet, methylene
large number of organic molecules.58 We have noticed that the blue, rhodamine B, and rhodamine 101 with removal
POPs developed in the present study have a strong affinity efficiencies of 99.0 ± 0.1, 95.4 ± 4.1, 90.9 ± 3.5, and 96.8 ±
toward the cationic quaternary ammonium salt. The size 2.2%, respectively (Figure 5d,e, Figure S29). On the other
selectivity in the CO2 conversion by the RN4-based POPs has hand, neutral molecules like p-nitrophenol and 2-naphthol (0.1
also been demonstrated (vide supra). These observations mM) were removed up to 80 and 76% by RN4-F, respectively
inspired us to investigate their potential for the charge-specific (1 mg mL−1, Figure S30). To explain the fast adsorption
size-selective uptake of organic micropollutants from water. kinetics of the cationic species, we estimated the sizes of the
We employed a large variety of organic dyes based on solubility micropollutants by the density functional theory (DFT)
in water, different sizes, and charges including cationic dyes, for calculations.60 The molecular dimension was measured from
8445 DOI: 10.1021/acs.chemmater.9b02563
Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

Figure 5. UV−vis absorption spectra of (a) cationic dye, methylene blue (MB, 0.1 mM in water), (b) anionic dye, Congo red (CR, 0.1 mM in
water) at different time intervals in presence of RN4-F (1 mg mL−1); inset: digital photographs of pristine dye solution (left) and after soaking with
RN4-F (right). (c) The percentage removal efficiency of MB in the presence of RN4-F, RN4-Az-OH, RN4-OH, and p-hydroxyphenylresorcin-
[4]arene (HRN4) monomers in 30 min; and (d) the percentage removal efficiency of various cationic dyes: methylene blue (MB), rhodamine B
(RhB), rhodamine 101 (Rh101), cresyl violet (CV), and anionic dyes: methyl blue (MeB), Congo red (CR), methyl orange (MO), and rose
Bengal (RB) employing RN4-F for 30 min. (e) The time-dependent adsorption of aqueous solutions of CV, MB, RhB, and Rh101 (0.1 mM) by
RN4-F (1 mg mL−1). The error bars in (c), (d), and (e) represent the standard deviation of three independent measurements. (f) The pore size
distribution obtained from NLDFT analysis of BET adsorption isotherms of RN4-F before and after adsorption with MB; inset: the calculated size
of MB molecule in X and Y directions (using density functional theory (DFT) and Material Studio version 6.1, space-filled model) depicting close
similarity with pore size distribution of pristine RN4-F. (g) UV−vis absorption spectra of 1:1 (v/v) mixed dye solution (MB: cationic dye, 0.1 mM
and MO: anionic dye, 0.1 mM) at different time intervals in the presence of RN4-F (1 mg mL−1); inset: digital photographs further suggesting the
charge-specific dye separation.

the geometry-optimized structure using Materials Studio RB (1.31 × 1.33 nm) as that of the pores. Dyes having similar
version 6.1 (Figure S31).61 The computed molecular molecular dimensions but opposite charges, for example,
dimensions of CV, MB, RhB, Rh 101, p-nitrophenol, and 2- cationic CV, MB, and anionic MO, showed different uptake by
naphthol are 1.48 × 0.83, 1.55 × 0.73, 1.50 × 1.44, 1.54 × RN4-F. Nearly 99% CV and 96% MB were sequestered by
1.33, 0.79 × 0.55, and 0.82 × 0.62 nm, respectively. Thus, the RN4-F, whereas only 33% of MO was sequestered by RN4-F.
pores around 0.7 and 1.4 nm (arising from the cavity of HRN4, This finding indicates that a positive charge on the guest assists
Figure S12),32 as revealed through the pore size distribution of predominantly in the selective uptake when molecular size is
RN4-F, are easily accessible to the dyes leading to the excellent comparable to the dimension of the pore. We further carried
uptake capacity (Figure 5d,f). The complete disappearance of out the dye adsorption measurement employing a 1:1 (v/v)
the peaks at 0.7 and 1.4 nm of MB-adsorbed RN4-F indicates mixed dye solution (cationic dye methylene blue: 0.1 mM,
that the intrinsic pores are occupied by the dye molecules of anionic dye methyl orange: 0.1 mM). A clear visual color
similar sizes (Figure 5f). CV being the smallest among the change from green to orange was observed within 15 min,
cationic dyes was adsorbed much faster than RhB and MB indicating the selective adsorption of the cationic dye (Figure
(Figure 5e). 55g). As mentioned earlier, efficient fluorine−cation inter-
A negligible uptake of 13.9 ± 0.4 and 17.0 ± 1.1% was actions in RN4-F induces charge-specific dye adsorption.26
observed for the anionic dyes with multiple charges, like We performed the batch-adsorption kinetic measurements
methyl blue (MeB) and Congo red (CR), respectively. The to investigate the mechanistic insight of dye adsorption by
larger sizes of MeB (2.41 × 1.76 nm) and CR (2.74 × 0.87 RN4-F. The experimental data were fitted quite well with the
nm) than that of the pores of RN4-F are responsible for the pseudo-second-order model (MB: R2 ∼0.99; RhB: R2 ∼0.98,
poor uptake capacity (Figure 5b,d). On the other hand, anionic Figures S32 and S33).62−64 The kinetic model was further
dyes with a unit charge, for example, methyl orange (MO) and substantiated by varying the initial dye concentrations (Figure
rose bengal (RB), were adsorbed moderately up to 33.2 ± 1.5 S34a, Table S4) as well as the adsorbent amounts (Figure
and 38.4 ± 0.6 %, respectively (Figure 5d). This can be S34b).62 The thermodynamic studies indicated that the dye
attributed to the comparable size of MO (1.59 × 0.55 nm) and adsorption by RN4-F followed the Langmuir model (Figure
8446 DOI: 10.1021/acs.chemmater.9b02563
Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

S35).18,63−65 The sorption process is likely to take place isotherms. The pore size distribution was obtained through nonlocal
through the energetically equivalent adsorption sites with the density functional theory (NLDFT) analysis. A Cary 100
monolayer formation of the adsorbate (referring to 1:1 spectrophotometer was employed for recording UV−vis absorption
complexation between the dye and RN4 building block, spectra.
Synthesis of Monomers. p-Hydroxyphenylresorcin[4]arene
Figure S35).18,65 The regeneration of RN4-F can easily be (HRN4) was synthesized following the procedure given below.
carried out by simple treatment with ethanol. The POP can be Resorcinol (4.1 mmol, 1 equiv) and concentrated HCl (3.5 mL) were
recycled multiple times with high retention efficiency (Figures dissolved in anhydrous ethanol (25 mL) and stirred at 0 °C for a few
S36 and S37). The selectivity, efficiency, and ease of recovery minutes. An ethanolic solution of p-hydroxybenzaldehyde (4.1 mmol,
suggest the advantage of knitting the macrocyclic building 1 equiv) was added dropwise to the cold solution and refluxed (85
blocks in POPs for water purification. °C) for 12 h. After cooling down to room temperature, the pink color


precipitate was filtered out and purified via reprecipitation in
deionized water. The reprecipitated purple-colored solid was washed
CONCLUSIONS with methanol, ethanol, acetone, and diethyl ether and dried under
Taking C-phenylresorcin[4]arene as a model zero-dimensional vacuum for 12 h at 50 °C (yield: 45%). Following the same protocol,
molecular solid, we demonstrated that connecting the 0D p-bromophenylresorcin[4]arene was synthesized using resorcinol and
porous cores through aromatic linkers led to the 3D-POPs p-bromobenzaldehyde (yield: 42%).
with the drastic improvement in the specific surface area from Synthesis of RN4-Az-OH. Benzidine (0.7 mmol, 2 equiv) was
156 to 1230 m2 g−1. The inherent guest-responsive properties taken in a 100 mL round bottom flask and was charged with 30 mL of
deionized water and 2 mL of concentrated HCl. The solution was
of the 0D preporous building blocks were not only retained but stirred for 30 min in an ice bath. Then, 20 mL of aqueous sodium
also enhanced many folds in the 3D porous polymers. We nitrite (1.4 mmol, 4 equiv) was added dropwise and stirred for 30 min
developed hierarchically mesoporous to ultramicroporous maintaining the temperature at 2−3 °C. The solution was neutralized
POPs by tuning the length and functionality of aromatic with dilute sodium carbonate solution. Then, it was added to the 30
linkers. Three environmentally relevant issues, namely, mL aqueous solution of p-hydroxyphenylresorcin[4]arene (0.35
selective gas uptake, catalytic conversion of CO2 to value- mmol, 1 equiv) and sodium carbonate (4.2 mmol, 12 equiv) and
added chemicals, and removal of micropollutants from water, stirred for 12 h at 2−3 °C. The solid sample was separated from the
were demonstrated to substantiate the advantage of reaction mixture by filtration and washed with dilute HCl solution.
interconnected pores originating through cross-linking of The dark grey-colored solid was thoroughly washed by multiple
Soxhlet extractions with methanol, THF, and ethanol, respectively,
macrocyclic cavitands. Azo-linked RN4-Az-OH was found to followed by freeze drying (yield: 90−95%).
be a remarkable metal-free recyclable catalyst for the Synthesis of RN4-OH. In a typical procedure, p-
cycloaddition of CO2 with several substituted epoxides under bromophenylresorcin[4]arene (0.18 mmol, 1 equiv), copper(I)
solvent-free reaction conditions (Table S7). Alkyne-linked iodide, and bis(triphenylphosphine)palladium(II) chloride were
microporous POP, RN4-OH having a high specific surface area taken in 5 mL degassed THF/DMF (1:1). Diisopropylamine (0.72
(720 m2 g−1) and residual palladium, showed a considerable mmol, 4 equiv) and 1,4-diethynylbenzene were dissolved in a 5 mL
H2 uptake (10 mmol g−1) at 1 atm (Table S6). Fluorinated dry, degassed THF. The solution containing the ethynyl linker (0.36
small linkers led to hydrophobic ultramicroporous POP, RN4- mmol, 2 equiv) was added dropwise to the solution containing the p-
F (1230 m2 g−1, Table S5), a highly robust and efficient bromophenylresorcin[4]arene with stirring to afford a dark gray
reaction mixture. The mixture was heated at 65 °C under N2 for 48 h.
material for charge-specific size-selective separation of organic
After cooling to room temperature, the reaction mixture was
dyes from water (Table S8). Thus, the approach of quenched with cold acidified methanol. The gravimetric filtration
“connecting the dots” paves the way for the development of was used to collect the precipitate. Then, it was washed with THF and
multifunctional macrocyclic cavitand-based 3D-POPs for methanol and followed by thorough washing by Soxhlet extraction for
targeting several environmental remediation problems like 24 h each in methanol, THF, chloroform, and ethanol. The resulting
catalytic conversion of greenhouse gases, selective separation of yellowish-brown solid was dried under vacuum at 80 °C for 12 h
micropollutants from a complex aqueous mixture, and water (yield ∼50%).
desalination. Synthesis of RN4-F. p-Hydroxyphenylresorcin[4]arene (0.35


mmol, 1 equiv) and potassium carbonate (2.8 mmol, 8 equiv) were
taken to a Schlenk tube, and 1.5 mL of dry DMF was added. A
EXPERIMENTAL SECTION solution of tetrafluoroterephthalonitrile (1 mmol, 3 equiv) in
General Methods. Solution-state 1H and 13C NMR spectra were anhydrous THF (13 mL) was added to the above mixture dropwise
acquired from Bruker Avance III 500 MHz NMR spectrometers. The while stirring under N2 atmosphere. The stirring was continued for 48
solid-state 13C NMR (CP/MAS) experiments were carried out on a h at 85 °C. Then, the light brown suspension was cooled. The residual
JEOL ECX2 400 MHz (field 9.4T, 4 mm probe) standard bore K2CO3 was quenched by washing with 1 N HCl until CO2 evolution
spectrometer. Infrared spectra were obtained in KBr discs on a stopped. The solid was filtered followed by soaking for 15 min each in
PerkinElmer model 2000 FTIR spectrometer in the 400−4000 cm−1 H2O, THF, and DCM. Then, the POP was dried through freeze
region. Thermogravimetric analysis (TGA) was carried out in a drying (yield: 60−70%).
nitrogen stream using a PerkinElmer TGA-6000 instrument at a Cycloaddition of CO2 with Epoxides. The reactions were
heating rate of 10 °C min−1. The powder X-ray diffraction (PXRD) carried out in a 35 mL Schlenk-sealed tube with a magnetic stirrer. At
data were recorded on a PANalytical Empyrean XRD instrument. The first, the catalyst (30 mg), tetra-n-butylammonium bromide (0.25
surface morphologies of the POPs were examined using a Carl Zeiss mmol), and epoxide (10 mmol) were added into a reactor. Then, the
(Ultraplus) field emission scanning electron microscope (FESEM). reactor was charged with CO2 (2.5 bar) and kept at 0 °C for 10 min.
An Oxford Instruments X-MaxN spectrometer attached to FESEM The CO2 pressure was maintained, and the reaction tube was placed
was used for energy dispersive X-ray spectroscopy (EDS). The to a preheated oil bath at 90 °C and stirred for the desired time. After
morphology of all the POPs was further studied using a FEI TALOS the reaction, the reactor was cooled to room temperature and the
200S instrument at a working voltage of 200 kV. A Quantachrome pressure was released. An aliquot of the reaction mixture was
Autosorb QUA211011 equipment was employed for the gas sorption dissolved in CDCl3 for the 1H NMR analysis to determine the
studies. POP samples were degassed at 60−80 °C for 12 h before the percentage (%) of conversion (Figures S54−S68). The reaction at 35
measurements. ASIQwin software was used for the analyses of °C and a 1 bar pressure was carried out using CO2 balloon pressure.

8447 DOI: 10.1021/acs.chemmater.9b02563


Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

The catalyst was separated through centrifugation, and the isolated thanks Ms. Parmeet Kaur Dhindsa (BS-MS student, Chemical
yield was estimated after silica gel column purification. RN4-Az-OH Science) of IISER Mohali for a preliminary exploration of dye
was recycled for 10 times. After each cycle, the catalyst was washed adsorption studies. A.G. and B.S. thank IISERB for fellowship.
with dichloromethane and dried under vacuum at 70 °C for 12 h.
Selective Micropollutant Separation. The batch adsorption
A.P. thanks Dr. J. A. Mondal, BARC, Dr. Jyotishman Dasgupta,
experiments of all the dyes were carried out at room temperature TIFR, and Kaleidoscope symposium 2018 for fruitful
under the dark conditions. The initial concentration of dyes was 0.1 discussion.


mM. Ten milligrams of POP was added into 10 mL of the dye
solution. First, the mixture was sonicated for 10 s in order to get a REFERENCES
homogeneous dispersion, and then, it was stirred at 800 rpm. The
mixture was filtered in different time intervals through Whatman 0.45 (1) Rouquerol, J.; Avnir, D.; Fairbridge, C. W.; Everett, D. H.;
μm membrane filters. The concentration of the dye in the filtrate was Haynes, J. H.; Pernicone, N.; Ramsay, J. D. F.; Sing, K. S. W.; Unger,
measured using the UV−vis spectroscopy. RN4-F was reused multiple K. K. Recommendations for the Characterization of Porous Solids
times for the dye uptake. It was regenerated by stirring the POP in (Technical Report). Pure Appl. Chem. 1994, 66, 1739−1758.
ethanol (15 mL) at 50 °C for 6 h and was recovered by filtration. (2) Kane, C. M.; Ugono, O.; Barbour, L. J.; Holman, K. T. Many
Computational Details. The ground-state geometries of all the Simple Molecular Cavitands Are Intrinsically Porous (Zero-Dimen-
dyes were optimized using density functional theory (DFT) at the sional Pore) Materials. Chem. Mater. 2015, 27, 7337−7354.
B3LYP/6-31G (d,p) level.60 The polarizable continuum model (3) Uekama, K.; Hirayama, F.; Irie, T. Cyclodextrin Drug Carrier
(PCM) was utilized to obtain the geometry of the dyes in the Systems. Chem. Rev. 1998, 98, 2045−2076.
aqueous medium. Material Studio version 6.1 was used to estimate (4) Silvi, S.; Arduini, A.; Pochini, A.; Secchi, A.; Tomasulo, M.;
the molecular dimensions from the geometry-optimized structure.61 Raymo, F. M.; Baroncini, M.; Credi, A. A Simple Molecular Machine


*
ASSOCIATED CONTENT
S Supporting Information
Operated by Photoinduced Proton Transfer. J. Am. Chem. Soc. 2007,
129, 13378−13379.
(5) Ogoshi, T.; Yamagishi, T. A.; Nakamoto, Y. Pillar-Shaped
Macrocyclic Hosts Pillar[n]arenes: New Key Players for Supra-
The Supporting Information is available free of charge on the
molecular Chemistry. Chem. Rev. 2016, 116, 7937−8002.
ACS Publications website at DOI: 10.1021/acs.chemma- (6) Zhang, Q.; Catti, L.; Tiefenbacher, K. Catalysis Inside the
ter.9b02563. Hexameric Resorcinarene Capsule. Acc. Chem. Res. 2018, 51, 2107−
Design strategy, synthesis, and full characterizations of 2114.
POPs through PXRD, FTIR, NMR, XPS spectra, (7) McKeown, N. B.; Budd, P. M. Polymers of Intrinsic
FESEM, TEM imaging, BET specific surface area, pore Microporosity (PIMs): Organic Materials for Membrane Separations,
size distributions, and selective gas uptake; CO 2 Heterogeneous Catalysis and Hydrogen Storage. Chem. Soc. Rev.
conversion, stabilities, and recyclabilities of catalysts; 2006, 35, 675−683.
(8) Thomas, A. Functional Materials: From Hard to Soft Porous
charge-specific size-selective dye adsorption from water,
Frameworks. Angew. Chem., Int. Ed. 2010, 49, 8328−8344.
kinetics, and thermodynamics of dye removal; and (9) Patra, A.; Scherf, U. Fluorescent Microporous Organic Polymers:
comparative accounts of macrocyclic cavitand-based Potential Testbed for Optical Applications. Chem. − Eur. J. 2012, 18,
POPs for task-specific applications with other reported 10074−10080.
porous materials (PDF) (10) Xu, Y.; Jin, S.; Xu, H.; Nagai, A.; Jiang, D. Conjugated

■ AUTHOR INFORMATION
Corresponding Author
Microporous Polymers: Design, Synthesis and Application. Chem. Soc.
Rev. 2013, 42, 8012−8031.
(11) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The
Chemistry and Applications of Metal-Organic Frameworks. Science
*E-mail: [email protected]. 2013, 341, 1230444.
ORCID (12) Slater, A. G.; Cooper, A. I. Function-led Design of New Porous
Arkaprabha Giri: 0000-0003-4356-2014 Materials. Science 2015, 348, aaa8075.
MD. Waseem Hussain: 0000-0001-6451-0964 (13) Wang, Z. J.; Ghasimi, S.; Landfester, K.; Zhang, K. A. I.
Molecular Structural Design of Conjugated Microporous Poly-
Bahadur Sk: 0000-0002-7098-5429 (Benzooxadiazole) Networks for Enhanced Photocatalytic Activity
Abhijit Patra: 0000-0003-3144-1813 with Visible Light. Adv. Mater. 2015, 27, 6265−6270.
Author Contributions (14) Bandyopadhyay, S.; Kundu, S.; Giri, A.; Patra, A. A smart
A.G. carried out the fabrication, characterizations, and analyses Photosensitizer Based on a Red Emitting Solution Processable Porous
of the materials developed in the study. W.H. helped in the Polymer: Generation of Reactive Oxygen Species. Chem. Commun.
synthesis of RN4-OH and CO2 conversion studies, and he 2018, 54, 9123−9126.
carried out the XPS analysis. B.S. contributed to the (15) Kandambeth, S.; Dey, K.; Banerjee, R. Covalent Organic
computational investigation. A.P. and A.G. wrote the manu- Frameworks: Chemistry beyond the Structure. J. Am. Chem. Soc.
2019, 141, 1807−1822.
script.
(16) Yuan, Y.; Zhu, G. Porous Aromatic Frameworks as a Platform
Funding for Multifunctional Applications. ACS Cent. Sci. 2019, 5, 409−418.
DST/CHM/2018/086 (DST/TM/WTI/WIC/2K17/82(G)) (17) Wu, J.; Xu, F.; Li, S.; Ma, P.; Zhang, X.; Liu, Q.; Fu, R.; Wu, D.
and SERB/CHM/2017/113 (file no. EMR/2017/000233). Porous Polymers as Multifunctional Material Platforms toward Task-
Notes Specific Applications. Adv. Mater. 2019, 31, 1802922.
The authors declare no competing financial interest. (18) Alsbaiee, A.; Smith, B. J.; Xiao, L.; Ling, Y.; Helbling, D. E.;


Dichtel, W. R. Rapid Removal of Organic Micropollutants from Water
by a Porous β-Cyclodextrin Polymer. Nature 2016, 529, 190−194.
ACKNOWLEDGMENTS
(19) Li, H.; Meng, B.; Chai, S.-H.; Liu, H.; Dai, S. Hyper-crosslinked
Financial support from DST-SERB, infrastructural support β-Cyclodextrin Porous Polymer: An Adsorption-facilitated Molecular
from IISERB, and the FIST supported TEM facility to the Catalyst Support for Transformation of Water-soluble Aromatic
Dept. of Chemistry, IISERB are gratefully acknowledged. A.G. Molecules. Chem. Sci. 2016, 7, 905−909.

8448 DOI: 10.1021/acs.chemmater.9b02563


Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

(20) Talapaneni, S. N.; Kim, D.; Barin, G.; Buyukcakir, O.; Je, S. H.; Blocks for Gas Storage and Separation. Chem. Mater. 2014, 26, 4023−
Coskun, A. Pillar[5]arene Based Conjugated Microporous Polymers 4029.
for Propane/Methane Separation through Host-Guest Complexation. (39) Hussain, W.; Bandyopadhyay, S.; Patra, A. Microporous
Chem. Mater. 2016, 28, 4460−4466. Organic Polymers Involving Thiadiazolopyridine for High and
(21) Shetty, D.; Jahovic, I.; Raya, J.; Ravaux, F.; Jouiad, M.; Olsen, J.- Selective Uptake of Greenhouse Gases at Low Pressure. Chem.
C.; Trabolsi, A. An Ultra-absorbent Alkyne-rich Porous Covalent Commun. 2017, 53, 10576−10579.
Polycalix[4]arene for Water Purification. J. Mater. Chem. A 2017, 5, (40) Ansón, A.; Lafuente, E.; Urriolabeitia, E.; Navarro, R.; Benito,
62−66. A. M.; Maser, W. K.; Martínez, M. T. Hydrogen Capacity of
(22) Su, K.; Wang, W.; Li, B.; Yuan, D. Azo-Bridged Calix[4]- Palladium-Loaded Carbon Materials. J. Phys. Chem. B 2006, 110,
resorcinarene-Based Porous Organic Frameworks with Highly 6643−6648.
Efficient Enrichment of Volatile Iodine. ACS Sustainable Chem. Eng. (41) Yao, B.-J.; Li, J.-T.; Huang, N.; Kan, J.-L.; Qiao, L.; Ding, L.-G.;
2018, 6, 17402−17409. Li, F.; Dong, Y.-B. Pd NP-Loaded and Covalently Cross-Linked COF
(23) Jin, Y.; Voss, B. A.; McCaffrey, R.; Baggett, C. T.; Noble, R. D.; Membrane Microreactor for Aqueous CBs Dechlorination at Room
Zhang, W. Microwave-assisted Syntheses of Highly CO2-selective Temperature. ACS Appl. Mater. Interfaces 2018, 10, 20448−20457.
Organic Cage Frameworks (OCFs). Chem. Sci. 2012, 3, 874−877. (42) Ramamurthy, V.; Sivaguru, J. Supramolecular Photochemistry
(24) Yang, H.; Du, Y.; Wan, S.; Trahan, G. D.; Jin, Y.; Zhang, W. as a Potential Synthetic Tool: Photocycloaddition. Chem. Rev. 2016,
Mesoporous 2D Covalent Organic Frameworks based on Shape- 116, 9914−9993.
persistent Arylene-ethynylene Macrocycles. Chem. Sci. 2015, 6, 4049− (43) Mandal, A. K.; Suresh, M.; Das, P.; Das, A. Restricted
4053. Conformational Flexibility of a Triphenylamine Derivative on the
(25) Du, Y.; Yang, H.; Whiteley, J. M.; Wan, S.; Jin, Y.; Lee, S.-H.; Formation of Host-Guest Complexes with Various Macrocyclic Hosts.
Zhang, W. Ionic Covalent Organic Frameworks with Spiroborate Chem. − Eur. J. 2012, 18, 3906−3917.
Linkage. Angew. Chem., Int. Ed. 2016, 55, 1737−1741. (44) Martínez-Rodríguez, L.; Garmilla, J. O.; Kleij, A. W. Cavitand-
(26) Byun, J.; Patel, H. A.; Thirion, D.; Yavuz, C. T. Charge-Specific Based Polyphenols as Highly Reactive Organocatalysts for the
Size-Dependent Separation of Water-Soluble Organic Molecules by Coupling of Carbon Dioxide and Oxiranes. ChemSusChem 2016, 9,
Fluorinated Nanoporous Networks. Nat. Commun. 2016, 7, 13377. 749−755.
(27) Tunstad, L. M.; Tucker, J. A.; Dalcanale, E.; Weiser, J.; Bryant, (45) Sakakura, T.; Kohno, K. The Synthesis of Organic Carbonates
J. A.; Sherman, J. C.; Helgeson, R. C.; Knobler, C. B.; Cram, D. J. from Carbon Dioxide. Chem. Commun. 2009, 1312−1330.
Host-Guest Complexation. 48. Octol Building Blocks for Cavitands (46) Martín, C.; Fiorani, G.; Kleij, A. W. Recent Advances in the
and Carcerands. J. Org. Chem. 1989, 54, 1305−1312. Catalytic Preparation of Cyclic Organic Carbonates. ACS Catal. 2015,
(28) Ji, G.; Yang, Z.; Zhang, H.; Zhao, Y.; Yu, B.; Ma, Z.; Liu, Z. 5, 1353−1370.
Hierarchically Mesoporous o-Hydroxyazobenzene Polymers: Syn- (47) Du, Y.; Yang, H.; Wan, S.; Jin, Y.; Zhang, W. A titanium-based
thesis and Their Applications in CO2 Capture and Conversion.
porous coordination polymer as a catalyst for chemical fixation of
Angew. Chem., Int. Ed. 2016, 55, 9685−9689.
CO2. J. Mater. Chem. A 2017, 5, 9163−9168.
(29) Trewin, A.; Cooper, A. I. Porous Organic Polymers: Distinction
(48) Buyukcakir, O.; Je, S. H.; Talapaneni, S. N.; Kim, D.; Coskun,
from Disorder? Angew. Chem., Int. Ed. 2010, 49, 1533−1535.
A. Charged Covalent Triazine Frameworks for CO2 Capture and
(30) Thommes, M.; Kaneko, K.; Neimark, A. V.; Olivier, J. P.;
Conversion. ACS Appl. Mater. Interfaces 2017, 9, 7209−7216.
Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K. S. W. Physisorption of
(49) Huang, K.; Zhang, J.-Y.; Liu, F.; Dai, S. Synthesis of Porous
Gases, with Special Reference to the Evaluation of Surface Area and
Polymeric Catalysts for the Conversion of Carbon Dioxide. ACS
Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem.
2015, 87, 1051−1069. Catal. 2018, 8, 9079−9102.
(31) Cychosz, K. A.; Guillet-Nicolas, R.; García-Martínez, J.; (50) Lyu, J.; Zhang, X.; Li, P.; Wang, X.; Buru, C. T.; Bai, P.; Guo,
Thommes, M. Recent Advances in the Textural Characterization of X.; Farha, O. K. Exploring the Role of Hexanuclear Clusters as Lewis
Hierarchically Structured Nanoporous Materials. Chem. Soc. Rev. Acidic Sites in Isostructural Metal−Organic Frameworks. Chem.
2017, 46, 389−414. Mater. 2019, 31, 4166−4172.
(32) Pei, W.-Y.; Xu, G.; Yang, J.; Wu, H.; Chen, B.; Zhou, W.; Ma, (51) Parlett, C. M. A.; Wilson, K.; Lee, A. F. Hierarchical Porous
J.-F. Versatile Assembly of Metal-Coordinated Calix[4]resorcinarene Materials: Catalytic Applications. Chem. Soc. Rev. 2013, 42, 3876−
Cavitands and Cages through Ancillary Linker Tuning. J. Am. Chem. 3893.
Soc. 2017, 139, 7648−7656. (52) Urakami, H.; Zhang, K.; Vilela, F. Modification of Conjugated
(33) Dawson, R.; Stöckel, E.; Holst, J. R.; Adams, D. J.; Cooper, A. I. Microporous Poly-benzothiadiazole for Photosensitized Singlet Oxy-
Microporous Organic Polymers for Carbon Dioxide Capture. Energy gen Generation in Water. Chem. Commun. 2013, 49, 2353−2355.
Environ. Sci. 2011, 4, 4239−4245. (53) Mai, W.; Sun, B.; Chen, L.; Xu, F.; Liu, H.; Liang, Y.; Fu, R.;
(34) Yassin, A.; Trunk, M.; Czerny, F.; Fayon, P.; Trewin, A.; Wu, D.; Matyjaszewski, K. Water-Dispersible, Responsive, and
Schmidt, J.; Thomas, A. Structure−Thermodynamic-Property Rela- Carbonizable Hairy Microporous Polymeric Nanospheres. J. Am.
tionships in Cyanovinyl-Based Microporous Polymer Networks for Chem. Soc. 2015, 137, 13256−13259.
the Future Design of Advanced Carbon Capture Materials. Adv. Funct. (54) Avram, L.; Cohen, Y. Spontaneous Formation of Hexameric
Mater. 2017, 27, 1700233. Resorcinarene Capsule in Chloroform Solution as Detected by
(35) Wang, W.; Zhou, M.; Yuan, D. Carbon Dioxide Capture in Diffusion NMR. J. Am. Chem. Soc. 2002, 124, 15148−15149.
Amorphous Porous Organic Polymers. J. Mater. Chem. A 2017, 5, (55) Lu, W.; Bosch, M.; Yuan, D.; Zhou, H.-C. Cost-Effective
1334−1347. Synthesis of Amine-Tethered Porous Materials for Carbon Capture.
(36) Patel, H. A.; Je, S. H.; Park, J.; Chen, D. P.; Jung, Y.; Yavuz, C. ChemSusChem 2015, 8, 433−438.
T.; Coskun, A. Unprecedented High-Temperature CO2 Selectivity in (56) Alves, M.; Mereau, R.; Grignard, B.; Detrembleur, C.; Jerome,
N2-phobic Nanoporous Covalent Organic Polymers. Nat. Commun. C.; Tassaing, T. A Comprehensive Density Functional Theory Study
2013, 4, 1357. of the Key Role of Fluorination and Dual Hydrogen Bonding in the
(37) Bandyopadhyay, S.; Anil, A. G.; James, A.; Patra, A. Activation of the Epoxide/CO2 Coupling by Fluorinated Alcohols.
Multifunctional Porous Organic Polymers: Tuning of Porosity, RSC Adv. 2016, 6, 36327−36335.
CO2, and H2 Storage and Visible-Light-Driven Photocatalysis. ACS (57) Zhang, D.-D.; Feng, X.; Gnanou, Y.; Huang, K.-W. Theoretical
Appl. Mater. Interfaces 2016, 8, 27669−27678. Mechanistic Investigation into Metal-Free Alternating Copolymeriza-
(38) Zhang, X.; Lu, J.; Zhang, J. Porosity Enhancement of tion of CO2 and Epoxides: The Key Role of Triethylborane.
Carbazolic Porous Organic Frameworks Using Dendritic Building Macromolecules 2018, 51, 5600−5607.

8449 DOI: 10.1021/acs.chemmater.9b02563


Chem. Mater. 2019, 31, 8440−8450
Chemistry of Materials Article

(58) Kobayashi, K.; Yamanaka, M. Self-assembled Capsules based on


Tetrafunctionalized Calix[4]resorcinarene Cavitands. Chem. Soc. Rev.
2015, 44, 449−466.
(59) He, X.; Male, K. B.; Nesterenko, P. N.; Brabazon, D.; Paull, B.;
Luong, J. H. T. Adsorption and Desorption of Methylene Blue on
Porous Carbon Monoliths and Nanocrystalline Cellulose. ACS Appl.
Mater. Interfaces 2013, 5, 8796−8804.
(60) Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical
Continuum Solvation Models. Chem. Rev. 2005, 105, 2999−3094.
(61) Fan, H.; Gu, J.; Meng, H.; Knebel, A.; Caro, J. High-Flux
Membranes Based on the Covalent Organic Framework COF-LZU1
for Selective Dye Separation by Nanofiltration. Angew. Chem., Int. Ed.
2018, 57, 4083−4087.
(62) Ho, Y. S.; McKay, G. Pseudo-second Order Model for Sorption
Processes. Process Biochem. 1999, 34, 451−465.
(63) Shetty, D.; Jahovic, I.; Raya, J.; Asfari, Z.; Olsen, J.-C.; Trabolsi,
A. Porous Polycalix[4]arenes for Fast and Efficient Removal of
Organic Micropollutants from Water. ACS Appl. Mater. Interfaces
2018, 10, 2976−2981.
(64) Roy, A.; Adhikari, B.; Majumder, S. B. Equilibrium, Kinetic, and
Thermodynamic Studies of Azo Dye Adsorption from Aqueous
Solution by Chemically Modified Lignocellulosic Jute Fiber. Ind. Eng.
Chem. Res. 2013, 52, 6502−6512.
(65) Alzate-Sánchez, D. M.; Smith, B. J.; Alsbaiee, A.; Hinestroza, J.
P.; Dichtel, W. R. Cotton Fabric Functionalized with a β-Cyclodextrin
Polymer Captures Organic Pollutants from Contaminated Air and
Water. Chem. Mater. 2016, 28, 8340−8346.

8450 DOI: 10.1021/acs.chemmater.9b02563


Chem. Mater. 2019, 31, 8440−8450

You might also like