Next Article in Journal
The Electro-Optical Properties and Adhesion Strength of Epoxy-Polymercaptan-Based Polymer Dispersed Liquid Crystal Films
Previous Article in Journal
Hydrothermal Synthesis, Crystal Structure, and Spectroscopic Properties of Pure and Eu3+-Doped NaY[SO4]2 ∙ H2O and Its Anhydrate NaY[SO4]2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two Co(II)-Based MOFs Constructed from Resorcin[4]Arene Ligand: Syntheses, Structures, and Heterogeneous Catalyst for Conversion of CO2

1
Department of Chemistry, College of Art and Science, Northeast Agricultural University, Harbin 150030, China
2
College of Chemistry and Materials Science, Inner Mongolia University for the Nationalities, Tongliao 028000, China
*
Authors to whom correspondence should be addressed.
Crystals 2021, 11(6), 574; https://doi.org/10.3390/cryst11060574
Submission received: 22 April 2021 / Revised: 8 May 2021 / Accepted: 18 May 2021 / Published: 21 May 2021

Abstract

:
Two Co(II)-based metal–organic frameworks (MOFs) with open channels, [(CH3)2NH2]2[Co5L(H2O)8]·4H2O (1) and [Co6L(DMF)2(H2O)8]·2H2O (2), were synthesized using resorcin[4]arene ligand (H12L). Compounds 1 and 2 exhibit different 3D microporous framework structures: 1 possesses two kinds of open channels parallel to the a-axis (ca. 5.0 × 5.0 Å) and the b-axis (ca. 4.0 × 6.0 Å), and 2 is an open framework with a window size of 5.6 × 5.6 Å. The activated crystal 1 involves many Lewis acid sites; thus, 1 shows prominent activity and recyclability for the reaction of carbon dioxide coupled with epoxides. Most strikingly, catalyst 1 can be reused for five successive cycles and provides outstanding catalytic activity.

Graphical Abstract

1. Introduction

Carbon dioxide (CO2) has caused various environmental and energy problems as a major greenhouse gas, but it is an ideal renewable C1 source in nature [1,2,3,4]. Therefore, considerable efforts have been devoted to capturing and converting CO2 into useful chemical products [5,6,7,8], such as CO2 absorption [9,10], photocatalysis [11,12,13], electrocatalysis [14,15,16], and organocatalysis [17,18,19]. Among these methods, CO2 coupling with epoxides is regarded as the most effective means because of the 100% atomic availability and the wide use of cyclic carbonates [20,21,22,23]. Homogeneous and heterogeneous catalysts have been used to catalyze this reaction, including transition-metal complexes, zeolites, organocatalysts, and so on [24,25,26]. Although homogeneous catalysts exhibit efficient catalytic activity for the reaction, the inherent limitations of catalyst separation have prevented their wide application [27]. To overcome these drawbacks, heterogeneous catalysts have been considered [28,29,30,31,32]. The cycloaddition reaction between CO2 and epoxides is a Lewis acid catalysis process; therefore, a catalyst with more active Lewis acid sites provides acceptable conversion of epoxides to cyclic carbonates.
Metal–organic frameworks (MOFs), as a kind of functional material, have attracted tremendous interest due to their large surface area, tunable pore structure, and rich active Lewis acid sites [33,34,35]. MOFs have a high adsorption capacity for CO2, which can increase the concentration of CO2 around the catalytic active sites. Additionally, framework pores can provide confined space for CO2 reaction [36,37]. Organic linkers play a crucial part in the synthesis of MOFs with a variety of pore size and chemical environments [38,39,40]. For this application, resorcin[4]arene is especially attractive because of its multiple coordinate sites and tunable structure. Many elegant structures have been obtained using functionalized resorcin[4]arene ligands [41,42,43,44,45,46,47,48].
Herein, we selected a functionalized-resorcin[4]arene (H12L) as a ligand, with twelve carboxylate groups in one ligand, so it has multiple possible coordination modes with metal ions. In this domain, two Co(II)-based microporous structures, [(CH3)2NH2]2[Co5L(H2O)8]·4H2O (1) and [Co6L(DMF)2(H2O)8]·2H2O (2), were synthesized using Co(BF4)2·6H2O and CoCl2·6H2O with the H12L ligand (Scheme 1). Remarkably, 1 shows outstanding catalytic capability for the conversion of CO2 as a heterogeneous catalyst.

2. Experimental

2.1. Materials and Methods

All the raw materials were obtained commercially. The method through which the H12L ligand was synthesized is consistent with the literature [49]. The PXRD patterns of 1 and 2 were collected using Cu Kα radiation (λ = 0.154 nm) on a Rigaku Dmax 2000 X-ray diffractometer. 1H NMR spectra were captured on a Bruker 600 MHz spectrometer in CDCl3 or DMSO-d6. TGA data were obtained using a TGA5500 analyzer (5 °C min−1, 25–600 °C, N2 flow). The C, H, and N elemental analyses were performed using a Vario MACRO cube analyzer. IR spectra were collected on a Thermo Scientific Nicolet 10. The CO2 gas sorption was performed on V-Sorb 2800S.

2.2. Synthesis of [(CH3)2NH2]2[Co5L(H2O)8]·4H2O (1)

H12L (0.023 g, 0.015 mmol), Co(BF4)2·6H2O (0.028 g, 0.08 mmol), 4 mL of H2O, and 4 mL of dimethylformamide (DMF) were mixed in a 15 mL Teflon reactor. The mixture was heated at 100 °C for 72 hours. The pink samples 1 were harvested by filtration (32% yield). Anal. calcd for C80H92N2O48Co5 (Mr = 2144.20): C, 44.77; H, 4.32; N, 1.31. Found: C, 44.68; H, 4.14; N, 1.29. IR data (KBr, cm−1): 3405 (s), 1606 (s), 1508 (s), 1423 (s), 1322 (m), 1286 (s), 1231 (m), 1184 (m), 1104 (m), 1064 (m), 929 (w), 858 (w), 827(w), 705 (w).

2.3. Synthesis of [Co6L(DMF)2(H2O)8]·2H2O (2)

H12L (0.015 g, 0.006 mmol) and CoCl2·6H2O (0.028 g, 0.12 mmol) were dispersed in DMF/H2O (8 mL, v/v = 6:2), and then placed in a 15 mL Teflon reactor. The mixture was heated at 110 °C for 72 hours. The pink samples 2 were obtained in a 9% yield. Anal. calcd for C82H86N2O48Co6 (Mr = 2221.10): C, 44.34; H, 3.90; N, 1.26. Found: C, 43.99; H, 3.86; N, 1.22. IR data (KBr, cm−1): 3415 (s), 1610 (s), 1502 (s), 1421 (s), 1334 (m), 1286 (s), 1162 (m), 1108 (m), 1064 (m), 1064 (m), 930 (w), 858 (w), 825(w), 708 (w).

2.4. Coupling of CO2 with Epoxides

To obtain the activated sample, catalyst 1 was immersed in acetone for 12 hours and then dried at 60 °C for 10 hours under vacuum. The reactions were executed in a 15 mL flask, the reaction system was refreshed with CO2 three times, and then the CO2 pressure was maintained at 1 atm. Epoxide (5 mmol), catalyst 1 (30 mg, 0.0014 mmol), and n-Bu4NBr (0.16 g, 0.50 mmol) were mixed in the flask, and then stirred at 80 °C for 8 hours. The conversion of the reactions were calculated by 1H NMR.

2.5. X-ray Crystallography

Diffraction data for compounds 1 and 2 were recorded at room temperature using an Oxford Diffraction Gemini R CCD diffractometer with Mo Kα radiation (λ = 0.71073 Å). The structures of 1 and 2 were solved by direct methods (SHELXS-2014) and refined on F2 by full-matrix least-squares using the SHELXS-2014 [50,51,52]. The solvent molecules were highly disordered, so the produced diffused electron densities were removed using the SQUEEZE program in PLATON [53]. Based on the TGA, electron diffraction density, and elemental analysis results, the solvent molecules were directly merged into the final molecular formula. The reflection peaks of hydrogen atoms on the solvent molecules were too weak to assign, so they were directly enclosed in the final molecular formula. Non-H atoms were refined anisotropically. Crystallographic data for 1 (CCDC 2078907) and 2 (CCDC 2078908) are summarized in Table 1, Tables S1 and S2.

3. Results and Discussion

3.1. Structure of [(CH3)2NH2]2[Co5L(H2O)8]·4H2O (1)

Compound 1 crystallizes in the triclinic system with space group P-1. Because of the disordered solvents, the SQUEEZE program in PLATON was used during the refinement. There are twelve water molecules and two [(CH3)2NH2]+ cations, produced by the decomposition and protonation of DMF, in the structure [54,55], which was calculated by elemental analysis, TGA, and electron diffraction density. The asymmetric structure of 1 is composed of two and a half CoII cations (Co1, Co2, and Co3), half a L12− ligand, and four coordinated water molecules (Figure 1a). All the CoII cations adopt a six-coordinate mode but different coordination environments. Co1 is coordinated with four water molecules and two O atoms from two L12− ligands; the occupancy of Co1 is 0.5. Co2 is linked with six O atoms from four L12− ligands. Co3 is surrounded by two water molecules and four O atoms from three L12− ligands. As shown in Figure 1b, each L12− ligand bridges sixteen CoII cations. In this manner, 1 shows a three-dimensional structure. As displayed in Figure 1c,d, there are two types of open channels in the framework with the window sizes of 5.0 × 5.0 Å (Figure 1c) and 4.0 × 6.0 Å (Figure 1d). The solvent-accessible volume is approximately 23.3% (2297.9 Å3), which was estimated by PLATON.

3.2. Structure of [Co6L(DMF)2(H2O)8]·2H2O (2)

The crystal 2 belongs to the triclinic system with space group P-1. The SQUEEZE function was used to remove the disordered solvents. The asymmetric structure of 2 comprises three CoII cations (Co1, Co2, and Co3), half a L12− ligand, and four coordinated water molecules (Figure 2a). Compared with Co2 and Co3, Co1 shows different coordination spheres: Co1 is coordinated with one coordinated water molecule and five O atoms from five L12− ligands; Co2 and Co3 both adopt a six-coordinate mode with one coordinated water molecule, one DMF molecule, and four O atoms from three L12− ligands. The Co–O bond lengths vary from 1.993(4) to 2.209(5) Å and the O–Co–O bond angles vary from 58.27(19)° to 180.00(12)°. As illustrated in Figure 2b, every L12− ligand coordinates with twenty-two CoII cations; as such, neighboring L12− ligands are linked by the CoII cations into an open framework. The window size is 5.6 × 5.6 Å along the a axis (Figure 2c,d). The solvent-accessible volume of compound 2 is ca. 20.0% based on the PLATON calculation.

3.3. Characterization of the Crystal Structure of 1 and 2

The TGA of compounds 1 and 2 was conducted under a N2 atmosphere. As displayed in Figure 3a, the TGA of compound 1 indicated that the weight loss before 240 °C is due to the DMF molecules and water molecules, and the framework begins to collapse after 240 °C. The TGA of compound 2 shows that the weight loss before 300 °C belongs to the DMF molecules and water molecules, then the weight loss from 300 °C is attributable to the framework decomposition. The PXRD pattern of 1 is consistent with the simulated one, which indicates that 1 is stable in air. Some characteristic peaks disappeared in the PXRD pattern of 2, which may have occurred due to an optimum growth orientation being chosen. The CO2 adsorption was performed at 273 K (Figure S4), and the CO2 uptake capacity was found to be ca. 0.48 mmol/g.

3.4. Coupling of CO2 with Epoxides

Given the high-density Lewis acid sites and high yield of compound 1, the heterogeneous catalytic performance of 1 was investigated for the coupling reaction of CO2 with epoxides. As shown in Scheme 2 and Table 2, gylcidylphenylether was selected as a typical substrate to obtain the optimum reaction conditions. Firstly, the reaction between the gylcidylphenylether and CO2 was performed in the presence of activated catalyst 1 (10 mg) and n-Bu4NBr (0.16 g) at 80 °C for 1 hour; the conversion was only 24% (entry 1, Figure S1a). Thus, the catalyst amount was increased from 10 to 20 and 30 mg, and the corresponding conversions were increased from 24% to 26% and 32%, respectively (entries 2 and 3, Figure S1b,c). The conversion was only 23% when the n-Bu4NBr was absent, which indicted that n-Bu4NBr is an important co-catalyst (entry 4, Figure S1d). When the reactions were performed at 25 and 50 °C, the conversions were 0% and 12%, respectively (entries 5 and 6, Figure S1e,f). To improve the conversion, the reaction time was extended to 2, 4, 6, and 8 hours; the corresponding conversions were 51%, 80%, 83%, and 98%, respectively (entries 7–10, Figure S1g–j). These experimental results showed that the optimum reaction conditions are 30 mg catalyst 1, 0.16 g n-Bu4NBr, and 1 atm CO2 reaction at 80 °C for 8 h.
Different epoxides were selected as the substrates to further examine the applicability of catalyst 1. The conversions of epichlorohydrin 2-ethyloxirane, 2-butyloxirane, 2-(butoxymethyl)oxirane, 1,2-epoxyethylbenzene and benzylglycidylether were 99%, 99%, 99%, 95%, 75%, and 93%, respectively (entries 1–6 in Table 3, Figure S2a–f). The reaction rate did not decrease when the length of the alkyl chain increased; these experimental results showed that the substrates do not enter the channel but react on the surface of the catalyst (entries 2–4 in Table 3, Figure S2b–d). Compared with other substrates, the reaction of 1,2-epoxyethylbenzene is relatively slow, which may due to the steric-hinderance effect (entries 5 in Table 3, Figure S2e). These results confirmed that the epoxides with variable alkyl chains or aromatic rings are all suitable substrates for the reaction.
To further explore the recyclability and stability of 1, recycling experiments were performed using gylcidylphenylether as the substrate. After the reaction, 1 was recovered from the mixture by centrifugation and filtration and then washed with dichloromethane three times. The catalyst 1 recovered from the reaction was reused for five consecutive cycles with conversions of more than 90% (Figure 4 and Figure S3a–c). This result proved that 1 is a heterogeneous and recyclable catalyst. The PXRD pattern of activated 1 is not consistent with the simulated one, which may due to the single-crystal to single-crystal transition [56]. However, the PXRD pattern of activated 1 after five recycles is consistent with the activated sample, which indicated that 1 is stable during the reaction process (Figure S5).

4. Conclusions

In summary, we synthesized two microporous MOFs (1 and 2) using Co(II) cations and functionalized resorcin[4]arene. The synthesized compounds 1 and 2 were characterized by single-crystal X-ray diffraction analysis, PXRD, IR, TGA, and elemental analysis. The activated catalyst 1 possesses a large number of unsaturated coordination CoII cations; thus, compound 1 is a promising heterogeneous catalyst for the CO2 conversion reaction. Most strikingly, 1 can be easily recovered and reused for five consecutive circles with high catalytic activity.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cryst11060574/s1, Figure S1: 1H NMR spectrum of the cycloaddition reaction between CO2 and gylcidylphenylether; Figure S2: 1H NMR spectrum of CO2 coupling with different epoxides using 1 as catalysts; Figure S3; 1H NMR spectrum of the cycloaddtion reaction between CO2 and gylcidylphenylether in different circles; Figure S4. CO2 total adsorption isotherm for 1; Figure S5; PXRD patterns of the 1; Table S1: Selected bond distances (Å) and angles (degrees) for 1; Table S2: Selected bond distances (Å) and angles (degrees) for 2. Crystallographic data of 1 and 2 (CIF).

Author Contributions

Conceptualization, B.-B.L., D.W., and F.Y.; data curation, B.-B.L.; formal analysis, B.-B.L. and D.W.; funding acquisition, B.-B.L. and X.H.; investigation B.-B.L.; methodology, B.-B.L.; project administration, B.-B.L.; resources, B.-B.L.; software, B.-B.L. and C.-J.F.; supervision, D.W. and F.Y.; validation, B.-B.L.; visualization, B.-B.L.; writing—original draft, B.-B.L. and X.H.; writing—review and editing, B.-B.L., D.W., and F.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Heilongjiang Province Postdoctoral Foundation (LBH-Z19117), the China Postdoctoral Science Foundation (2020M670874), and the Natural Science Foundation of Inner Mongolia (2020BS02015).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or supplementary material.

Acknowledgments

We would like to faithfully thank Northeast Normal University for carried out the XRD measurement.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Artz, J.; Müller, T.E.; Thenert, K.; Kleinekorte, J.; Meys, R.; Sternberg, A.; Bardow, A.; Leitne, W. Sustainable Conversion of Carbon Dioxide: An Integrated Review of Catalysis and Life Cycle Assessment. Chem. Rev. 2018, 118, 434–504. [Google Scholar] [CrossRef] [PubMed]
  2. He, J.; Janaky, C. Recent Advances in Solar-Driven Carbon Dioxide Conversion: Expectations versus Reality. ACS Energy Lett. 2020, 5, 1996–2014. [Google Scholar] [CrossRef] [PubMed]
  3. Sumida, K.; Rogow, D.L.; Mason, J.A.; McDonald, T.M.; Bloch, E.D.; Herm, Z.R.; Bae, T.-H.; Long, J.R. Carbon Dioxide Capture in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 724–781. [Google Scholar] [CrossRef] [PubMed]
  4. Francke, R.; Schille, B.; Roemelt, M. Homogeneously Catalyzed Electroreduction of Carbon Dioxide Methods, Mechanisms, and Catalysts. Chem. Rev. 2018, 118, 4631–4701. [Google Scholar] [CrossRef]
  5. Yan, X.; Chen, C.; Wu, Y.; Liu, S.; Chen, Y.; Feng, R.; Zhang, J.; Han, B. Efficient electroreduction of CO2 to C2+ products on CeO2 modified CuO. Chem. Sci. 2021, 12, 6638–6645. [Google Scholar] [CrossRef]
  6. Jiang, X.; Nie, X.; Guo, X.; Song, C.; Chen, J.G. Recent Advances in Carbon Dioxide Hydrogenation to Methanol via Heterogeneous Catalysis. Chem. Rev. 2020, 120, 7984–8034. [Google Scholar] [CrossRef]
  7. Ra, E.C.; Kim, K.Y.; Kim, E.H.; Lee, H.; An, K.; Lee, J.S. Recycling Carbon Dioxide through Catalytic Hydrogenation: Recent Key Developments and Perspectives. ACS Catal. 2020, 10, 11318–11345. [Google Scholar] [CrossRef]
  8. Dey, S.; Bhunia, A.; Breitzke, H.; Groszewicz, P.B.; Buntkowsky, G.; Janiak, C. Two linkers are better than one: Enhancing CO2 capture and separation with porous covalent triazine-based frameworks from mixed nitrile linkers. J. Mater. Chem. A 2017, 5, 3609–3620. [Google Scholar] [CrossRef]
  9. Ding, M.; Flaig, R.W.; Jiang, H.-L.; Yaghi, O.M. Carbon capture and conversion using metal-organic frameworks and MOF-based materials. Chem. Soc. Rev. 2019, 48, 2783–2828. [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. Fang, Z.-B.; Liu, T.-T.; Liu, J.; Jin, S.; Wu, X.-P.; Gong, X.-Q.; Wang, K.; Yin, Q.; Liu, T.-F.; Cao, R.; et al. Boosting Interfacial Charge-Transfer Kinetics for Efficient Overall CO2 Photoreduction via Rational Design of Coordination Spheres on Metal-Organic Frameworks. J. Am. Chem. Soc. 2020, 142, 12515–12523. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, S.-S.; Huang, H.-H.; Liu, M.; Yao, S.; Guo, S.; Wang, J.-W.; Zhang, Z.-M.; Lu, T.-B. Encapsulation of Single Iron Sites in a Metal-Porphyrin Framework for High-Performance Photocatalytic CO2 Reduction. Inorg. Chem. 2020, 59, 6301–6307. [Google Scholar] [CrossRef] [PubMed]
  13. Qin, D.; Zhou, Y.; Wang, W.; Zhang, C.; Zeng, G.; Huang, D.; Wang, L.; Wang, H.; Yang, Y.; Lei, L.; et al. Recent advances in two-dimensional nanomaterials for photocatalytic reduction of CO2: Insights into performance, theories and perspective. J. Mater. Chem. A 2020, 8, 19156–19195. [Google Scholar] [CrossRef]
  14. Zhu, Q.; Yang, D.; Liu, H.; Sun, X.; Chen, C.; Bi, J.; Liu, J.; Wu, H.; Han, B. Hollow metal organic framework-mediated in-situ architecture of copper dendrites for enhanced CO2 electroreduction. Angew. Chem. Int. Ed. 2020, 59, 8896–8901. [Google Scholar] [CrossRef] [PubMed]
  15. Das, S.; Pérez-Ramírez, J.; Gong, J.; Dewangan, N.; Hidajat, K.; Gates, B.C.; Kawi, S. Core–shell structured catalysts for thermocatalytic, photocatalytic, and electrocatalytic conversion of CO2. Chem. Soc. Rev. 2020, 49, 2937–3004. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, L.; Li, X.-X.; Lang, Z.-L.; Liu, Y.; Liu, J.; Yuan, L.; Lu, W.-Y.; Xia, Y.-S.; Dong, L.-Z.; Yuan, D.-Q.; et al. Enhanced Cuprophilic Interactions in Crystalline Catalysts Facilitate the Highly Selective Electroreduction of CO2 to CH4. J. Am. Chem. Soc. 2021, 143, 3808–3816. [Google Scholar] [CrossRef]
  17. Song, Q.-W.; Zhou, Z.-H.; He, L.-N. Efficient, selective and sustainable catalysis of carbon dioxide. Green Chem. 2017, 19, 3707–3728. [Google Scholar] [CrossRef]
  18. Shaikh, R.R.; Pornpraprom, S.; D’Elia, V. Catalytic Strategies for the Cycloaddition of Pure, Diluted, and Waste CO2 to Epoxides under Ambient Conditions. ACS Catal. 2018, 8, 419–450. [Google Scholar] [CrossRef]
  19. Kamphuis, A.J.; Picchioni, F.; Pescarmona, P.P. CO2-fixation into cyclic and polymeric carbonates: Principles and applications. Green Chem. 2019, 21, 406–448. [Google Scholar] [CrossRef] [Green Version]
  20. Xu, W.; Chen, H.; Jie, K.; Yang, Z.; Li, T.; Dai, S. Entropy-Driven Mechanochemical Synthesis of Polymetallic ZeoliticImidazolate Frameworks for CO2 Fixation. Angew. Chem. Int. Ed. 2019, 58, 5018–5022. [Google Scholar] [CrossRef]
  21. Ji, H.; Naveen, K.; Lee, W.; Kim, T.S.; Kim, D.; Cho, D.-H. Pyridinium-Functionalized Ionic Metal-Organic Frameworks Designed as Bifunctional Catalysts for CO2 Fixation into Cyclic Carbonates. ACS Appl. Mater. Interfaces 2020, 12, 24868–24876. [Google Scholar] [CrossRef] [PubMed]
  22. Lu, X.-B.; Darensbourg, D.J. Cobalt Catalysts for The Coupling of CO2 and Epoxides to Provide Polycarbonates and Cyclic Carbonates. Chem. Soc. Rev. 2012, 41, 1462–1484. [Google Scholar] [CrossRef]
  23. Li, P.-Z.; Wang, X.-J.; Liu, J.; Lim, J.S.; Zou, R.; Zhao, Y.L. A triazole-containing metal-organic framework as a highly effective and substrate size-dependent catalyst for CO2 conversion. J. Am. Chem. Soc. 2016, 138, 2142–2145. [Google Scholar] [CrossRef] [PubMed]
  24. Ding, M.; Jiang, H.-L. Incorporation of Imidazolium-Based Poly(ionic liquid)s into a Metal-Organic Framework for CO2 Capture and Conversion. ACS Catal. 2018, 8, 3194–3201. [Google Scholar] [CrossRef]
  25. Zhou, L.-J.; Sun, W.; Yang, N.-N.; Li, P.; Gong, T.; Sun, W.-J.; Sui, Q.; Gao, E.-Q. A Facile and Versatile “Click” Approach Toward Multifunctional Ionic Metal-organic Frameworks for Efficient Conversion of CO2. ChemSusChem 2019, 12, 2202–2210. [Google Scholar] [CrossRef] [PubMed]
  26. Kuruppathparambil, R.R.; Babu, R.; Jeong, H.M.; Hwang, G.-Y.; Jeong, G.S.; Kim, M.-I.; Kim, D.-W.; Park, D.-W. A Solid Solution Zeolitic Imidazolate Framework as A Room Temperature Efficient Catalyst for the Chemical Fixation of CO2. Green Chem. 2016, 18, 6349–6356. [Google Scholar] [CrossRef]
  27. Cuesta-Aluja, L.; Masdeu-Bultó, A.M. Iron(III) Versatile Catalysts for Cycloaddition of CO2 to Epoxides and Epoxidation of Alkenes. ChemistrySelect 2016, 1, 2065–2070. [Google Scholar] [CrossRef]
  28. Yuan, Y.; Li, J.; Sun, X.; Li, G.; Liu, Y.; Verma, G.; Ma, S. Indium-Organic Frameworks Based on Dual Secondary Building Units Featuring Halogen-Decorated Channels for Highly Effective CO2 Fixation. Chem. Mater. 2019, 31, 1084–1091. [Google Scholar] [CrossRef]
  29. Li, J.; Ren, Y.; Yue, C.; Fan, Y.; Qi, C.; Jiang, H. Highly Stable Chiral Zirconium-Metallosalen Frameworks for CO2 Conversion and Asymmetric C-H Azidation. ACS Appl. Mater. Interfaces 2018, 10, 36047–36057. [Google Scholar] [CrossRef]
  30. Lu, B.-B.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. A Polyoxovanadate-Resorcin[4]arene-Based Porous Metal-Organic Framework as an Efficient Multifunctional Catalyst for the Cycloaddition of CO2 with Epoxides and the Selective Oxidation of Sulfides. Inorg. Chem. 2017, 56, 11710–11720. [Google Scholar] [CrossRef] [PubMed]
  31. Cao, J.-P.; Xue, Y.-S.; Li, N.-F.; Gong, J.-J.; Kang, R.-K.; Xu, Y. Lewis Acid Dominant Windmill-Shaped V8 Clusters: A Bifunctional Heterogeneous Catalyst for CO2 Cycloaddition and Oxidation of Sulfides. J. Am. Chem. Soc. 2019, 141, 19487–19497. [Google Scholar] [CrossRef]
  32. Zhao, Y.-Q.; Liu, Y.-Y.; Ma, J.-F. Polyoxometalate-Based Organic−Inorganic Hybrids as Heterogeneous Catalysts for Cycloaddition of CO2 with Epoxides and Oxidative Desulfurization Reactions. Cryst. Growth Des. 2021, 21, 1019–1027. [Google Scholar] [CrossRef]
  33. Drake, T.; Ji, P.; Lin, W. Site Isolation in Metal-Organic Frameworks Enables Novel Transition Metal Catalysis. ACC. Chem. Res. 2018, 51, 2129–2138. [Google Scholar] [CrossRef]
  34. Pei, W.-Y.; Xu, G.; Yang, J.; Wu, H.; Chen, B.; Zhou, W.; Ma, J.-F. Versatile Assembly of Metal-Coordinated Calix[4]resorcinarene Cavitands and Cages through Ancillary Linker Tuning. J. Am. Chem. Soc. 2017, 139, 7648–7656. [Google Scholar] [CrossRef]
  35. Wei, Y.-S.; Zhang, M.; Zou, R.; Xu, Q. Metal-Organic Framework-Based Catalysts with Single Metal Sites. Chem. Rev. 2020, 120, 12089–12174. [Google Scholar] [CrossRef] [PubMed]
  36. He, H.; Perman, J.A.; Zhu, G.; Ma, S. Metal-Organic Frameworks for CO2 Chemical Transformations. Small 2016, 12, 6309–6324. [Google Scholar] [CrossRef] [PubMed]
  37. Singh, G.; Lee, J.; Karakoti, A.; Bahadur, R.; Yi, J.; Zhao, D.; AlBahily, K.; Vinu, A. Emerging trends in porous materials for CO2 capture and conversion. Chem. Soc. Rev. 2020, 49, 4360–4404. [Google Scholar] [CrossRef]
  38. Lu, W.; Wei, Z.; Gu, Z.-Y.; Liu, T.-F.; Park, J.; Park, J.; Tian, J.; Zhang, M.; Zhang, Q.; Thomas, G., III; et al. Tuning the structure and function of metal-organic frameworks via linker design. Chem. Soc. Rev. 2014, 43, 5561–5593. [Google Scholar] [CrossRef]
  39. Jiang, W.; Yang, J.; Liu, Y.-Y.; Song, S.-Y.; Ma, J.-F. A porphyrin-based porous rtl metal-organic framework as an efficient catalyst for the cycloaddition of CO2 to epoxides. Chem. Eur. J. 2016, 22, 16991–16997. [Google Scholar] [CrossRef] [PubMed]
  40. Wu, H.; Yang, J.; Su, Z.-M.; Batten, S.R.; Ma, J.-F. An exceptional 54-fold interpenetrated coordination polymer with 103-srs network topology. J. Am. Chem. Soc. 2011, 133, 11406–11409. [Google Scholar] [CrossRef]
  41. Kalaj, M.; Cohen, S.M. Postsynthetic Modification: An Enabling Technology for the Advancement of Metal-Organic Frameworks. ACS Cent. Sci. 2020, 6, 1046–1057. [Google Scholar] [CrossRef] [PubMed]
  42. Lu, B.-B.; Chen, X.-Y.; Feng, C.-J.; Chang, J.; Ye, F. Palladium Nanoparticles Immobilized on a Resorcin[4]arene-Based Metal-Organic Framework for Hydrogenation of Nitroarenes. ACS Appl. Nano Mater. 2021, 4, 2278–2284. [Google Scholar] [CrossRef]
  43. Han, X.; Xu, Y.-X.; Yang, J.; Xu, X.; Li, C.-P.; Ma, J.-F. Metal-Assembled, Resorcin[4]arene-Based Molecular Trimer for Efficient Removal of Toxic Dichromate Pollutants and Knoevenagel Condensation Reaction. ACS Appl. Mater. Interfaces 2019, 11, 15591–15597. [Google Scholar] [CrossRef]
  44. Liu, J.-H.; Shen, Q.-T.; Yang, J.; Yu, M.-Y.; Ma, J.-F. Polyoxometalate-Templated Cobalt-Resorcin[4]arene Frameworks: Tunable Structure and Lithium-Ion Battery Performance. Inorg. Chem. 2021, 60, 3729–3740. [Google Scholar] [CrossRef] [PubMed]
  45. Su, K.; Wang, W.; Du, S.; Ji, C.; Zhou, M.; Yuan, D. Reticular Chemistry in the Construction of Porous Organic Cages. J. Am. Chem. Soc. 2020, 142, 18060–18072. [Google Scholar] [CrossRef]
  46. Lu, B.-B.; Yang, J.; Che, G.-B.; Pei, W.-Y.; Ma, J.-F. Highly Stable Copper(I)-Based Metal-Organic Framework Assembled with Resorcin[4]arene and Polyoxometalate for Efficient Heterogeneous Catalysis of Azide-Alkyne “Click” Reaction. ACS Appl. Mater. Interfaces 2018, 10, 2628–2636. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, H.; Yang, J.; Liu, Y.-Y.; Song, S.; Ma, J.-F. A Family of Metal-Organic Frameworks with A New Chair Conformation Resorcin[4]arene-Based Ligand: Selective Luminescent Sensing of Amine and Aldehyde Vapors, and Solvent-Mediated Structural Transformations. Cryst. Growth Des. 2016, 16, 3244–3255. [Google Scholar] [CrossRef]
  48. Lv, L.-L.; Yang, J.; Zhang, H.-M.; Liu, Y.-Y.; Ma, J.-F. Metal-Ion Exchange, Small-Molecule Sensing, Selective Dye Adsorption, and Reversible Iodine Uptake of Three Coordination Polymers Constructed by A New Resorcin[4]arene-Based Tetracarboxylate. Inorg. Chem. 2015, 54, 1744–1755. [Google Scholar] [CrossRef]
  49. Lu, B.-B.; Jiang, W.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. Resorcin[4]arene-Based Microporous Metal-Organic Framework as an Efficient Catalyst for CO2 Cycloaddition with Epoxides and Highly Selective Luminescent Sensing of Cr2O72−. ACS Appl. Mater. Interfaces 2017, 9, 39441–39449. [Google Scholar] [CrossRef]
  50. Sheldrick, G.M. SHELXS-2014, Program for the Crystal Structure Solution; University of Göttingen: Göttingen, Germany, 2014. [Google Scholar]
  51. Sheldrick, G.M. SHELXL-2014, Program for the Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 2014. [Google Scholar]
  52. Farrugia, L.J. WINGX: A Windows Program for Crystal Structure Analysis; University of Glasgow: Glasgow, UK, 1988. [Google Scholar]
  53. Spek, A.L. PLATON SQUEEZE: A tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Cryst. 2015, C71, 9–18. [Google Scholar]
  54. Hawxwell, S.M.; Brammer, L. Solvent hydrolysis leads to an unusual Cu(II) metal-organic framework. CrystEngComm 2006, 8, 473–476. [Google Scholar] [CrossRef]
  55. Murphy, C.A.; Cameron, T.S.; Cooke, M.W.; Aquino, M.A.S. Synthesis and structure of trans-dimethylammonium bis(oxalato)diaquaruthenate(III) tetrahydrate. Inorg. Chim. Acta 2000, 305, 225–229. [Google Scholar] [CrossRef]
  56. Li, Q.-Q.; Liu, H.; Zheng, T.-T.; Liu, P.; Song, J.-X.; Wang, Y.-Y. The effect of coordinated solvent molecules on metal coordination environments in singlecrystal-to-single-crystal transformations. CrystEngComm 2020, 22, 6750–6775. [Google Scholar] [CrossRef]
Scheme 1. Synthetic strategy for compounds 1 and 2.
Scheme 1. Synthetic strategy for compounds 1 and 2.
Crystals 11 00574 sch001
Figure 1. (a) Coordination environments around Co(II) in 1. Symmetry codes: (#1) −x − 2,−y + 2,−z − 1, (#2) −x − 1,−y + 1,−z, (#3) −x,−y + 1,−z, (#4) x + 1,y,z, (#5) x − 1,y,z, and (#6) −x − 1,−y + 2,−z. (b) Coordination of the L12− ligand. (c) The 1D channel of 1. (d) The 3D framework of 1 down the b-axis.
Figure 1. (a) Coordination environments around Co(II) in 1. Symmetry codes: (#1) −x − 2,−y + 2,−z − 1, (#2) −x − 1,−y + 1,−z, (#3) −x,−y + 1,−z, (#4) x + 1,y,z, (#5) x − 1,y,z, and (#6) −x − 1,−y + 2,−z. (b) Coordination of the L12− ligand. (c) The 1D channel of 1. (d) The 3D framework of 1 down the b-axis.
Crystals 11 00574 g001
Figure 2. (a) Coordination environments around Co(II) in 2. Symmetry codes: (#1) −x + 1,−y + 1,−z − 1, (#2) x − 1,y,z + 1, (#3) x − 1,y − 1,z + 1, (#4) −x,−y + 1,−z, (#5) x,y − 1,z, (#7) x + 1,y,z − 1, (#8) x,y + 1,z, and (#9) x + 1,y + 1,z − 1. (b) Coordination of the L12− ligand. The 3D structures of 2 down the a-axis (c) and c-axis (d).
Figure 2. (a) Coordination environments around Co(II) in 2. Symmetry codes: (#1) −x + 1,−y + 1,−z − 1, (#2) x − 1,y,z + 1, (#3) x − 1,y − 1,z + 1, (#4) −x,−y + 1,−z, (#5) x,y − 1,z, (#7) x + 1,y,z − 1, (#8) x,y + 1,z, and (#9) x + 1,y + 1,z − 1. (b) Coordination of the L12− ligand. The 3D structures of 2 down the a-axis (c) and c-axis (d).
Crystals 11 00574 g002
Figure 3. (a) Thermogravimetric curves of 1 and 2. PXRD analysis of (b) 1 and (c) 2.
Figure 3. (a) Thermogravimetric curves of 1 and 2. PXRD analysis of (b) 1 and (c) 2.
Crystals 11 00574 g003
Scheme 2. Coupling of CO2 with gylcidylphenylether.
Scheme 2. Coupling of CO2 with gylcidylphenylether.
Crystals 11 00574 sch002
Figure 4. Catalyst recycling test.
Figure 4. Catalyst recycling test.
Crystals 11 00574 g004
Table 1. X-ray crystal data and structure refinements parameters of 1 and 2.
Table 1. X-ray crystal data and structure refinements parameters of 1 and 2.
Parameters12
FormulaC80 H92 O48 N2 Co5C82 H86 O48 N2 Co6
Mr2144.202221.10
Cryst systTriclinicTriclinic
Space groupP-1P-1
a (Å)10.5320(6)11.2564(6)
b (Å)13.2619(7)15.9785(9)
c (Å)18.1118(10)16.5583(10)
α (°)70.883(5)62.347(6)
β (°)74.056(5)73.527(5)
γ (°)85.860(4)70.219(5)
V (Å3)2297.9(2)2453.6(3)
Z11
Dcalc (g cm−3)1.5501.503
F(000)11051138
Rint0.04980.0446
GOF on F21.2111.175
R1 a [I > 2σ(I)]0.08470.0663
wR2 b (all data)0.18830.1559
aR1 = Σ||Fo| − |Fc|||Fo|. b wR2 = {Σ[w(Fo2 − Fc2)2]/Σw(Fo2)2]}1/2.
Table 2. Coupling of CO2 with gylcidylphenylether under different conditions a.
Table 2. Coupling of CO2 with gylcidylphenylether under different conditions a.
Entry 1 (mg)Temperature (°C)Time (h)Conversion (%) b
11080124
22080126
33080148
4080123
5302510
63050112
73080251
83080480
93080683
103080898
a Reaction conditions: gylcidylphenylether (5.00 mmol, 0.75 g), n-Bu4NBr (0.50 mmol 0.16 g), and CO2 (1 atm). b Isolated conversions were calculated by 1H NMR.
Table 3. Coupling of CO2 with different epoxides a.
Table 3. Coupling of CO2 with different epoxides a.
EntryEpoxidesProductsConversion (%) b
1 Crystals 11 00574 i001 Crystals 11 00574 i00299
2 Crystals 11 00574 i003 Crystals 11 00574 i00499
3 Crystals 11 00574 i005 Crystals 11 00574 i00699
4 Crystals 11 00574 i007 Crystals 11 00574 i00895
5 Crystals 11 00574 i009 Crystals 11 00574 i01075
6 Crystals 11 00574 i011 Crystals 11 00574 i01293
a Reaction conditions: epoxides (5.00 mmol), CO2 pressure (1 atm), activated catalyst 1 (30 mg, 0.07 mmol based on CoII cations), n-Bu4NBr (0.16 g, 0.50 mmol), 8 h, and 80 °C. b Isolated conversions were calculated by 1H NMR.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lu, B.-B.; Han, X.; Feng, C.-J.; Wang, D.; Ye, F. Two Co(II)-Based MOFs Constructed from Resorcin[4]Arene Ligand: Syntheses, Structures, and Heterogeneous Catalyst for Conversion of CO2. Crystals 2021, 11, 574. https://doi.org/10.3390/cryst11060574

AMA Style

Lu B-B, Han X, Feng C-J, Wang D, Ye F. Two Co(II)-Based MOFs Constructed from Resorcin[4]Arene Ligand: Syntheses, Structures, and Heterogeneous Catalyst for Conversion of CO2. Crystals. 2021; 11(6):574. https://doi.org/10.3390/cryst11060574

Chicago/Turabian Style

Lu, Bing-Bing, Xue Han, Cheng-Jie Feng, Duo Wang, and Fei Ye. 2021. "Two Co(II)-Based MOFs Constructed from Resorcin[4]Arene Ligand: Syntheses, Structures, and Heterogeneous Catalyst for Conversion of CO2" Crystals 11, no. 6: 574. https://doi.org/10.3390/cryst11060574

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