Next Article in Journal
Social Botomics: A Systematic Ensemble ML Approach for Explainable and Multi-Class Bot Detection
Next Article in Special Issue
Adsorption of Methyl Orange on a Novel Palygorskite/UiO-66 Nanocomposite
Previous Article in Journal
Selective Electrochemical Regeneration of Aqueous Amine Solutions to Capture CO2 and to Convert H2S into Hydrogen and Solid Sulfur
Previous Article in Special Issue
Preparation and Application of ZIF-8 Thin Layers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controllable Synthesis of 1, 3, 5-tris (1H-benzo[d]imidazole-2-yl) Benzene-Based MOFs

1
Department of Chemical Engineering, Changwon National University, Changwon-Si 51140, Korea
2
Department of Materials Science and Engineering, Ulsan National Institute of Science and Technology (UNIST), 50 UNIST-gil, Ulju-gun, Ulsan 689-798, Korea
3
Industry-University Cooperation Foundation, Changwon National University, Changwon-Si 51140, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this research work.
Appl. Sci. 2021, 11(21), 9856; https://doi.org/10.3390/app11219856
Submission received: 15 September 2021 / Revised: 13 October 2021 / Accepted: 19 October 2021 / Published: 21 October 2021
(This article belongs to the Special Issue Synthesis and Applications of Metal-Organic Frameworks (MOFs))

Abstract

:
The growing interest in metal–organic frameworks (MOFs) in both industrial and scientific circles has increased in the last twenty years, owing to their crystallinity, structural versatility, and controlled porosity. In this study, we present three novel MOFs obtained from the 1, 3, 5-tris (1H-benzo[d]imidazole-2-yl) benzene (TIBM) organic linker. The formed TIBM crystal powders were characterized by scanning electron microscopy (SEM) to estimate the morphology of the particles, powder X-ray diffraction (XRD) to confirm the crystal structure, Brunauer–Emmett–Teller (BET) method for structural analysis, and thermogravimetric measurements to examine the thermal stability. The TIBM-Cu MOF showed excellent CO2 (3.60 mmol/g) adsorption capacity at 1 bar and 298 K, because of the open Cu site, compared to TIBM-Cr (1.6 mmol/g) and TIBM-Al (2.1 mmol/g). Additionally, due to the high porosity (0.3–1.5 nm), TIBM-Cu MOF showed a considerable CO2/N2 selectivity (53) compared to TIBM-Al (35) and TIBM-Cr (10).

1. Introduction

The design and implementation of new porous materials for carbon dioxide (CO2) separation by selective adsorption is a rapidly increasing research area, because of its importance in energy and environment-related applications [1,2,3]. The greenhouse effect responsible for global warming is mainly related to CO2 emissions. Globally, CO2 release is rapidly increasing because of the combustion of carbon-based fuels (oil, coal, and natural gas) and chemical reactions in petrochemical industries, steel, and cement [4]. Post-combustion CO2 capture is considered the most effective technique for minimizing CO2 emitted from industrial and energy-related sources. To this end, different technologies have been applied for CO2 capture, such as membrane separation [5,6] liquid ammonia, and amine absorption [7,8] and adsorption [9,10]. Membrane separation is not suitable to large-scale applications because of its short lifetime, limited performance at low pressure, and poor stability in acid gas environments [11]. The absorption process has been extensively used in power plants, thanks to the high CO2 selectivity promoted by amine functional groups; however, substantial downsides include equipment corrosion, high energy intake, and toxic ammonia loss [12].
Several materials have been investigated for their promising CO2 capture and catalytic conversion properties, including supported silica [13,14], zeolites [15,16], metal oxides [17,18], bio-waste-derived carbons [19,20], metal–organic frameworks (MOFs) [21,22], porous carbons [23,24], and recently, a new class of porous materials, i.e., porous organic polymers (POPs). The last materials further encompass crystalline covalent organic frameworks (COFs) [25,26], triazine-based organic frameworks (CTFs) [27,28], conjugated microporous polymers (CMPs) [29,30], porous aromatic frameworks (PAFs) [31,32,33], and hyper-cross-linked polymers (HCPs) [34,35]. Amongst the cited systems, MOFs have attracted significant attention, owing to their high surface area, pore volume, and surface functionality. Their unique crystal structure is composed of metal centers connected by organic linkers [36] that can be tuned to give rise to wide applications in gas storage [37,38,39], separation [39,40], drug delivery [41], and catalysis [42].
Generally, MOFs present high CO2 adsorption capacities (2–5 mmol/g) at 298 K and high pressure, while the CO2 capture capacity at below 1 bar is still challenging [42]. Therefore, many methods have been developed to obtain MOFs with improved adsorption capacity, such as metal doping in the metal–organic framework [43] and chemical functionalization with amine groups, such as polyethyleneimine (PEI) and tetraethylenepentamine (TEPA) [44,45,46]. Nevertheless, during the preparation, other aspects must be considered, such as recyclability, stability, and mild regeneration conditions. Researchers have highlighted alternative synthetic methods to modify MOFs stability, flexibility, and particle size [47,48]. The materials were prepared using a solvothermal method, which requires high-temperature treatment of metal salts and organic ligands in water or organic solvents [49]. Solvent properties, such as polarity and hydrophobic/hydrophilicity, affect the reaction mechanism. In particular, low-boiling-point solvents generate high vapor pressure under a reaction in a sealed reactor. The vapor pressure in a sealed reactor increases as the temperature increases, which causes the formation of a supercritical fluid, in which liquid and gas exist simultaneously. This supercritical-fluid phase improves the mixing of chemical reagents in the solvent, and causes product formation by crystallization, which does not normally occur. Hence, the use of low-volatility solvents increases the mobility of the dissolved ions and allows better mixing of the reagents [50,51,52]. The surface functionalities and structural topologies of MOFs are determined by various types of metal ions and the network-forming ligands. Concerning this, alteration in the functionality and structure of MOFs is achieved using different secondary building units (SBUs), specific functional groups, and multi-topic building blocks [53].
Several novel structures have been developed to improve the MOFs performance. Chen et al. proposed an NJU-Bai MOF with amide-functionalized carboxylate linkers and measured a significant improvement in CO2/N2 selectivity (166.7) at 298 K [54]. Wang et al. reported a covalent triazine framework (CTF) with N-heteroaromatic building blocks, which exhibited a CO2 capacity of 3.48 mmol/g at 273 K and 1 bar [55]. Suresh et al. studied an amide-functionalized microporous organic polymer (Am-MOP) with a CO2 capacity of 40 mL/g at 195 K and 1 bar. To improve CO2 binding affinity, Zheng et al. introduced polar acylamide groups in the framework, and the obtained Porous Coordination Network (PCN) series MOF exhibited a CO2 capacity of 20–24 mmol/g at 20 bar and 298 K [56]. Martin Schroder and coworkers prepared a series of isoreticular MFM MOFs functionalized with amide group, among which MFM-126 showed an excellent CO2 adsorption capacity of 4.63 mmol/g at 298 K and 1 bar [56].
In this work, we report the preparation of novel MOFs consisting of 1, 3, 5-tris (1H-benzo[d]imidazole-2-yl) benzene (TIBM) as an organic linker and Al, Cr, and Cu as metal ions. The synthesis of the organic core integrated with imidazole rings, which possess secondary and tertiary amines, was crucial for the protonation of metal ions. In order to allow the correct metal incorporation and obtain high-quality materials, the imidazole rings of the ligand have to be equally distant and symmetrically anchored to the phenyl ring [57,58,59,60,61,62]. Additionally, the use of the TIBM linkers with different metal precursors affects the pore size and the CO2/N2 adsorption selectivity of the prepared system. In this regard, we synthesized three novel MOFs using different metal precursors (MOF-Al, MOF-Cr, and MOF-Cu) for CO2 and N2 capture. The obtained materials were examined using FT-IR, X-ray powder diffraction (XRPD), scanning electron microscopy (SEM), and thermogravimetric analysis (TGA). The volumetric method was used for CO2 and N2 adsorption capacity measurements.

2. Results

2.1. Characterization of the TIBM Based MOFs

2.1.1. Morphological Properties

N2 adsorption isotherms at 77 K (Figure 1a) were measured using a volumetric BET instrument (BELSORP-max, MicrotracBEL, Osaka, Japan) in order to characterize the structural properties of the MOFs. All the isotherms showed a typical Type I isotherm, corresponding to a microporous structure. The specific surface area and pore volume of the TIBM MOFs were determined by––the Brunauer Emmett Teller (BET) method. TIBM-Cr presented the largest surface area and pore volume among the other MOFs (Table 1). The pore size distribution of the TIBM MOFs was determined by the non-local density functional theory (NLDFT) method. TIBM-Cr MOFs showed the largest pore size range (1.0–4.0 nm), compared to TIBM-Cu (0.3 to 1.5 nm) and TIBM-Al (1.0 to 3.0 nm) (Figure S1b in Supporting Information). The majority of pores appeared at 2 nm for TIBM-Al and TIBM-Cr, and at 1 nm for TIBM-Cu.
Figure 2 reports the SEM images of the three TIBM MOF samples. The average particle diameters of the TIBM MOFs were determined using Image J software (NIH, Bethesda, MD, USA) based on the pixel distance of each image, which is correlated with the scale bar. TIBM-Cr exhibited a smaller particle size (0.25 µm) than TIBM-Cu (28.29 µm) and TIBM-Al (0.61 µm).
The MOF TIBM-Al showed a clear hexagonal crystal structure with sharp edges, spherical and highly porous nanoclusters were observed for TIBM-Cr, and a rhombic crystal structure was detected for TIBM-Cu. The TIBM-MOFs surfaces were found to be smooth, without evident cracks due to strong chelation between the central metal ion and TIBM amine linkers.

2.1.2. X-ray Diffraction (XRD) Analysis

The quality of TIBM-MOFs was optimized, and the samples were successfully doped with three metal ions by ionic-covalent bonds. The XRD pattern (Figure 3) of the TIBM-MOFs was compatible with the simulated pattern of UiO-66 from the Cambridge Structural Database (CCDC 837796) [63]. The patterns of TIBM-Cu and TIBM-Al are consistent with those of Cu-BTC and HKUST-1, respectively, while for TIBM-Cr, a resemblance to that of MIL-101 (Fe) and Fe-BTC was found [64,65,66,67,68,69,70,71], thus confirming the successful synthesis and integrity of the crystal structure after the coordination with TIBM (Table 2). In particular, the chelation was indicated by a broad peak and a sharp peak at 2θ values of 9° and 18°, respectively, providing clear evidence of the crystalline nature of the TIBM-Al MOF. The Cr-TIBM MOF exhibited blunt peak intensities at 2θ values of 10° and 18°–20°, showing an increase in the crystalline degree of the MOFs [67,68]. On the other hand, the Cu-TIBM MOF showed sharp crystalline peaks, comparable to those of the Cu-based MOFs [64,65].

2.1.3. FTIR Analysis

The FTIR analysis allowed characterizing molecular interactions and bonding formation in the MOF frameworks (Figure 4). The strong stretching bands at 490–500 cm−1 were attributed to metal-hydrogen bonds, particularly those of Cr and Cu metal ions. The bands at 720–724 cm−1 and 750–754 cm−1 correspond to =C-H bond modes in phenyl rings [74,75]. The characteristic MOFs bands, related to metal-ion-bound second and third amines (>NH-M-N and >N-M-N of MOF), appear in the range 1090–1100 cm−1. Stretching bands due to C=C and C-H deformations of the phenyl rings were observed at 1399 cm−1. The strong vibration mode at 1455 cm−1 related to -NH and metal ions was attributed to the bidentate behavior of the N-M-N moiety. These characteristic peaks match well with the previously reported FTIR analysis of MOF-199 [28]. The strong resonance band exhibited by TIBM-Al was attributed to strong H-bonding of hydroxyl groups in the porous TIBM-Al material, as compared to TIBM-Cr and TIBM-Cu. This strong H-bonding occurs in Al-metal-based MOFs such as MIL-53(Al) and MIL-96 (Al), as compared to the Zn-based ZIF-8 and Zr-based UiO-66, and results in a strong resonance band [76,77,78]. The spectra for TIBM-Cr appear noticeably different after 3500 cm−1, as compared to TIBM-Al and TIBM-Cu. The small peaks that were observed in the TIBM-Al and TIBM-Cu spectra after 3500 cm−1 were attributed to the presence of crystalline water or the acidic -OH in carboxylic groups; these peaks do not appear in the case of TIBM-Cr [79,80]. The FTIR spectrum of the TIBM linker showed the typical peaks for NH wagging at 910 cm1, for C=C at 1420 cm1, for C=N at 1442 cm1, and 1610 cm1, for CO-NH at 1713 cm1, and for NH at 3448 cm1 (Figure 4b).

2.1.4. Thermal Analysis

During the TIBM MOFs thermogravimetric analysis (Figure 5), small drops in the range 50–150 °C were caused by dehydration. A further loss of 48% (300–500 °C) for TIBM-Cu, 28% (500–650 °C) for TIBM-Al, and 40% (400–600 °C) for TIBM-Cr, was ascribed to the decomposition of chelated imidazolium moieties [81,82]. According to the large drop temperatures, TIBM-Al presented better thermal stability than the other TIBM MOFs.

2.2. CO2 and N2 Adsorption Measurements

To examine the CO2-capture performance of all the samples, CO2 and N2 adsorption properties were measured at 298 K and 0–1 bar. As reported in Figure 6, TIBM-Cu showed the highest CO2 adsorption capacity (3.60 mmol/g) at 1 bar, compared to the TIBM-Al (2.04 mmol/g) and TIBM-Cr (1.67 mmol/g). This result is ascribable to both the metal site exposure and the metal oxide chelation of the TIBM host precursor for Cu2+ ions [83]. The Cu-Cu magnetic interaction of TIBM-Cu is stronger than the Al-Al and Cr-Cr magnetic interactions of TIBM-Al and TIBM-Cr, respectively, and results in a strong interaction between CO2 and the two available electrons of the Cu2+ metal in the TIBM-Cu MOF [84,85]. Conversely, TIBM-Cr showed the highest surface area (2141.1 m2/g) and total pore volume 2.116 cm3/g, compared to the other examined MOFs (Table 1), while a reverse adsorption trend was found for N2 adsorption for the TIBM MOFs. With regards to the CO2/N2 adsorption selectivity, which is defined as the ratio between CO2 adsorption capacity at 0.15 bar and N2 adsorption capacity at 0.85 bar, TIBM-Cu showed the highest value (53.69) compared to TIBM-Cr (11.16) and TIBM-Al (33.33), as reported in Figure 7, mainly in virtue of its high CO2 adsorption capacity, and concurrent low N2 adsorption capacity. In addition, CO2/N2 selectivity is different for each of the TIBM MOFs and depends on different factors, such as MOF pore size and available open metal sites for CO2. The pore sizes of TIBM-Cu (0.5 and 1 nm) are more suited to selective adsorption of CO2 (0.33 nm) over N2 (0.35 nm) than TIMB-Al (1.9 nm) and TIBM-Cr (2.2 nm), owing to the size-selective separation. The CO2/N2 adsorption selectivity of TIBM-Cu is promising, compared to other MOFs reported in the literature (e.g., 6 for CuDABCO [86], 8 for ZIF-8 [87], 12 for MIL-101 (Cr) [88], 18 for MOF-5 [89], 16.5 for MOF-177 [89]). However, the adsorption capacities of the TIBM MOFs are comparable to those of previously reported MOFs (Table 3). TIBM-Cu was selected for the performance evaluation, showing that the CO2 adsorption capacity (Figure 8) gradually decreased with increasing temperature from 298 to 338 K.

3. Materials and Methods

3.1. Chemicals

All chemicals were used as received, without further purification. Trimesic acid, chromium nitrate nonahydrate (99%), hydrofluoric acid (ACS reagent ≥ 48%), deionized water, copper acetate (trace metals basis ≥ 99.99%), ethanol (absolute alcohol ≥ 99.8%), aluminum chloride hexahydrate (reagent plus ≥ 99%), zinc acetate dihydrate (ACS reagent ≥ 98%), N,N-dimethylformamide (anhydrous, ≥ 99.8%), polyphosphoric acid (reagent grade), o-phenylenediamine (≥ 98%), sodium bicarbonate solution (Bioreagent 7.5%), and methanol (ACS reagent) were purchased from Sigma Aldrich (St. Louis, MO, USA).

3.2. Preparation of 1, 3, 5-tris (1H-benzo[d]imidazole-2-yl) Benzene

O-phenylenediamine (7.2 g, 0.06 mol) was added to a solution of trimesic acid (4 g, 0.03 mmol) in polyphosphoric acid (PPA) (50 mL), and the reaction mixture was heated at 230 °C for 12 h (Figure 9). The resultant yellowish-colored reaction mixture was poured into ice-cold water (500 mL); upon stirring, the obtained brown precipitate was collected. The precipitate was neutralized by adding 20% sodium bicarbonate solution and filtered by centrifugation (4000 rpm). The brown solid converted into a white solid (82% yield) after recrystallization with methanol [110]; Mp. 280 °C. 1H NMR (400 MHz, DMSO-d6/TMS, ppm) δ 7.24–7.52 (m, 6H), δ 7.62–7.92 (m, 6H), 8.9 (s, 3H), 13.2 (s, 3H); 13C (100 MHz, DMSO-d6/TMS, ppm) δ 115.2, 119.2, 122.2, 123.3, 125.5, 131.6, 135.8, 144.2, 159.1; ESI-MS: ([M + H]+) m/z 427.18 (100%), found m/z 427.32 (100%), Chemical Formula: C27H19N6+, (Figure S1). The TIBM was synthesized as described in the literature [110]; the obtained pale yellow solid presented FT-IR and 1H-NMR spectra compatible with the ones previously reported [110,111].

3.3. Synthesis of TIBM-Cr

Chromium nitrate nonahydrate (0.6340 g) and TIBM (0.438 g) were dissolved in 20 mL deionized water by sonication for 10 min, and hydrofluoric acid (HF) (60 µL) was added to the mixture. The reaction mixture was transferred into a Teflon autoclave reactor sealed in a stainless steel vessel and maintained at 483 K for 48 h. The fine green-colored precipitate was washed three times in hot ethanol and five times in hot water. The final TBIM-Cr MOF was dried at 373 K and evacuated at 423 K for 12 h.

3.4. Synthesis of TIBM-Cu

Copper acetate (1.4 g) and TIBM (0.78 g) were dissolved in 30 mL water/ethanol (2:1) solution by sonication. HF (120 µL) was added to the mixture as a module. The reaction mixture was transferred into a Teflon autoclave reactor sealed in a stainless-steel vessel and kept at 423 K for 24 h. The blue-gray-colored precipitate was washed five times in ethanol and dried at 373 K, and then evacuated at 393 K for 12 h.

3.5. Synthesis of TIBM-Al

Aluminum chloride hexahydrate was dehydrated at 373 K for 10 h to remove water from the metal precursor. The dehydrated Al–metal precursor (1.2 g) and TIBM (0.626 g) were mixed in 30 mL water/ethanol (1:1) by sonication for 10 min. HF (60 µL) was added to the mixture as a module. The reaction mixture was transferred into a Teflon autoclave reactor sealed in a stainless-steel vessel, and maintained at 423 K for 48 h. The white-colored powder precipitate was washed five times in ethanol, subsequently dried at 373 K, and evacuated at 393 K for 12 h (Figure 10).

3.6. Characterization

Fourier-transform infrared spectra (FTIR) in the range of 400–4000 cm−1 were obtained with a Spectrum Two, PerkinElmer, UK spectrometer. A mixture of TIBM MOF power and KBr in the weight ratio of 1:99 was used to prepare the sample. The TIBM-MOFs particle morphology and crystal size were determined by means of SEM (Merlin compact, Carl Zeiss) at an accelerating voltage of 1 kV/10 kV, with a current of 10 μA. A volumetric method was used to analyze morphological properties, such as specific surface area, pore size distribution, and pore volume, using a BELSORP-mini (Microtrac BEL, Osaka, Japan) based on N2 adsorption isotherm at 77 K. XRD analysis (X’Pert Pro-MPD, PANalytical, Almelo, The Netherlands) was performed to determine the TIBM-MOFs crystallinity. For the thermal stability test, the samples were heated to 800 °C at a heating rate of 20 °C/min under N2 atmosphere (50 mL/min), using a TGA instrument (Scinco TGA N1000, Twin Lakes, WI, USA). 1H-NMR spectra were obtained with a Bruker 400 MHz NMR spectrometer in CDCl3, using tetramethylsilane (TMS) as an internal standard. Mass spectra were recorded on a Bruker Daltonik, Bremen, Germany, operated in linear mode with a pulsed nitrogen laser (337 nm, pulse frequency, 2 Hz).

3.7. CO2 and N2 Adsorption Capacity Measurements

CO2 and N2 adsorption on TIBM-MOFs were measured at 298–338 K at the pressure of 0–1 bar using a volumetric apparatus (BELSORP-mini, MicrotracBEL, Osaka, Japan). TIBM MOFs were first evacuated at 393 K (TIBM-Cu, TIBM-Al) and 423 K (TIBM-Cr) for 12 h, to remove impurities. A water circulating jacket connected to a thermostatic bath was used to control the measurement temperature with a precision of ±0.01 °C. A reference gas (helium) was used to determine the free space in the sample holder.

4. Conclusions

In summary, we have developed novel TIBM MOFs using a simple and inexpensive solvothermal process. The TIBM linker was prepared via condensation method, using trimesic acid and o-phenylenediamine. Interestingly, metal-modified TIBM MOFs (Cu-TIBM, Cr-TIBM, Al-TIBM) showed modifications in characteristic properties such as morphology, surface area, and pore size distribution. In particular, TIBM-Cu showed a remarkable CO2 adsorption capacity (3.60 mmol/g) and selectivity (53), compared to TIBM-Cr and TIBM-Al. Due to the presence of open metal sites, N atoms of the imidazole functional groups, and an ideal pore size (0.3–1.5 nm) for selective CO2 adsorption (the size of the pore aperture is similar to the size of CO2), the TIBM-Cu CO2 adsorption capacity is higher than that of previously reported MOFs, including MOF-5 (2.1 mmol/g), ZIF-8 (0.8 mmol/g), MIL-101 (Cr) (1.8 mmol/g) [112], MIL-101 (Cr, Mg) (2.0 mmol/g) [113], UiO-66 (2.5 mmol/g) [112], UiO-66- NH2 (3 mmol/g) [114], and MOF-177 (0.8 mmol/g) [115,116]. Moreover, the TIBM-Cr MOF high surface area (2141 m2/g) and pore volume (2.116 cm3/g) points to the material potential for further applications.

Supplementary Materials

The following figures are available online at https://www.mdpi.com/article/10.3390/app11219856/s1, Figure S1: NMR spectra of TIBM linker.

Author Contributions

Writing—original draft preparation, conceptualization, and methodology, S.G. and R.K.C.; validation, S.G. and R.G.; supervision and writing—review and editing, S.H. and S.K.; investigation and resources, S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Changwon National University from 2021 to 2022.

Data Availability Statement

Not applicable.

Acknowledgments

This research was supported by the Changwon National University from 2021 to 2022.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Keith, D.W. Why capture CO2 from the atmosphere? Science 2009, 325, 1654–1655. [Google Scholar] [CrossRef]
  2. Raupach, M.R.; Marland, G.; Ciais, P.; Le Quéré, C.; Canadell, J.G.; Klepper, G.; Field, C.B. Global and regional drivers of accelerating CO2 emissions. Proc. Natl. Acad. Sci. USA 2007, 104, 10288–10293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Canadell, J.G.; Le Quéré, C.; Raupach, M.R.; Field, C.B.; Buitenhuis, E.T.; Ciais, P.; Conway, T.J.; Gillett, N.P.; Houghton, R.; Marland, G. Contributions to accelerating atmospheric CO2 growth from economic activity, carbon intensity, and efficiency of natural sinks. Proc. Natl. Acad. Sci. USA 2007, 104, 18866–18870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Quadrelli, R.; Peterson, S. The energy-climate challenge: Recent trends in CO2 emissions from fuel combustion. Energy Policy 2007, 35, 5938–5952. [Google Scholar] [CrossRef]
  5. Ma, J.; Ying, Y.; Yang, Q.; Ban, Y.; Huang, H.; Guo, X.; Xiao, Y.; Liu, D.; Li, Y.; Yang, W. Mixed-matrix membranes containing functionalized porous metal-organic polyhedrons for the effective separation of CO2-CH4 mixture. Chem. Commun. 2015, 51, 4249–4251. [Google Scholar] [CrossRef]
  6. Zou, X.; Zhang, F.; Thomas, S.; Zhu, G.; Valtchev, V.; Mintova, S. Co3 (HCOO)6 microporous metal-organic framework membrane for separation of CO2/CH4 mixtures. Chem. Eur. J. 2011, 17, 12076–12083. [Google Scholar] [CrossRef]
  7. Conway, W.; Bruggink, S.; Beyad, Y.; Luo, W.; Melián-Cabrera, I.; Puxty, G.; Feron, P. CO2 absorption into aqueous amine blended solutions containing monoethanolamine (MEA), N, N-dimethylethanolamine (DMEA), N, N-diethylethanolamine (DEEA) and 2-amino-2-methyl-1-propanol (AMP) for post-combustion capture processes. Chem. Eng. Sci. 2015, 126, 446–454. [Google Scholar] [CrossRef] [Green Version]
  8. Pinto, D.D.; Monteiro, J.G.-S.; Johnsen, B.; Svendsen, H.F.; Knuutila, H. Density measurements and modelling of loaded and unloaded aqueous solutions of MDEA (N-methyldiethanolamine), DMEA (N, N-dimethylethanolamine), DEEA (diethylethanolamine) and MAPA (N-methyl-1, 3-diaminopropane). Int. J. Greenh. Gas Control 2014, 25, 173–185. [Google Scholar] [CrossRef] [Green Version]
  9. Zhao, H.; Luo, X.; Zhang, H.; Sun, N.; Wei, W.; Sun, Y. Carbon-based adsorbents for post-combustion capture: A review. Greenh. Gases Sci.Technol. 2018, 8, 11–36. [Google Scholar] [CrossRef]
  10. Yu, J.; Xie, L.-H.; Li, J.-R.; Ma, Y.; Seminario, J.M.; Balbuena, P.B. CO2 capture and separations using MOFs: Computational and experimental studies. Chem. Rev. 2017, 117, 9674–9754. [Google Scholar] [CrossRef]
  11. He, X.J.E. A review of material development in the field of carbon capture and the application of membrane-based processes in power plants and energy-intensive industries. Energy Sustain. Soc. 2018, 8, 34. [Google Scholar] [CrossRef]
  12. Lawal, O.; Bello, A.; Idem, R. The role of methyl diethanolamine (MDEA) in preventing the oxidative degradation of CO2 loaded and concentrated aqueous monoethanolamine (MEA)-MDEA blends during CO2 absorption from flue gases. Ind. Eng. Chem. Res. 2005, 44, 1874–1896. [Google Scholar] [CrossRef]
  13. Jeong, W.; Kim, J. Understanding the mechanisms of CO2 adsorption enhancement in pure silica zeolites under humid conditions. J. Phys. Chem. C 2016, 120, 23500–23510. [Google Scholar] [CrossRef]
  14. Ravi, S.; Kang, D.H.; Roshan, R.; Tharun, J.; Kathalikkattil, A.C.; Park, D.W. Organic sulphonate salts tethered to mesoporous silicas as catalysts for CO2 fixation into cyclic carbonates. Catal. Sci. Technol. 2015, 5, 1580–1587. [Google Scholar] [CrossRef]
  15. Jee, S.E.; Sholl, D.S. Carbon dioxide and methane transport in DDR zeolite: Insights from molecular simulations into carbon dioxide separations in small pore zeolites. J. Am. Chem. Soc. 2009, 131, 7896–7904. [Google Scholar] [CrossRef]
  16. Su, F.; Lu, C.; Kuo, S.-C.; Zeng, W. Adsorption of CO2 on amine-functionalized Y-type zeolites. Energy Fuels 2010, 24, 1441–1448. [Google Scholar] [CrossRef]
  17. Koirala, R.; Gunugunuri, K.R.; Pratsinis, S.E.; Smirniotis, P.G. Effect of zirconia doping on the structure and stability of CaO-based sorbents for CO2 capture during extended operating cycles. J. Phys. Chem. C 2011, 115, 24804–24812. [Google Scholar] [CrossRef]
  18. Wang, Q.; Tay, H.H.; Zhong, Z.; Luo, J.; Borgna, A. Synthesis of high-temperature CO2 adsorbents from organo-layered double hydroxides with markedly improved CO2 capture capacity. Energy Environ. Sci. 2012, 5, 7526–7530. [Google Scholar] [CrossRef]
  19. Plaza, M.; González, A.; Pevida, C.; Pis, J.; Rubiera, F. Valorisation of spent coffee grounds as CO2 adsorbents for postcombustion capture applications. Appl. Energy 2012, 99, 272–279. [Google Scholar] [CrossRef] [Green Version]
  20. Wang, J.; Heerwig, A.; Lohe, M.R.; Oschatz, M.; Borchardt, L.; Kaskel, S. Fungi-based porous carbons for CO2 adsorption and separation. J. Mater. Chem. 2012, 22, 13911–13913. [Google Scholar] [CrossRef]
  21. Yu, J.; Balbuena, P.B. How impurities affect CO2 capture in metal-organic frameworks modified with different functional groups. ACS Sustain. Chem. Eng. 2015, 3, 117–124. [Google Scholar] [CrossRef]
  22. Hwang, G.-Y.; Roshan, R.; Ryu, H.-S.; Jeong, H.-M.; Ravi, S.; Kim, M.-I.; Park, D.-W. A highly efficient zeolitic imidazolate framework catalyst for the co-catalyst and solvent free synthesis of cyclic carbonates from CO2. J. CO2 Util. 2016, 15, 123–130. [Google Scholar] [CrossRef]
  23. Chen, J.; Yang, J.; Hu, G.; Hu, X.; Li, Z.; Shen, S.; Radosz, M.; Fan, M. Enhanced CO2 capture capacity of nitrogen-doped biomass-derived porous carbons. ACS Sustain. Chem. Eng. 2016, 4, 1439–1445. [Google Scholar] [CrossRef]
  24. Wei, J.; Zhou, D.; Sun, Z.; Deng, Y.; Xia, Y.; Zhao, D. A controllable synthesis of rich nitrogen-doped ordered mesoporous carbon for CO2 capture and supercapacitors. Adv. Funct. Mater. 2013, 23, 2322–2328. [Google Scholar] [CrossRef]
  25. Cote, A.P.; Benin, A.I.; Ockwig, N.W.; O’Keeffe, M.; Matzger, A.J.; Yaghi, O.M. Porous, crystalline, covalent organic frameworks. Science 2005, 310, 1166–1170. [Google Scholar] [CrossRef] [Green Version]
  26. Furukawa, H.; Yaghi, O.M. Storage of hydrogen, methane, and carbon dioxide in highly porous covalent organic frameworks for clean energy applications. J. Am. Chem. Soc. 2009, 131, 8875–8883. [Google Scholar] [CrossRef]
  27. Ren, S.; Bojdys, M.J.; Dawson, R.; Laybourn, A.; Khimyak, Y.Z.; Adams, D.J.; Cooper, A.I. Porous, fluorescent, covalent triazine-based frameworks via room-temperature and microwave-assisted synthesis. Adv. Mater. 2012, 24, 2357–2361. [Google Scholar] [CrossRef]
  28. Ren, S.; Dawson, R.; Laybourn, A.; Jiang, J.-X.; Khimyak, Y.; Adams, D.J.; Cooper, A.I. Functional conjugated microporous polymers: From 1, 3, 5-benzene to 1, 3, 5-triazine. Polym. Chem. 2012, 3, 928–934. [Google Scholar] [CrossRef]
  29. Cooper, A.I. Conjugated microporous polymers. Adv. Mater. 2009, 21, 1291–1295. [Google Scholar] [CrossRef]
  30. Xu, F.; Chen, X.; Tang, Z.; Wu, D.; Fu, R.; Jiang, D. Redox-active conjugated microporous polymers: A new organic platform for highly efficient energy storage. ChemComm 2014, 50, 4788–4790. [Google Scholar] [CrossRef] [Green Version]
  31. Ben, T.; Ren, H.; Ma, S.; Cao, D.; Lan, J.; Jing, X.; Wang, W.; Xu, J.; Deng, F.; Simmons, J.M. Targeted synthesis of a porous aromatic framework with high stability and exceptionally high surface area. Angew. Chem. 2009, 121, 9621–9624. [Google Scholar] [CrossRef]
  32. Ren, H.; Ben, T.; Wang, E.; Jing, X.; Xue, M.; Liu, B.; Cui, Y.; Qiu, S.; Zhu, G. Targeted synthesis of a 3D porous aromatic framework for selective sorption of benzene. ChemComm 2010, 46, 291–293. [Google Scholar] [CrossRef] [PubMed]
  33. Ben, T.; Pei, C.; Zhang, D.; Xu, J.; Deng, F.; Jing, X.; Qiu, S. Gas storage in porous aromatic frameworks (PAFs). Energy Environ. Sci. 2011, 4, 3991–3999. [Google Scholar]
  34. Monnereau, L.; Nieger, M.; Muller, T.; Bräse, S. Tetrakis-(4-thiyphenyl) methane: Origin of a Reversible 3D-Homopolymer. Adv. Funct. Mater. 2014, 24, 1054–1058. [Google Scholar]
  35. Dawson, R.; Ratvijitvech, T.; Corker, M.; Laybourn, A.; Khimyak, Y.Z.; Cooper, A.I.; Adams, D.J. Microporous copolymers for increased gas selectivity. Polym. Chem. 2012, 3, 2034–2038. [Google Scholar] [CrossRef]
  36. Li, S.; Huo, F. Metal-organic framework composites: From fundamentals to applications. Nanoscale 2015, 7, 7482–7501. [Google Scholar] [CrossRef]
  37. Collins, D.J.; Zhou, H.-C. Hydrogen storage in metal-organic frameworks. J. Mater. Chem. 2007, 17, 3154–3160. [Google Scholar]
  38. Van den Berg, A.W.; Areán, C.O. Materials for hydrogen storage: Current research trends and perspectives. ChemComm 2008, 6, 668–681. [Google Scholar]
  39. Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastre, J. Metal-organic frameworks—Prospective industrial applications. J. Mater. Chem. 2006, 16, 626–636. [Google Scholar]
  40. Pan, L.; Olson, D.H.; Ciemnolonski, L.R.; Heddy, R.; Li, J. Separation of hydrocarbons with a microporous metal-organic framework. Angew. Chem. Int. Ed. 2006, 45, 616–619. [Google Scholar] [CrossRef]
  41. Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Baati, T.; Eubank, J.F.; Heurtaux, D.; Clayette, P.; Kreuz, C. Porous metal-organic-framework nanoscale carriers as a potential platform for drug delivery and imaging. Nat. Mater. 2010, 9, 172–178. [Google Scholar] [CrossRef]
  42. Khutia, A.; Janiak, C. Programming MIL-101Cr for selective and enhanced CO2 adsorption at low pressure by postsynthetic amine functionalization. Dalton Trans. 2014, 43, 1338–1347. [Google Scholar] [CrossRef]
  43. Botas, J.A.; Calleja, G.; Sánchez-Sánchez, M.; Orcajo, M.G. Cobalt doping of the MOF-5 framework and its effect on gas-adsorption properties. Langmuir 2010, 26, 5300–5303. [Google Scholar] [CrossRef]
  44. Liu, Q.; Ding, Y.; Liao, Q.; Zhu, X.; Wang, H.; Yang, J. Fast synthesis of Al fumarate metal-organic framework as a novel tetraethylenepentamine support for efficient CO2 capture. Colloids Surf. A Physicochem. Eng. Asp. 2019, 579, 123645. [Google Scholar] [CrossRef]
  45. Chen, C.; Kim, J.; Park, D.-W.; Ahn, W.-S. Ethylenediamine grafting on a zeolite-like metal organic framework (ZMOF) for CO2 capture. Mater. Lett. 2013, 106, 344–347. [Google Scholar] [CrossRef]
  46. Le, Y.; Guo, D.; Cheng, B.; Yu, J. Amine-functionalized monodispersed porous silica microspheres with enhanced CO2 adsorption performance and good cyclic stability. J. Colloid Interface Sci. 2013, 408, 173–180. [Google Scholar] [CrossRef]
  47. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341. [Google Scholar] [CrossRef] [Green Version]
  48. Lu, W.; Wei, Z.; Gu, Z.-Y.; Liu, T.-F.; Park, J.; Park, J.; Tian, J.; Zhang, M.; Zhang, Q.; Gentle, T., III. Tuning the structure and function of metal-organic frameworks via linker design. Chem. Soc. Rev. 2014, 43, 5561–5593. [Google Scholar] [CrossRef] [PubMed]
  49. Stock, N.; Biswas, S. Synthesis of metal-organic frameworks (MOFs): Routes to various MOF topologies, morphologies, and composites. Chem. Rev. 2012, 112, 933–969. [Google Scholar] [CrossRef]
  50. Lee, Y.-R.; Kim, J.; Ahn, W.-S. Synthesis of metal-organic frameworks: A mini review. Korean J. Chem. Eng. 2013, 30, 1667–1680. [Google Scholar] [CrossRef]
  51. Kashyap, A.; Singh, N.K.; Soni, M.; Soni, A. Deposition of Thin Films by Chemical Solution-Assisted Techniques. In Chemical Solution Synthesis for Materials Design and Thin Film Device Applications; Elsevier: Amsterdam, The Netherlands, 2021; pp. 79–117. [Google Scholar]
  52. Nunes, D.; Pimentel, A.; Santos, L.; Barquinha, P.; Pereira, L.; Fortunato, E.; Martins, R. Synthesis, Design, and Morphology of Metal Oxide Nanostructures. In Metal Oxide Nanostructures; Elsevier: Amsterdam, The Netherlands, 2019; pp. 21–57. [Google Scholar]
  53. Zhou, Y.; Han, L. Recent advances in naphthalenediimide-based metal-organic frameworks: Structures and applications. Coord. Chem. Rev. 2020, 430, 213665. [Google Scholar] [CrossRef]
  54. Chen, C.; Zhang, M.; Zhang, W.; Bai, J. Stable amide-functionalized metal-organic framework with highly selective CO2 adsorption. Inorg. chem. 2019, 58, 2729–2735. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, G.; Leus, K.; Zhao, S.; Van Der Voort, P. Newly designed covalent triazine framework based on novel N-heteroaromatic building blocks for efficient CO2 and H2 capture and storage. ACS Appl. Mater. Interfaces 2017, 10, 1244–1249. [Google Scholar] [CrossRef]
  56. Suresh, V.M.; Bonakala, S.; Atreya, H.S.; Balasubramanian, S.; Maji, T.K. Amide functionalized microporous organic polymer (Am-MOP) for selective CO2 sorption and catalysis. ACS Appl. Mater. Interfaces 2014, 6, 4630–4637. [Google Scholar] [CrossRef] [PubMed]
  57. Dubey, P.; Kumar, C.R.; Babu, B. Solid phase synthesis of benzimidazole ketones and benzimidazole chalcones under solvent-free conditions. Indian J. Chem. 2003, 35, 3128–3130. [Google Scholar] [CrossRef]
  58. Dubey, P.; Naidu, A.; Kumar, C.R. Studies on syntheses of 1-alkyl-2-substitutedthiazolylbenzimidazoles. Indian J. Chem. 2003, 42B, 931–934. [Google Scholar] [CrossRef]
  59. Ramaiah, K.; Dubey, P.; Ramanatham, J.; Kumar, C.R.; Grossert, J. Benzimidazolium dichromates: Efficient reagents for selective oxidation of alcohols to carbonyl compounds. Indian J. Chem. 2003, 42B, 1765–1767. [Google Scholar] [CrossRef]
  60. Dubey, P.; Kumar, C.R.; Reddy, P. Syntheses of 1-alkyl-2-(substituted-2-pyridyl) benzimidazoles. Indian J. Chem. 2003, 42B, 2115–2118. [Google Scholar] [CrossRef]
  61. Dubey, P.; Naidu, A.; Kumar, C.R.; Reddy, P. Preparation of 4-(1-alkyl-benzo [d] imidazole-2-yl)-2-phenyl-2, 3 dihydrobenzo (b) [1, 4] thiazepines. Indian J. Chem. 2003, 42B, 1701–1705. [Google Scholar] [CrossRef]
  62. Dubey, P.; Kumar, R.; Kumar, C.R.; Grossert, J.; Hooper, D. Condensation of o-phenylenediamine with cinnamic acids. Synth. Commun. 2001, 31, 3439–3446. [Google Scholar] [CrossRef]
  63. Lemaire, P.C.; Lee, D.T.; Zhao, J.; Parsons, G.N. Reversible low-temperature metal node distortion during atomic layer deposition of Al2O3 and TiO2 on UiO-66-NH2 metal-organic framework crystal surfaces. ACS Appl. Mater. Interfaces 2017, 9, 22042–22054. [Google Scholar] [CrossRef]
  64. Wang, Y.; Cao, W.; Wang, L.; Zhuang, Q.; Ni, Y. Electrochemical determination of 2, 4, 6-trinitrophenol using a hybrid film composed of a copper-based metal organic framework and electroreduced graphene oxide. Microchim. Acta 2018, 185, 1–9. [Google Scholar] [CrossRef]
  65. Nivetha, R.; Sajeev, A.; Paul, A.M.; Gothandapani, K.; Gnanasekar, S.; Jacob, G.; Sellappan, R.; Raghavan, V.; Pitchaimuthu, S.; Jeong, S.K.; et al. Cu based Metal Organic Framework (Cu-MOF) for electrocatalytic hydrogen evolution reaction. Mater. Res. Express. 2020, 7, 114001. [Google Scholar] [CrossRef]
  66. Yang, A.; Li, P.; Zhong, J. Facile preparation of low-cost HKUST-1 with lattice vacancies and high-efficiency adsorption for uranium. RSC Adv. 2019, 9, 10320–10325. [Google Scholar] [CrossRef] [Green Version]
  67. Mahdipoor, H.R.; Halladj, R.; Babakhani, E.G.; Amjad-Iranagh, S.; Ahari, J.S. Synthesis, characterization, and CO2 adsorption properties of metal organic framework Fe-BDC. RSC Adv. 2021, 11, 5192–5203. [Google Scholar] [CrossRef]
  68. Torres, N.; Galicia, J.; Plasencia, Y.; Cano, A.; Echevarría, F.; Desdin-Garcia, L.; Reguera, E.J.C.; Physicochemical, S.A.; Aspects, E. Implications of structural differences between Cu-BTC and Fe-BTC on their hydrogen storage capacity. Colloids Surf. A Physicochem. Eng. Asp 2018, 549, 138–146. [Google Scholar] [CrossRef]
  69. Hindocha, S.; Poulston, S. Study of the scale-up, formulation, ageing and ammonia adsorption capacity of MIL-100 (Fe), Cu-BTC and CPO-27 (Ni) for use in respiratory protection filters. Faraday Discuss. 2017, 201, 113–125. [Google Scholar] [CrossRef] [Green Version]
  70. Chen, M.-L.; Zhou, S.-Y.; Xu, Z.; Ding, L.; Cheng, Y.-H. Metal-organic frameworks of MIL-100 (Fe, Cr) and MIL-101 (Cr) for aromatic amines adsorption from aqueous solutions. Molecules 2019, 24, 3718. [Google Scholar] [CrossRef] [Green Version]
  71. Simonsson, I.; Gärdhagen, P.; Andrén, M.; Tam, P.L.; Abbas, Z. Experimental investigations into the irregular synthesis of iron (iii) terephthalate metal-organic frameworks MOF-235 and MIL-101. Dalton Trans. 2021, 50, 4976–4985. [Google Scholar] [CrossRef]
  72. Israr, F.; Chun, D.; Kim, Y.; Kim, D.K. High yield synthesis of Ni-BTC metal-organic framework with ultrasonic irradiation: Role of polar aprotic DMF solvent. Ultrason. Sonochem. 2016, 31, 93–101. [Google Scholar] [CrossRef]
  73. Li, X.; Liu, H.; Jia, X.; Li, G.; An, T.; Gao, Y. Novel approach for removing brominated flame retardant from aquatic environments using Cu/Fe-based metal-organic frameworks: A case of hexabromocyclododecane (HBCD). Sci. Total Environ. 2018, 621, 1533–1541. [Google Scholar] [CrossRef]
  74. Cheedarala, R.K.; Park, E.J.; Park, Y.B.; Park, H.W. Highly wettable CuO: Graphene oxide core–shell porous nanocomposites for enhanced critical heat flux. Phys. Status Solidi B 2015, 212, 1756–1766. [Google Scholar] [CrossRef]
  75. Nahian, S.A.; Cheedarala, R.K.; Ahn, K.K. A study of sustainable green current generated by the fluid-based triboelectric nanogenerator (FluTENG) with a comparison of contact and sliding mode. Nano Energy 2017, 38, 447–456. [Google Scholar] [CrossRef]
  76. Hadjiivanov, K.I.; Panayotov, D.A.; Mihaylov, M.Y.; Ivanova, E.Z.; Chakarova, K.K.; Andonova, S.M.; Drenchev, N.L. Power of infrared and raman spectroscopies to characterize metal-organic frameworks and investigate their interaction with guest molecules. Chem. Rev. 2020, 121, 1286–1424. [Google Scholar] [CrossRef]
  77. Autie-Castro, G.; Autie, M.; Rodríguez-Castellón, E.; Aguirre, C.; Reguera, E. Cu-BTC and Fe-BTC metal-organic frameworks: Role of the materials structural features on their performance for volatile hydrocarbons separation. Colloids Surf. A Physicochem. Eng. Asp. 2015, 481, 351–357. [Google Scholar] [CrossRef]
  78. Gopi, S.; Al-Mohaimeed, A.M.; Al-onazi, W.A.; Elshikh, M.S.; Yun, K. Metal organic framework-derived Ni-Cu bimetallic electrocatalyst for efficient oxygen evolution reaction. J. King Saud Univ. Sci. 2021, 33, 101379. [Google Scholar] [CrossRef]
  79. Salama, R.S.; El-Hakama, S.A.; Samraa, S.E.; El-Dafrawya, S.M.; Ahmeda, A.I. Adsorption, equilibrium and kinetic studies on the removal of methyl orange dye from aqueous solution by using of copper metal organic framework (Cu-BDC). Int. J. Modern Chem. 2018, 10, 195–207. [Google Scholar]
  80. Andonova, S.; Ivanova, E.; Yang, J.; Hadjiivanov, K. Adsorption Forms of CO2 on MIL-53 (Al) and MIL-53 (Al)–OH x As Revealed by FTIR Spectroscopy. J. Phys. Chem. C 2017, 121, 18665–18673. [Google Scholar] [CrossRef]
  81. Cheedarala, R.K.; Ahn, K.K. Double characteristic BNO-SPI-TENGs for robust contact electrification by vertical contact separation mode through ion and electron charge transfer. Nano Energy 2018, 44, 430–437. [Google Scholar] [CrossRef]
  82. Cheedarala, R.K.; Parvez, A.N.; Ahn, K.K. Electric impulse spring-assisted contact separation mode triboelectric nanogenerator fabricated from polyaniline emeraldine salt and woven carbon fibers. Nano Energy 2018, 53, 362–372. [Google Scholar] [CrossRef]
  83. Peedikakkal, A.M.P.; Aljundi, I.H. Mixed-Metal Cu-BTC Metal–Organic Frameworks as a Strong Adsorbent for Molecular Hydrogen at Low Temperatures. ACS Omega 2020, 5, 28493–28499. [Google Scholar] [CrossRef]
  84. Peng, Y.; Huang, H.; Zhang, Y.; Kang, C.; Chen, S.; Song, L.; Liu, D.; Zhong, C. A versatile MOF-based trap for heavy metal ion capture and dispersion. Nat. Commun. 2018, 9, 1–9. [Google Scholar] [CrossRef] [Green Version]
  85. Ongari, D.; Tiana, D.; Stoneburner, S.J.; Gagliardi, L.; Smit, B. Origin of the Strong Interaction between Polar Molecules and Copper (II) Paddle-Wheels in Metal Organic Frameworks. J. Phys. Chem. 2017, 121, 15135–15144. [Google Scholar] [CrossRef] [Green Version]
  86. Mishra, P.; Edubilli, S.; Mandal, B.; Gumma, S. Adsorption of CO2, CO, CH4 and N2 on DABCO based metal organic frameworks. Microporous Mesoporous Mater. 2013, 169, 75–80. [Google Scholar] [CrossRef]
  87. Zhang, Z.; Xian, S.; Xi, H.; Wang, H.; Li, Z. Improvement of CO2 adsorption on ZIF-8 crystals modified by enhancing basicity of surface. Chem. Eng. Sci. 2011, 66, 4878–4888. [Google Scholar] [CrossRef]
  88. Munusamy, K.; Sethia, G.; Patil, D.V.; Rallapalli, P.B.S.; Somani, R.S.; Bajaj, H.C. Sorption of carbon dioxide, methane, nitrogen and carbon monoxide on MIL-101 (Cr): Volumetric measurements and dynamic adsorption studies. Chem. Eng. J. 2012, 195, 359–368. [Google Scholar] [CrossRef]
  89. Saha, D.; Bao, Z.; Jia, F.; Deng, S. Adsorption of CO2, CH4, N2O, and N2 on MOF-5, MOF-177, and zeolite 5A. Environ. Sci. Technol. 2010, 44, 1820–1826. [Google Scholar] [CrossRef]
  90. Flaig, R.W.; Osborn Popp, T.M.; Fracaroli, A.M.; Kapustin, E.A.; Kalmutzki, M.J.; Altamimi, R.M.; Fathieh, F.; Reimer, J.A.; Yaghi, O.M. The chemistry of CO2 capture in an amine-functionalized metal-organic framework under dry and humid conditions. J. Am. Chem. Soc. 2017, 139, 12125–12128. [Google Scholar] [CrossRef] [Green Version]
  91. Lee, Y.G.; Moon, H.R.; Cheon, Y.E.; Suh, M.P. A comparison of the H2 sorption capacities of isostructural metal-organic frameworks with and without accessible metal sites: [{Zn2 (abtc)(dmf) 2} 3] and [{Cu2 (abtc)(dmf) 2} 3] versus [{Cu2 (abtc)} 3]. Angew. Chem. 2008, 120, 7855–7859. [Google Scholar] [CrossRef]
  92. Sumida, K.; Horike, S.; Kaye, S.S.; Herm, Z.R.; Queen, W.L.; Brown, C.M.; Grandjean, F.; Long, G.J.; Dailly, A.; Long, J.R. Hydrogen storage and carbon dioxide capture in an iron-based sodalite-type metal-organic framework (Fe-BTT) discovered via high-throughput methods. Chem. Sci. 2010, 1, 184–191. [Google Scholar] [CrossRef]
  93. Hou, L.; Shi, W.-J.; Wang, Y.-Y.; Guo, Y.; Jin, C.; Shi, Q.-Z. A rod packing microporous metal–organic framework: Unprecedented ukv topology, high sorption selectivity and affinity for CO2. ChemComm 2011, 47, 5464–5466. [Google Scholar] [CrossRef]
  94. Mallick, A.; Saha, S.; Pachfule, P.; Roy, S.; Banerjee, R. Selective CO2 and H2 adsorption in a chiral magnesium-based metal organic framework (Mg-MOF) with open metal sites. J. Mater. Chem. 2010, 20, 9073–9080. [Google Scholar] [CrossRef]
  95. Zhang, Z.; Xiang, S.; Rao, X.; Zheng, Q.; Fronczek, F.R.; Qian, G.; Chen, B. A rod packing microporous metal-organic framework with open metal sites for selective guest sorption and sensing of nitrobenzene. ChemComm 2010, 46, 7205–7207. [Google Scholar] [CrossRef]
  96. Kim, S.-N.; Kim, J.; Kim, H.-Y.; Cho, H.-Y.; Ahn, W.-S. Adsorption/catalytic properties of MIL-125 and NH2-MIL-125. Catal. Today 2013, 204, 85–93. [Google Scholar] [CrossRef]
  97. Qin, J.-S.; Du, D.-Y.; Li, W.-L.; Zhang, J.-P.; Li, S.-L.; Su, Z.-M.; Wang, X.-L.; Xu, Q.; Shao, K.-Z.; Lan, Y.-Q. N-rich zeolite-like metal-organic framework with sodalite topology: High CO2 uptake, selective gas adsorption and efficient drug delivery. Chem. Sci. 2012, 3, 2114–2118. [Google Scholar] [CrossRef]
  98. Zhang, G.; Wei, G.; Liu, Z.; Oliver, S.R.; Fei, H. A robust sulfonate-based metal-organic framework with permanent porosity for efficient CO2 capture and conversion. Chem. Mater. 2016, 28, 6276–6281. [Google Scholar] [CrossRef]
  99. Liao, P.-Q.; Zhou, D.-D.; Zhu, A.-X.; Jiang, L.; Lin, R.-B.; Zhang, J.-P.; Chen, X.-M. Strong and dynamic CO2 sorption in a flexible porous framework possessing guest chelating claws. J. Am. Chem. Soc. 2012, 134, 17380–17383. [Google Scholar] [CrossRef]
  100. Jiang, Z.-R.; Wang, H.; Hu, Y.; Lu, J.; Jiang, H.-L. Polar group and defect engineering in a metal-organic framework: Synergistic promotion of carbon dioxide sorption and conversion. ChemSusChem 2015, 8, 878–885. [Google Scholar] [CrossRef]
  101. Debatin, F.; Thomas, A.; Kelling, A.; Hedin, N.; Bacsik, Z.; Senkovska, I.; Kaskel, S.; Junginger, M.; Müller, H.; Schilde, U. In Situ Synthesis of an Imidazolate-4-amide-5-imidate Ligand and Formation of a Microporous Zinc–Organic Framework with H2-and CO2-Storage Ability. Angew. Chem. 2010, 122, 1280–1284. [Google Scholar] [CrossRef]
  102. Prasad, T.K.; Suh, M.P. Control of Interpenetration and Gas-Sorption Properties of Metal-Organic Frameworks by a Simple Change in Ligand Design. Chem. Eur. J. 2012, 18, 8673–8680. [Google Scholar] [CrossRef]
  103. Zheng, S.-T.; Bu, J.T.; Li, Y.; Wu, T.; Zuo, F.; Feng, P.; Bu, X. Pore space partition and charge separation in cage-within-cage indium—Organic frameworks with high CO2 uptake. J. Am. Chem. Soc. 2010, 132, 17062–17064. [Google Scholar] [CrossRef]
  104. Choi, H.S.; Suh, M.P. Highly selective CO2 capture in flexible 3D coordination polymer networks. Angew. Chem. Int. Ed. 2009, 48, 6865–6869. [Google Scholar] [CrossRef]
  105. Hu, Z.; Zhang, K.; Zhang, M.; Guo, Z.; Jiang, J.; Zhao, D. A Combinatorial Approach towards Water-Stable Metal-Organic Frameworks for Highly Efficient Carbon Dioxide Separation. ChemSusChem 2014, 7, 2791–2795. [Google Scholar] [CrossRef]
  106. Park, J.; Yuan, D.; Pham, K.T.; Li, J.-R.; Yakovenko, A.; Zhou, H.-C. Reversible alteration of CO2 adsorption upon photochemical or thermal treatment in a metal-organic framework. J. Am. Chem. Soc. 2012, 134, 99–102. [Google Scholar] [CrossRef]
  107. Demessence, A.; D’Alessandro, D.M.; Foo, M.L.; Long, J.R. Strong CO2 binding in a water-stable, triazolate-bridged metal-organic framework functionalized with ethylenediamine. J. Am. Chem. Soc. 2009, 131, 8784–8786. [Google Scholar] [CrossRef]
  108. Park, H.J.; Cheon, Y.E.; Suh, M.P. Post-Synthetic Reversible Incorporation of Organic Linkers into Porous Metal-Organic Frameworks through Single-Crystal-to-Single-Crystal Transformations and Modification of Gas-Sorption Properties. Chem. Eur. J. 2010, 16, 11662–11669. [Google Scholar] [CrossRef]
  109. Hong, D.H.; Suh, M.P. Enhancing CO2 Separation Ability of a Metal-Organic Framework by Post-Synthetic Ligand Exchange with Flexible Aliphatic Carboxylates. Chem. Eur. J. 2014, 20, 426–434. [Google Scholar] [CrossRef]
  110. Chandrashekhar, N.; Thomas, B.; Gayathri, V.; Ramanathan, K.; Nanje Gowda, N.M. Synthesis and NMR spectral assignments of novel nitrogen and sulfur heterocyclic compounds. Magn. Reson. Chem. 2008, 46, 769–774. [Google Scholar] [CrossRef]
  111. Xiong, J.-F.; Luo, S.-H.; Huo, J.-P.; Liu, J.-Y.; Chen, S.-X.; Wang, Z.-Y. Design, Synthesis, and Characterization of 1, 3, 5-Tri (1 H-benzo [d] imidazol-2-yl) benzene-Based Fluorescent Supramolecular Columnar Liquid Crystals with a Broad Mesomorphic Range. J. Org. Chem. 2014, 79, 8366–8373. [Google Scholar] [CrossRef]
  112. Lin, Y.; Kong, C.; Chen, L. Direct synthesis of amine-functionalized MIL-101 (Cr) nanoparticles and application for CO2 capture. RSC Adv. 2012, 2, 6417–6419. [Google Scholar] [CrossRef]
  113. Gaikwad, S.; Kim, S.-J.; Han, S. CO2 capture using amine-functionalized bimetallic MIL-101 MOFs and their stability on exposure to humid air and acid gases. Microporous Mesoporous Mater. 2019, 277, 253–260. [Google Scholar] [CrossRef]
  114. Cao, Y.; Zhao, Y.; Lv, Z.; Song, F.; Zhong, Q. Preparation and enhanced CO2 adsorption capacity of UiO-66/graphene oxide composites. J. Ind. Eng. Chem. 2015, 27, 102–107. [Google Scholar] [CrossRef]
  115. Walton, K.S.; Millward, A.R.; Dubbeldam, D.; Frost, H.; Low, J.J.; Yaghi, O.M.; Snurr, R.Q. Understanding inflections and steps in carbon dioxide adsorption isotherms in metal-organic frameworks. J. Am. Chem. Soc. 2008, 130, 406–407. [Google Scholar] [CrossRef] [PubMed]
  116. Gaikwad, S.; Kim, Y.; Gaikwad, R.; Han, S. Enhanced CO2 capture capacity of amine-functionalized MOF-177 metal organic framework. J. Environ. Chem. Eng. 2021, 9, 105523. [Google Scholar] [CrossRef]
Figure 1. (a) N2 adsorption isotherms and (b) pore size distribution of TIBM-Al, TIBM-Cr, and MOF-Cu at 77 K.
Figure 1. (a) N2 adsorption isotherms and (b) pore size distribution of TIBM-Al, TIBM-Cr, and MOF-Cu at 77 K.
Applsci 11 09856 g001
Figure 2. SEM images of (a) TIBM-Al, (b) TIBM-Cr, and (c) TIBM-Cu.
Figure 2. SEM images of (a) TIBM-Al, (b) TIBM-Cr, and (c) TIBM-Cu.
Applsci 11 09856 g002
Figure 3. XRD patterns of MOF-Al, MOF-Cr, and MOF-Cu. (The dashed boxes represent the characteristic peaks that are matched with those of the reported MOFs).
Figure 3. XRD patterns of MOF-Al, MOF-Cr, and MOF-Cu. (The dashed boxes represent the characteristic peaks that are matched with those of the reported MOFs).
Applsci 11 09856 g003
Figure 4. FTIR spectra of (a) TIBM−Al, TIBM−Cr, and TIBM−Cu, and (b) TIBM.
Figure 4. FTIR spectra of (a) TIBM−Al, TIBM−Cr, and TIBM−Cu, and (b) TIBM.
Applsci 11 09856 g004aApplsci 11 09856 g004b
Figure 5. Thermal degradation graphs for TIBM-Al, TIBM-Cr, and TIBM-Cu.
Figure 5. Thermal degradation graphs for TIBM-Al, TIBM-Cr, and TIBM-Cu.
Applsci 11 09856 g005
Figure 6. (a) CO2 and (b) N2 adsorption capacities at 298 K for TIBM-Al, TIBM-Cr, and TIBM-Cu MOFs.
Figure 6. (a) CO2 and (b) N2 adsorption capacities at 298 K for TIBM-Al, TIBM-Cr, and TIBM-Cu MOFs.
Applsci 11 09856 g006aApplsci 11 09856 g006b
Figure 7. CO2/N2 selectivity for TIBM-Al, TIBM-Cr, and TIBM-Cu. Error bars represent the standard deviation (n = 3).
Figure 7. CO2/N2 selectivity for TIBM-Al, TIBM-Cr, and TIBM-Cu. Error bars represent the standard deviation (n = 3).
Applsci 11 09856 g007
Figure 8. CO2 adsorption capacity of TIBM-Cu MOF at increasing temperatures.
Figure 8. CO2 adsorption capacity of TIBM-Cu MOF at increasing temperatures.
Applsci 11 09856 g008
Figure 9. Synthesis of TIBM linker.
Figure 9. Synthesis of TIBM linker.
Applsci 11 09856 g009
Figure 10. Synthesis of TIBM-MOFs (i.e., M= Al, Cr, and Cu).
Figure 10. Synthesis of TIBM-MOFs (i.e., M= Al, Cr, and Cu).
Applsci 11 09856 g010
Table 1. Surface area and pore volume of TIBM-MOFs.
Table 1. Surface area and pore volume of TIBM-MOFs.
Sample NameBET Surface Area (m2/g)Total Pore Volume (cm3/g)
TIBM-Al505.190.681
TIBM-Cr2141.12.116
TIBM-Cu1073.50.778
Table 2. Matching the characteristics peaks of TIMB MOFs with reported MOFs.
Table 2. Matching the characteristics peaks of TIMB MOFs with reported MOFs.
Synthesized MOFsMatching Peak Angles (2θ)Reported MOFs with Matching PeaksKnown Structure
TIBM-Al9.1, 12.5–20, 24–29Cu-BTC, Ni-BTC [67,68,72]Hexagonal, cubic
TIBM-Cu9.4–20, 28.8, 35.3, 38.9Cu-BTC, HKUST-1 [64,65]Cubic, pyramidal
TIBM-Cr10.9, 14, 20, 24Fe-BTC, MIL-100, MOF-235 [67,68,69,71,73]Irregular spherical
Table 3. CO2 adsorption capacities of different MOFs.
Table 3. CO2 adsorption capacities of different MOFs.
AdsorbentCondition (Temperature/Pressure)CO2 Adsorption Capacity (mmol/g)Reference
TIBM-Cu298 K/1 bar3.6This work
TIBM-Al298 K/1 bar2.0This work
TIBM-Cr298 K/1 bar1.6This work
IRMOF-74-III-(CH2NH2)2298 K/1 bar3[90]
SNU-5195 K /1 bar2.6[91]
Fe-BTT298 K/1 bar3.1[92]
[Cd2L1(H2O)]2293 K /1 bar2.1[93]
[Mg(3,5-PDC)(H2O)]298 K/1 bar0.6[94]
[Zn4(OH)2(1,2,4-BTC)2(H2O)2]295 K/1 bar1.9[95]
NH2-MIL-125298 K/1 bar2.2[96]
IRMOF-74-III-(CH2NH2)2298 K/1 bar3[90]
IFMC-1298 K/1 bar2.7[97]
TMOF-1298 K/1 bar1.4[98]
MAF-23298 K/1 bar2.5[99]
USTC-253298 K/1 bar2.1[100]
[Zn(L2)]n298 K/1 bar2.1[101]
USTC-253-TFA298 K/1 bar2.9[100]
SNU-71298 K/1 bar1[102]
CPM-5299 K/1 bar2.4[103]
SNU-M10298 K/1 bar2.1[104]
UiO-66-SO3H-0.15298 K/1 bar2.2[105]
SHF-61295 K/1 bar1[106]
Cu-BTTri-ens298 K/1 bar1.3[107]
SNU-31298 K/1 bar0.6[108]
UiO-66-AD4298 K/1 bar1.9[109]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gaikwad, S.; Cheedarala, R.K.; Gaikwad, R.; Kim, S.; Han, S. Controllable Synthesis of 1, 3, 5-tris (1H-benzo[d]imidazole-2-yl) Benzene-Based MOFs. Appl. Sci. 2021, 11, 9856. https://doi.org/10.3390/app11219856

AMA Style

Gaikwad S, Cheedarala RK, Gaikwad R, Kim S, Han S. Controllable Synthesis of 1, 3, 5-tris (1H-benzo[d]imidazole-2-yl) Benzene-Based MOFs. Applied Sciences. 2021; 11(21):9856. https://doi.org/10.3390/app11219856

Chicago/Turabian Style

Gaikwad, Sanjit, Ravi Kumar Cheedarala, Ranjit Gaikwad, Soonho Kim, and Sangil Han. 2021. "Controllable Synthesis of 1, 3, 5-tris (1H-benzo[d]imidazole-2-yl) Benzene-Based MOFs" Applied Sciences 11, no. 21: 9856. https://doi.org/10.3390/app11219856

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop