Construction of two metal-organic frameworks by rigid bis(triazole) and carboxylate mixed-ligands and their catalytic properties for CO2 cycloaddition reaction

Weichen WANG Chunhua GONG Junyong ZHANG Yanfeng BI Hao XU Jingli XIE

Citation:  Weichen WANG, Chunhua GONG, Junyong ZHANG, Yanfeng BI, Hao XU, Jingli XIE. Construction of two metal-organic frameworks by rigid bis(triazole) and carboxylate mixed-ligands and their catalytic properties for CO2 cycloaddition reaction[J]. Chinese Journal of Inorganic Chemistry, 2024, 40(7): 1377-1386. doi: 10.11862/CJIC.20230415 shu

刚性双(三氮唑)和羧酸混合配体构建的两种金属有机骨架材料及其在CO2环加成反应中的催化性质

    通讯作者: 宫春华, gongch@jxnhu.edu.cn
    毕研峰, biyanfeng@lnpu.edu.cn
    谢景力, jlxie@mail.zjxu.edu.cn
  • 基金项目:

    国家自然科学基金 21771088

摘要: 通过酸碱混合配体策略合成了2例含刚性双三氮唑配体的金属有机骨架(MOF)材料:{[Zn2(L)(TP)2(H2O)·H2O]}n (1)和[Zn(L)(HTMA)]n (2),其中L=4,4'-(3,3'-dimethyl-(1,1'-biphenyl)-4,4'-diyl)bis(4H-1,2,4-triazole),H2TP=对苯二甲酸,H3TMA=1,3,5-均苯三甲酸。用单晶X射线衍射表征其结构。结构分析表明,MOF 1显示出3,6-双节点的二维结构,其拓扑符号为(42·6)2(48·66·8),MOF 2呈现为经典的sql二维拓扑结构。在温和条件下,2对CO2与环氧化物的环加成反应具有优异的催化活性,且重复使用至少3次后仍然保持其催化性能。

English

  • Recently, the development of feasible carbon dioxide capture and storage/sequestration technologies has received widespread attention to reduce greenhouse gas emissions[1-8]. Also, CO2 is considered to be a more useful and environmentally friendly source of C1 than other C1 feedstocks in industry[9-14]. Besides capturing and storing atmospheric CO2 by using functional nanoporous materials such as activated carbons[15-16], zeolites[17], metal-organic frameworks (MOFs) [18-20], and porous organic polymers[21], the direct cycloaddition of CO2 with epoxides into value-added chemicals such as cyclic carbonate may be a more attractive strategy for reducing and utilizing CO2[8, 22]. Although many hetero-geneous catalysts such as zeolite, metal oxide, and functional graphene have been applied to carbon dioxide cycloaddition reaction system, most catalysts have low catalytic activity, and the reaction conditions often require high temperature (> 100 ℃) and high pressure (> 3 MPa)[23-27]. Therefore, it is pertinent to explore new and efficient heterogeneous catalysts that can effectively convert CO2 under mild conditions.

    As a novel type of heterogeneous catalysts, MOFs have been widely explored in the cycloaddition of CO2 with epoxides, due to their specific advantages such as high specific surface area, and multifunctional structure[28-36]. However, MOF catalysts still have various practical problems in CO2 cycloaddition, such as low activity, poor stability, difficult recycling, and harsh reaction conditions, which directly put the hurdle for the application of MOF materials[37-38]. Therefore, it is imperative to explore stable, efficient, and recyclable MOF catalysts, especially those capable of converting CO2 into cyclic carbonate under mild conditions.

    In our previous work, a new 2D Ag-MOF with the open Lewis acid silver site was obtained, and it could effectively promote the CO 2 cycloaddition reaction of N-methylpropargylamine to generate the product of 3-methyl-5-methyleneoxazolidin-2-one under relatively mild conditions (75 ℃ and 101.325 kPa) [39]. Also, the Ag-MOF can be recycled at least three times and still maintain the stability of the structure. Following this research line, a rigid bis(triazole) ligand (L) has been combined with Zn2+ to achieve two novel MOFs by using mixed-ligand strategy in the presence of aromatic carboxylic acid ligands, namely {[Zn2(L) (TP)2(H2O)·H2O]}n (1) and [Zn(L)(HTMA)]n (2), where L=4, 4′-(3, 3′-dimethyl-(1, 1′-biphenyl)-4, 4′-diyl)bis(4H-1, 2, 4-triazole), H2TP=terephthalic acid, H3TMA=1, 3, 5-benzenedicarboxylic acid (Scheme 1). Structural analysis reveals that 1 displays a 3, 6-connected 2D structure with a new topological point symbol of (42·6)2(48·66·8), while 2 presents a 2D sql topology structure. Interestingly, due to the good stability and open Lewis acid Zn sites, 2 exhibits efficient catalytic activity for CO2 conversion into cyclic carbonates under mild conditions. Moreover, 2 can be recycled at least three times without significant loss of catalytic activity and retain the original frame structure.

    Scheme1

    Scheme1.  Structures of ligand L (base) and different auxiliary ligands (acid)

    All the chemicals were received as reagent grade and used without further purification. IR spectrum was recorded on a Varian 640 FT/IR spectrometer with KBr powder in a 4 000-500 cm-1 range. Powder X-ray diffraction (PXRD) patterns were collected on a DX-2600 spectrometer with Mo radiation (λ=0.071 073 nm) at 296 K (U=40 kV, I=30 mA, 2 θ=5°-50°). Thermal gravimetric analyses (TGA) were carried out under N2 flow on an SDT 2960 differential thermal analyzer at a heating rate of 10 ℃·min-1. The concentration of leaching zinc ions was determined by the ICP-MS of Agilent 7800. 1H NMR spectrum was recorded on a Varian (400 MHz) spectrometer. And the conversion was calculated by 1H NMR.

    Zn(NO3)2·6H2O (14.8 mg, 0.05 mmol), H2TP (8.5 mg 0.05 mmol), L (15.8 mg 0.05 mmol) were dissolved in a mixed solution of DMF/H2O (1 mL/2 mL) at room temperature. The resulting solution was sonicated for 30 min, sealed in a Teflon-lined autoclave under autogenous pressure, heated to 80 ℃, and maintained for 2 d. After slowly cooling down to room temperature, colorless crystals of MOF 1 were collected and washed three times with distilled water. Yield: 63.2%, based on Zn(NO3)2·6H2O). Anal. Calcd. for C34H28Zn2N6O10(%): C, 50.33; H, 3.48; N, 10.36. Found(%): C, 50.24; H, 3.61; N, 10.23. IR (KBr powder, cm-1): 3 392(m), 2 918 (w), 1 614(m), 1 574(s), 1 512(w), 1 385(s), 1 299(m), 1 033(m), 848(m), 755(s), 580(m).

    Colorless block crystals of MOF 2 were obtained through a similar procedure by using H3TMA (7.80 mg, 0.05 mmol) instead. Yield: 75.6%, based on Zn(NO3)2·6H2O). Anal. Calcd. for C27H20ZnN6O6(%): C, 54.98; H, 3.42; N, 14.25. Found(%): C, 54.26; H, 3.78; N, 13.95. IR (KBr powder, cm-1): 3 455(m), 3 162(w), 2 923(w), 1 715(vs), 1 574(s), 1 447(m), 1 353(vs), 1 232(m), 1 038 (m), 883(m), 729(m).

    The single-crystal X-ray diffraction data of MOF 1 were collected on a Bruker D8 QUEST instrument equipped with graphite-monochromated Cu (λ = 0.154 184 nm) at 296(2) K radiation. The suitable crystal of MOF 2 was selected and characterized on an Oxford Diffraction Gemini R Ultra diffractometer with graphite-monochromated Cu (λ =0.154 184 nm) at 296(2) K. The structures were solved by SHELXS (direct methods) and refined by SHELXL (full-matrix least-squares techniques) in the Olex2 package. The topological structures of all complexes are calculated by the TOPOS 4.0 program package[40]. Crystal data and structure refinements of MOFs 1 and 2 are provided in Table S1 (Supporting information). Selected bond lengths and angles are listed in Table S2 and S3.

    Single-crystal X-ray diffraction analysis reveals that MOF 1 crystallizes in the orthorhombic Pbca space group. The asymmetric unit of 1 contained two Zn(Ⅱ) ions, one L ligand, two TP2- ligands, a coordinated water molecule, and a lattice water molecule. There are two independent Zn(Ⅱ) ions (Zn1 and Zn2) with different coordination environments in 1. The Zn1 ion is four-coordinated by two oxygen atoms from two different TP2- ligands, a nitrogen atom from the L ligand, and an oxygen atom from the coordinated water molecule to form a distorted tetrahedron configuration. The Zn2 ion is five-coordinated by three oxygen atoms from two different TP2- ligands, and two nitrogen atoms from two different L ligands to form a trigonal bipyramid (Fig. 1a). The bond lengths of Zn—O range from 0.193 3 to 0.239 7 nm and the bond lengths of Zn—N range from 0.204 6 to 0.211 6 nm, respectively. The carboxylate groups of the deprotonated TP2- ligands display chelating bidentate and monodentate coordination modes. The L ligand has only one coordination mode and coordinates with three Zn ions. Zn1 and Zn2 in the asymmetric unit are bridged by the L ligand. The adjacent TP2- ligands are linked to form two different 1D chains (Fig. 1b and 1c). The coordinated L ligands are fixed between the two chains to assemble a 2D layered network (Fig. 1d). Simplify the structure of 1 to get its topological network structure (Fig. 2), in which each [Zn2 N4] unit is regarded as the 6-connected node, and all Zn1 are considered as 3-connected nodes. As a result, the framework is an unprecedented 3, 6-connected network with the Schläfli symbol (42·6)2(48·66·8) by software TOPOS.

    Figure 1

    Figure 1.  Structure of MOF 1: (a) coordination environment of Zn2+ ion; (b, c) two different 1D chains; (d) 2D layer

    Symmetry codes: 0.5-x, 0.5+y, z; 1.5-x, -0.5+y, z; 1-x, 1-y, z.

    Figure 2

    Figure 2.  Topological simplification diagram of MOF 1

    Single-crystal X-ray diffraction analysis reveals that MOF 2 crystallizes in the monoclinic P21/c space group. The asymmetric unit of 2 contained a Zn(Ⅱ) ion, an L ligand, and an HTMA2- ligand. The Zn(Ⅱ) ion is four-coordinated with a [ZnO2N2] coordination environ-ment and displays a tetrahedron configuration, consist-ing of two oxygen atoms (O1, O4) from two HTMA2- ligands and two nitrogen atoms (N1, N4) from two L ligands (Fig. 3a). The bond lengths of Zn—O range from 0.194 0 to 0.195 1 nm and the bond lengths of Zn—N range from 0.203 0 to 0.203 4 nm, respectively. The carboxylate groups of deprotonated HTMA2- form 1D chains with Zn2+ ions by single-dentate coordination (Fig. 3b). The coordinated L ligands are fixed between the two chains to assemble a 2D wave-type layered network (Fig. 3c). The structure of 2 is simplified by treating L and HTMA2- as linkers, which can be viewed as a sql topology (Fig. 4).

    Figure 3

    Figure 3.  Structure of MOF 2: (a) coordination environment of Zn2+ ion; (b) 1D chain;(c) 2D layer

    Symmetry codes: 1-x, y, z; 1+x, 0.5+y, 0.5+z.

    Figure 4

    Figure 4.  Topological simplification diagram of MOF 2

    The simulated powder X-ray diffraction and experimental results of MOFs 1 and 2 are shown in Fig.S3. The main peak positions of the experimental patterns of 1 and 2 were consistent with their simulated patterns, demonstrating the single-phase purity of the products.

    IR spectra of MOFs 1 and 2 are shown in Fig.S4. The IR spectra of the MOFs showed broad bands at about 3 392 cm-1 for 1, and 3 455 cm-1 for 2, which can be ascribed to the O—H stretching vibration of the water molecules. No absorption peak around 1 700 cm-1 for —COOH was observed for 1, indicating that all the carboxyl groups of H2TP are deprotonated. By contrast, for 2, the presence of a band at 1 715 cm-1 indicates that the HTMA2- is protonated. The characteristic peaks of methyl groups were at 2 918 cm-1 for 1 and 2 923 cm-1 for 2. Characteristic bands at 1 512 cm-1 for 1 and 1 537 cm-1 for 2 are assigned to the νC=N stretching vibration of 1, 2, 4-triazole. The strong characteristic bands at 1 574, 1 385 cm-1 for 1 and 1 574, 1 433 cm-1 for 2 can be considered as the asymmetric and symmetric vibrations of the carboxyl groups, respectively. These IR spectra are in agreement with the results of elemental analysis and single-crystal X-ray diffraction results.

    The TGA experiments for MOFs 1 and 2 were carried out in a range of 25-800 ℃. As shown in Fig.S5a, the TG curve of 1 displayed two-step thermal decomposing behaviours. The first weight loss of 4.39% (25-285 ℃) is due to the loss of one lattice H2O and one coordinated water molecule (Calcd. 4.44%). After 340 ℃, the skeletons of 1 began to collapse because of the decomposition of the organic ligands. As shown in Fig.S5b, the TG curve of 2 displayed high thermal stability and one-step thermal decomposing behaviours. As the temperature rose to 360 ℃, the skeletons of 2 began to collapse because of the decomposition of organic ligands. Thus, these results reveal that MOFs 1 and 2 have good thermal stability.

    Crystal structures of MOFs 1 and 2 show quite a few interesting features, for example, (ⅰ)the presence of Lewis acidic Zn2+ metal sites in their framework, which may activate the substrate, (ⅱ)the presence of channels that enable free diffusion of substrates to the catalytic sites, (ⅲ)the presence of a π-electron-rich framework architecture that may strengthen the structure and could facilitate the interactions between the catalytic framework and substrates/reagents. These properties provide the basis for exploring their catalytic activity for the reaction of CO2 cycloaddition with epoxide. Initially, by using propylene oxide and CO2 as model substrates for the studied cycloaddition reaction, 2 was tried as the catalyst to test its activity. By optimizing the reaction conditions, we have examined the influence of catalyst loading, the importance of cocatalyst, and reaction time under mild conditions (atmospheric pressure and room temperature) (Table 1).

    Table 1

    Table 1.  Orthogonal experiments of CO2 cycloaddition reactionsa
    下载: 导出CSV
    Entry xcatalyst/% xTBAB/% Time/h Conversionb/%
    1 0.1 0 24 0
    2 0 8.0 24 32.48
    3 0.1 1.0 24 30.30
    4 0.1 3.0 24 37.88
    5 0.1 5.0 24 52.36
    6 0.1 6.0 24 66.67
    7 0.1 7.0 24 89.30
    8 0.1 8.0 24 > 99
    9 0.08 8.0 24 69.44
    10 0.05 8.0 24 47.85
    11 0.1 8.0 18 67.57
    12 0.1 8.0 12 38.91
    13 0.1 8.0 6 18.42
    anepoxide=20 mmol; bThe conversion of the epoxide was determined by 1H NMR analysis (Fig.S8-S21).

    The reaction was conducted in a 10 mL Schlenk tube with 20 mmol of propylene oxide and CO2 under atmospheric pressure at room temperature. In the presence of 8.0% of co-catalyst (TBAB), 0.1% loading of catalyst 2 afforded above 99% maximum conversion within 24 h (Table 1, entry 8). When the reactions were performed in the absence of TBAB or 2, no substrates or only a little were converted (Table 1, entries 1 and 2). Therefore, it was assumed that the synergistic effect of 2 and TBAB should be responsible for the high catalytic activity of the system under mild conditions. Under the same condition, when 1 was used as a catalyst, the conversion of propylene oxide was 89.28% (Table 2, entry 1). The different coordination geometry of Lewis acid Zn metal centers in 1 and 2 lead to certain differences in their catalytic effects for CO2 cycloaddition. Additionally, control experiments indicated that other standard zinc catalysts (ZnCl2 and Zn(NO3)2·6H2O) provided low conversions with the same amount of catalysts (Table 2, entries 3 and 4).

    Table 2

    Table 2.  Catalytic activity of various catalysts for propylene oxidea
    下载: 导出CSV
    Entry Catalyst Epoxide Product Conversionb/% TONc
    1 1 89.28 892.8
    2 2 100.0 1 000.0
    3 ZnCl2 51.55 515.5
    4 Zn(NO3)2·6H2O 52.36 523.6
    a Reaction conditions: propylene oxide (20 mmol), xcatalyst=0.1%, xTBAB=8%, RT, atmospheric pressure (balloon), 24 h; b The conversion of the epoxide was determined by 1H NMR analysis; c TON=nproduct/ncatalyst (Fig.S22-S24).

    To explore the potential generality of the catalytic system, several typical epoxide substrates (i.e., 1-chloro-2, 3-epoxypropane, 1-bromo-2, 3-epoxypropane, and styrene oxide) were examined in this cycloaddition reaction. Under the optimized conditions, the conversions of the three substrates were 57.80%, 55.87%, and 14.53% (Table 3, entries 2-4), respectively. The lower catalytic efficiency may be associated with the large size of the substrates, resulting in ineffective binding with vacant zinc sites.

    Table 3

    Table 3.  Cycloaddition reactions of CO2 and various epoxides with catalyst 2a
    下载: 导出CSV
    Entry Epoxide Product Conversionb/% TONc
    1 100.0 10 00.0
    2 57.80 578.0
    3 55.87 558.7
    4 14.53 145.3
    aReaction conditions: epoxide (20 mmol), xcatalyst=0.1%, xTBAB=8%, RT, atmospheric pressure (balloon), 24 h; b The con-version of the epoxide was determined by 1H NMR analysis; c TON=nproduct/ncatalyst (Fig.S25-Fig.S27).

    To investigate the heterogeneous nature of the zinc catalyst, MOF 2 was filtered from the reaction mixture after 6 h. The resulting supernatant continued to react for another 18 h, and there were no observable catalytic conversions (Fig. S6). Additionally, the catalyst leakage in collected supernatant was tested by the ICP-MS analysis, and the result showed only 0.25 mg·L-1 of free Zn ion, demonstrating that there was almost no leaching of Zn ion during the reaction process. Those results showed that 2 is a heterogeneous catalyst in the catalytic system. As shown in Fig.S7a, to evaluate the recyclability of 2, experiments were performed by using the recovered catalyst of 2. The obtained solid can still maintain 95.23% conversion after three recycles. The PXRD results revealed that the catalyst keeps its crystallinity after catalytic experiments (Fig.S7b).

    The catalytic mechanism was further inspected based on literature reports (Fig. 5) [36-37, 41-42]. Firstly, epoxide interacts with the Lewis acid Zn2+ center via the oxygen atom and activates the ternary ring. Then, the epoxide opens the ring with the help of the Zn2+ sites to form the O···Zn adducts (Ⅰ). Secondly, the Br- produced by TBAB attacks as nucleophiles on the sterically less-crowded carbon atom to destroy the terpolymer epoxy ring (Ⅱ). In the third step, CO2 is inserted and forms the alkyl carbonate anion (Ⅲ). Subsequently, the intra-molecular ring closes to produce a cyclic carbonate product (Ⅳ). Finally, the epoxy carbonate breaks away from the Lewis acid Zn2+ center and reproduces the catalyst, thus allowing the CO2 cycloaddition reaction to enter the next catalytic cycle.

    Figure 5

    Figure 5.  Plausible mechanism for the cycloaddition of epoxides with CO2 catalyzed by MOF 2

    In summary, two novel metal-organic frameworks have been prepared by using a mixed-ligand strategy and structurally characterized. The presence of different acidic ligands results in two MOFs with interesting topologies. MOF 1 displays a 3, 6-connected 2D structure with a new topological point symbol of (42·6)2(48·66·8), while MOF 2 presents a 2D sql topology. Remarkably, 1 and 2 were exploited as catalytic materials in the fixation of CO2 to cyclic carbonates under mild conditions (RT and ambient pressure). In particular, 2 has the highest catalytic efficiency (above 99%) in 24 h for cycloaddition of propylene oxide, and its structure remains stable after three cycles. In addition, 2 has certain catalytic activity for other epoxides such as 1-chloro-2, 3-epoxypropane, 1-bromo-2, 3-epoxypropane, and styrene oxide, demonstrating its generality for the catalytic reaction. This work provides a further step towards the efficient use of CO2 under mild condi-tions and lays a foundation for achieving more heteroge-neous catalysts in this particular area.

    Supporting information is available at http://www.wjhxxb.cn


    1. [1]

      Jacobson M Z. Review of solutions to global warming, air pollution, and energy security[J]. Energy Environ. Sci., 2009, 2:  148-173. doi: 10.1039/B809990C

    2. [2]

      Jones W D. Carbon capture and conversion[J]. J. Am. Chem. Soc., 2020, 142:  4955-4957. doi: 10.1021/jacs.0c02356

    3. [3]

      Zou Y H, Huang Y B, Si D H, Yin Q, Wu Q J, Weng Z X, Cao R. Porous metal-organic framework liquids for enhanced CO2 adsorption and catalytic conversion[J]. Angew. Chem. Int. Ed., 2021, 60(38):  20915-20920. doi: 10.1002/anie.202107156

    4. [4]

      Li J, Ma Y G, McCarthy M C, Sculley J, Yu J L, Jeong H, Balbuena P B, Zhou H. Carbon dioxide capture-related gas adsorption and separation in metal-organic frameworks[J]. Coord. Chem. Rev., 2011, 255:  1791-1823. doi: 10.1016/j.ccr.2011.02.012

    5. [5]

      Wu Q J, Si D H, Sun P P, Dong S Z, Chen Q, Ye S H, Sun D, Cao R, Huang Y B. Atomically precise copper nanoclusters for highly efficient electroreduction of CO2 towards hydrocarbons via breaking the coordination symmetry of Cu site[J]. Angew. Chem. Int. Ed., 2023, 62(36):  e2023068.

    6. [6]

      Liang J, Xie Y Q, Wang X S, Wang Q, Liu T T, Huang Y B, Cao R. An imidazolium-functionalized mesoporous cationic metal-organic framework for cooperative CO2 fixation into cyclic carbonate[J]. Chem. Commun., 2018, 54:  342-345. doi: 10.1039/C7CC08630J

    7. [7]

      Trickett C A, Helal A, Al-Maythalony B A, Yamani Z H, Cordova K E, Yaghi O M. The chemistry of metal-organic frameworks for CO2 capture, regeneration and conversion[J]. Nat. Rev. Mater., 2017, 2(8):  17045. doi: 10.1038/natrevmats.2017.45

    8. [8]

      Guo F, Zhang X L. Metal-organic frameworks for the energy-related conversion of CO2 into cyclic carbonates[J]. Dalton Trans., 2020, 49:  9935-9947. doi: 10.1039/D0DT01516D

    9. [9]

      Tian D W, Liu B Y, Gan Q Y, Li H R, Darensbourg D J. Formation of cyclic carbonates from carbon dioxide and epoxides coupling reactions efficiently catalyzed by robust, recyclable one-component aluminum-salen complexes[J]. ACS Catal., 2012, 2:  2029-2035. doi: 10.1021/cs300462r

    10. [10]

      Yang H M, Zhang X, Zhang G Y, Fei H H. An alkaline-resistant Ag(Ⅱ)-anchored pyrazolate-based metal-organic framework for chemical fixation of CO2[J]. Chem. Commun., 2018, 54:  4469-4472. doi: 10.1039/C8CC01461B

    11. [11]

      Liu X H, Ma J G, Niu Z, Yang G M, Cheng P. An efficient nanoscale heterogeneous catalyst for the capture and conversion of carbon dioxide at ambient pressure[J]. Angew. Chem. Int. Ed., 2015, 127:  1002-1005. doi: 10.1002/ange.201409103

    12. [12]

      Grignard B, Gennen S, Jerome C, Kleij A W. Detrembleur C. Advances in the use of CO2 as a renewable feedstock for the synthesis of polymers[J]. Chem. Soc. Rev., 2019, 48:  4466-4514.

    13. [13]

      赵丹, 廖再添, 张旺, 陈治洲, 孙为银. 功能化金属有机框架材料催化二氧化碳转化研究进展[J]. 无机化学学报, 2021,37,(7): 1153-1176. ZHAO D, LIAO Z T, ZHANG W, CHEN Z Z, SUN W Y. Progress in functional metal-organic frameworks for catalytic conversion of carbon dioxide[J]. Chinese J. Inorg. Chem., 2021, 37(7):  1153-1176.

    14. [14]

      Hanusch J M, Kerschgens I P, Huber F, Neuburger M, Gademann K. Pyrrolizidines for direct air capture and CO2 conversion[J]. Chem. Commun., 2019, 55:  949-952. doi: 10.1039/C8CC08574A

    15. [15]

      Zhang Y M, Li B Y, Williams K, Gao W Y, Ma S Q. A new microporous carbon material synthesized via thermolysis of a porous aromatic framework embedded with an extra carbon source for low-pressure CO2 uptake[J]. Chem. Commun., 2013, 49:  10269-10271. doi: 10.1039/c3cc45252b

    16. [16]

      Bae Y, Snurr R Q. Development and evaluation of porous materials for carbon dioxide separation and capture[J]. Angew. Chem. Int. Ed., 2011, 50:  11586-11596. doi: 10.1002/anie.201101891

    17. [17]

      Hudson M R, Queen W L, Mason J A, Fickel D W, Lobo R F, Brown C M. Unconventional, highly selective CO2 adsorption in zeolite SSZ1-3[J]. J. Am. Chem. Soc., 2012, 134:  1970-1973. doi: 10.1021/ja210580b

    18. [18]

      Gao W Y, Chen Y, Niu Y H, Williams K, Cash L, Perez P J, Wojtas L, Cai J F, Chen Y S, Ma S Q. Crystal engineering of an nbo topology metal-organic framework for chemical fixation of CO2 under ambient conditions[J]. Angew. Chem. Int. Ed., 2014, 53:  2615-2619. doi: 10.1002/anie.201309778

    19. [19]

      Xiang S C, He Y B, Zhang Z J, Wu H, Zhou W, Krishna R, Chen B L. Microporous metal-organic framework with potential for carbon dioxide capture at ambient conditions[J]. Nat. Commun., 2012, 3:  954. doi: 10.1038/ncomms1956

    20. [20]

      Nugent P, Belmabkhout Y, Burd S D, Cairns A J, Luebke R, Forrest K, Pham T, Ma S, Space B, Wojtas L, Eddaoudi M, Zaworotko M J. Porous materials with optimal adsorption thermodynamics and kinetics for CO2 separation[J]. Nature, 2013, 495:  80-84. doi: 10.1038/nature11893

    21. [21]

      Lu W, Sculley J P, Yuan D Q, Krishna R, Wei Z W, Zhou H C. Polyamine-tethered porous polymer networks for carbon dioxide capture from flue gas[J]. Angew. Chem. Int. Ed., 2012, 51:  7480-7484. doi: 10.1002/anie.201202176

    22. [22]

      Choi J C, He L N, Yasuda H Y, Sakakura T. Selective and high yield synthesis of dimethyl carbonate directly from carbon dioxide and methanol[J]. Green Chem., 2002, 4:  230-234. doi: 10.1039/b200623p

    23. [23]

      Xie Y, Wang T T, Liu X H, Zou K, Deng W Q. Capture and conversion of CO2 at ambient conditions by a conjugated microporous polymer[J]. Nat. Commun., 2013, 4:  1960. doi: 10.1038/ncomms2960

    24. [24]

      Lin S, Diercks C S, Zhang Y B, Kornienko N, Nichols E M, Zhao Y B, Paris A R, Kim D, Yang P D, Yaghi O M, Chang C J. Covalent organic frameworks comprising cobalt porphyrins for catalytic CO2 reduction in water[J]. Science, 2015, 349:  1208-1213. doi: 10.1126/science.aac8343

    25. [25]

      Yamaguchi K, Ebitani K, Yoshida T, Yoshida H, Kaneda K. Mg-Al mixed oxides as highly active acid-base catalysts for cycloaddition of carbon dioxide to epoxides[J]. J. Am. Chem. Soc., 1999, 121:  4526-4527. doi: 10.1021/ja9902165

    26. [26]

      Yano T, Matsui H, Koike T, Ishiguro H, Fujihara H, Yoshihara M, Maeshima T. Magnesium oxide-catalysed reaction of carbon dioxide with an epoxide with retention of stereochemistry[J]. Chem. Commun., 1997, 12:  1129-1130.

    27. [27]

      Tang L, Zhang S B, Wu Q L, Wang X R, Wu H, Jiang Z Y. Heterobimetallic metal-organic framework nanocages as highly efficient catalysts for CO2 conversion under mild conditions[J]. J. Mater. Chem. A, 2018, 6:  2964-2973. doi: 10.1039/C7TA09082J

    28. [28]

      Li P Z, Wang X J, Liu J, Lim J S, Zou R Q, Zhao Y L. A triazole-containing metal-organic framework as a highly effective and substrate size-dependent catalyst for CO2 conversion[J]. J. Am. Chem. Soc., 2016, 138:  2142-2145. doi: 10.1021/jacs.5b13335

    29. [29]

      潘伟, 马驰枭, 周川江, 张俊勇, 石彦波, 徐昊, 朱敦如, 谢景力. 基于2, 6-二(4-羧基苯亚甲基)环己酮的金属-有机框架化合物的合成与表征[J]. 无机化学学报, 2021,37,(5): 953-960. PAN W, MA C X, ZHOU C J, ZHANG J Y, SHI Y B, XU H, ZHU D R, XIE J L. Synthesis and characterization of metalorganic framework based on 2, 6-bis (4carboxybenzylidene) cyclo-hexanone[J]. Chinese J. Inorg. Chem., 2021, 37(5):  953-960.

    30. [30]

      Song J L, Zhang Z F, Hu S Q, Wu T B, Jiang T, Han B X. MOF-5/n-Bu4NBr: An efficient catalyst system for the synthesis of cyclic carbonates from epoxides and CO2 under mild conditions[J]. Green Chem., 2009, 11:  1031-1036. doi: 10.1039/b902550b

    31. [31]

      Kim J, Kim S, Jang H G, Seo G, Ahn W. CO2 Cycloaddition of styrene oxide over MOF catalysts[J]. Appl. Catal. A-Gen., 2013, 453:  175180.

    32. [32]

      Liang J, Chen R P, Wang X Y, Liu T T, Wang X S, Huang Y B, Cao R. Postsynthetic ionization of an imidazole-containing metal-organic framework for the cycloaddition of carbon dioxide and epoxides[J]. Chem. Sci., 2017, 8:  1570-1575. doi: 10.1039/C6SC04357G

    33. [33]

      Xue Z M, Jiang J Y, Ma M G, Li M F, Mu T C. Gadolinium-based metal-organic framework as an efficient and heterogeneous catalyst to activate epoxides for cycloaddition of CO2 and alcoholysis[J]. ACS Sustain. Chem. Eng., 2017, 5:  2623-2631. doi: 10.1021/acssuschemeng.6b02972

    34. [34]

      Parmar B, Patel P, Kureshy R I, Khan N H, Suresh E. Sustainable heterogeneous catalysts for CO2 utilization by using dual ligand Zn/Cd metal-organic frameworks[J]. Chem.-Eur. J., 2018, 24:  15831-15839. doi: 10.1002/chem.201802387

    35. [35]

      He H M, Sun Q, Gao W Y, Perman J A, Sun F X, Zhu G S, Aguila B, Forrest K, Space B, Ma S Q. A stable metal-organic framework featuring a local buffer environment for carbon dioxide fixation[J]. Angew. Chem. Int. Ed., 2018, 57:  4657-4662. doi: 10.1002/anie.201801122

    36. [36]

      Rani P, Husain A, Bhasin K K, Kumar G. Metal-organic framework-based selective molecular recognition of organic amines and fixation of CO2 into cyclic carbonates[J]. Inorg. Chem., 2022, 61:  6977-6994. doi: 10.1021/acs.inorgchem.2c00367

    37. [37]

      Sun X D, Gu J M, Yuan Y, Yu C Y, Li J T, Shan H Y, Li G H, Liu Y L. A stable mesoporous Zr-Based metal organic framework for highly efficient CO2 conversion[J]. Inorg. Chem., 2019, 58:  7480-7487. doi: 10.1021/acs.inorgchem.9b00701

    38. [38]

      Das R, Ezhil T, Nagaraja C M. Design of bifunctional zinc(Ⅱ)-organic framework for efficient coupling of CO2 with terminal/internal epoxides under mild conditions[J]. Cryst. Growth Des., 2022, 22:  598-607. doi: 10.1021/acs.cgd.1c01148

    39. [39]

      Ma C X, Pan W, Zhang J Y, Zeng X H, Gong C H, Xu H T, Shen R P, Zhu D R, Xie J L. Metal-organic frameworks derived from chalcone dicarboxylic acid: New topological characters and initial catalytic properties[J]. Inorg. Chim. Acta, 2022, 543:  121166. doi: 10.1016/j.ica.2022.121166

    40. [40]

      Blatov V A, Ilyushin G D, Blatova O A, Anurova N A, Ivanov-Schits A K, Dem'Yanets L N. Analysis of migration paths in fast-ion conductors with voronoi-dirichlet partition[J]. Acta Crystallogr. Sect. B, 2006, B62:  1010-1018.

    41. [41]

      Liang J, Xie Y Q, Wu Q, Wang X Y, Liu T T, Li H F, Huang Y B, Cao R. Zinc porphyrin/imidazolium integrated multivariate zirconium metal-organic frameworks for transformation of CO2 into cyclic carbonates[J]. Inorg. Chem., 2018, 57:  2584-2593. doi: 10.1021/acs.inorgchem.7b02983

    42. [42]

      Zhou Z, He C, Xiu J H, Yang L, Duan C Y. Metal-organic polymers containing discrete single-walled nanotube as a heterogeneous catalyst for the cycloaddition of carbon dioxide to epoxides[J]. J. Am. Chem. Soc., 2015, 137:  15066-15069. doi: 10.1021/jacs.5b07925

  • Scheme1  Structures of ligand L (base) and different auxiliary ligands (acid)

    Figure 1  Structure of MOF 1: (a) coordination environment of Zn2+ ion; (b, c) two different 1D chains; (d) 2D layer

    Symmetry codes: 0.5-x, 0.5+y, z; 1.5-x, -0.5+y, z; 1-x, 1-y, z.

    Figure 2  Topological simplification diagram of MOF 1

    Figure 3  Structure of MOF 2: (a) coordination environment of Zn2+ ion; (b) 1D chain;(c) 2D layer

    Symmetry codes: 1-x, y, z; 1+x, 0.5+y, 0.5+z.

    Figure 4  Topological simplification diagram of MOF 2

    Figure 5  Plausible mechanism for the cycloaddition of epoxides with CO2 catalyzed by MOF 2

    Table 1.  Orthogonal experiments of CO2 cycloaddition reactionsa

    Entry xcatalyst/% xTBAB/% Time/h Conversionb/%
    1 0.1 0 24 0
    2 0 8.0 24 32.48
    3 0.1 1.0 24 30.30
    4 0.1 3.0 24 37.88
    5 0.1 5.0 24 52.36
    6 0.1 6.0 24 66.67
    7 0.1 7.0 24 89.30
    8 0.1 8.0 24 > 99
    9 0.08 8.0 24 69.44
    10 0.05 8.0 24 47.85
    11 0.1 8.0 18 67.57
    12 0.1 8.0 12 38.91
    13 0.1 8.0 6 18.42
    anepoxide=20 mmol; bThe conversion of the epoxide was determined by 1H NMR analysis (Fig.S8-S21).
    下载: 导出CSV

    Table 2.  Catalytic activity of various catalysts for propylene oxidea

    Entry Catalyst Epoxide Product Conversionb/% TONc
    1 1 89.28 892.8
    2 2 100.0 1 000.0
    3 ZnCl2 51.55 515.5
    4 Zn(NO3)2·6H2O 52.36 523.6
    a Reaction conditions: propylene oxide (20 mmol), xcatalyst=0.1%, xTBAB=8%, RT, atmospheric pressure (balloon), 24 h; b The conversion of the epoxide was determined by 1H NMR analysis; c TON=nproduct/ncatalyst (Fig.S22-S24).
    下载: 导出CSV

    Table 3.  Cycloaddition reactions of CO2 and various epoxides with catalyst 2a

    Entry Epoxide Product Conversionb/% TONc
    1 100.0 10 00.0
    2 57.80 578.0
    3 55.87 558.7
    4 14.53 145.3
    aReaction conditions: epoxide (20 mmol), xcatalyst=0.1%, xTBAB=8%, RT, atmospheric pressure (balloon), 24 h; b The con-version of the epoxide was determined by 1H NMR analysis; c TON=nproduct/ncatalyst (Fig.S25-Fig.S27).
    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  219
  • HTML全文浏览量:  18
文章相关
  • 发布日期:  2024-07-10
  • 收稿日期:  2023-10-31
  • 修回日期:  2024-04-17
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章