Urothermal Syntheses of Chiral Zincic Benzotriazole-5-carboxylate Frameworks for Iodine Uptake

Zai-Chun HUANG Xiao-Fang WANG Bi Foua Claude Alain GOHI Yu-Feng LI Yao KANG Zhi-Qiang JIANG

Citation:  Zai-Chun HUANG, Xiao-Fang WANG, Bi Foua Claude Alain GOHI, Yu-Feng LI, Yao KANG, Zhi-Qiang JIANG. Urothermal Syntheses of Chiral Zincic Benzotriazole-5-carboxylate Frameworks for Iodine Uptake[J]. Chinese Journal of Structural Chemistry, 2021, 40(2): 199-206. doi: 10.14102/j.cnki.0254–5861.2011–2796 shu

Urothermal Syntheses of Chiral Zincic Benzotriazole-5-carboxylate Frameworks for Iodine Uptake

English

  • During the past several decades, crystalline metal-organic frameworks (MOFs) had received intensive attention due to their specific properties, as well as intriguing structures, in which inorganic component and organic linker are assembled together by strong coordination bond and regularly arranged in designated location in the lattice[1-5]. But up to now, a lot of efforts are still devoted to synthesize this new crystalline material due to several potential applications in many highly active fields, such as gas storage/separation, small/large molecule absorption, fluorescence sensor, drug delivery and catalysis[6-19]. Although both metal units and organic components can play a crucial role in the unique shape and prospective function, the structural evolution and stability of framework can be also affected by solvents or synthesis method. The different combination of solvents and their dosage proportion are usually used as effective strategy to demonstrate the structural evolution[20-25], which can be further explained by one of Li's recent works[26]. In this reported work, a series of intriguing multicomponent trinuclear copper pyrazolate metal-organic frameworks were assembled by combining N, N-dimethylformamide (DMF), N, N-dimethylacetamide (DMA) and N-methyl-2-pyrrolidone (NMP). In addition to focusing on the solvent, choosing suitable synthetic method is also useful to various architectures based on similar metal units and organic linkers[27, 28]. In 2010, Zhang and Bu developed a new synthetic method, named urothermal synthesis[29]. During this synthetic process, urea derivatives have competed with homologues or other solvents like DMF or DEF to coordinate to metal sites, which can be utilized to control the structure of framework.

    As a subclass of MOFs, crystalline chiral MOFs, especially homochiral MOF materials, have received explosive attention, owing to their potential applications in enantioselective processes, heterogeneous asymmetric catalysis and sensor technology in optoelectronics[30-35]. However, such chiral MOFs can generally be constructed from chiral organic ligands, and only a limited number of chiral MOFs were synthesized by achiral organic linker[36-39]. After being inspired by the strategy that N-donor ligand and O-donor ligands could be simultaneously employed to extend structure and functionality, many aromatic organic ligands containing carboxylic and nitrogen donor groups, such as imidazoledi-carboxylic and triazoledicarboxylic acids, have been researched extensively in constructing a series of metal-organic materials[40-43]. However, it is rare to use above-mentioned achiral carboxylic and nitrogen donor ligand to fabricate chiral frameworks, which suggests that the preparation of chiral framework with this kind of organic ligand is still faced with a great challenge because many uncertain problems originating from versatile linker factors and variable metallic coordination configuration have yet to be overcome. Despite the fact that it still plays an essential role in the field of biosciences, functional materials and so on, iodine also has certain drawbacks, such as radioactive 129I, which is a source of severe environmental pollution. Iodine uptake is therefore important to counter its harmful effects on the environment. As far as we know, it is very rare that chiral MOFs were employed to absorb iodine.

    In this work, we used benzotriazole-5-carboxylate acid as a multidentate organic ligand to assemble chiral MOFs. In fact, merely a handful of framework with this ligand can be achieved, let alone chiral backbone, which may be modified through a new synthetic method. Herein, three complexes, namely [Zn(btca)Cl]·DMA·N(CH3)2 (1), [Zn(btca)(Ac-)]· N(CH3)2 (2) and [Zn(btca)]·H2O (3), were obtained. It was interesting to note that their structures demonstrated the characteristic of structural diversities based on similar coordination mode in which 1 and 2 showed two-dimensional (2D) achiral network with fes topology, while 3 presented a three-dimensional chiral porous framework with eta topology. This meaningful result could be assigned to the urothermal synthesis with ethyleneurea (e-urea) and 1, 3-dimethyl-3, 4, 5, 6-tetrahydro-2(1H)-pyrimidinone (DMPU) as template agents. Meanwhile, compound 3 exhibited an excellent property of iodine uptake.

    All reagents and solvents were of analytic grade and were purchased from Chinese commercial sources and used without further purification. Diffraction data were collected by using Saturn724+CCD and XCalibur E CCD diffractometers, respectively. The powder diffraction data were collected on a MiniFlex2 X-ray diffractometer.

    2.2.1   Syntheses of compounds 1 and 2

    ZnCl2 (0.1207 g, 0.89 mmol) for 1, Zn(NO3)2·6H2O (0.1810 g, 0.63 mmol) for 2, benzotriazole-5-carboxylate acid (btca, 0.0862 g, 0.49 mmol), pyrazine (0.0301 g, 0.38 mmol) and N, N-dimethylacetamide (DMA, 2 mL) were spilled into a 20 mL vial and stirred for 10 minutes. The mixture was kept into the oven at 130 ℃ for five days. Colorless transparent block-like crystals 1 were obtained. Compound 2 was synthesized similar to 1 with adding additional solvents of methanol (2 mL), acetonitrile (2 mL) and water (2 mL).

    2.2.2   Synthesis of compound 3

    Compound 3 was synthesized by urothermal synthesis. Urothermal reaction of Zn(NO3)2·6H2O (0.1807 g, 0.63 mmol), benzotriazole-5-carboxylate acid (btca, 0.0859 g, 0.48 mmol), pyrazine (0.0300 g, 0.38 mmol), e-urea (1.2 g) and DMPU (3 mL) in a 20 mL vial at 130 ℃ for 3 days produced colorless crystals and then cooled to room temperature. After washing with alcohol, the resulting compound 3 was obtained.

    Single-crystal X-ray diffraction data collection was performed on the computer-controlled Saturn 724+CCD diffractometer and XCalibur E CCD single-crystal diffractometer with graphite-monochromatic Mo (λ = 0.71073 Å) at T = 293(2) K, respectively. The structure was solved by direct methods and refined by full-matrix least-squares methods on F2 by using the SHELX-97 program package. All non-hydrogen atoms were refined anisotropically. Hydrogen atoms were generated geometrically. The crystallographic data, and related bond lengths and bond angles for compounds 1~3 are listed in Tables 1 and 2.

    Table 1

    Table 1.  Crystal Data and Structure Refinement for Complexes 1~3
    DownLoad: CSV
    Complex 1 2 3
    Formula C13H18ClN5O3Zn C11H12N4O4Zn C7H3N3O3Zn
    Fw 393.14 329.62 242.49
    Crystal system Orthorhombic Orthorhombic Trigonal
    Space group Pbca Pbca P3221
    a (Å) 16.2563(9) 10.189(8) 10.0172(3)
    b (Å) 9.870(2) 16.464(14) 10.0172(3)
    c (Å) 22.871(2) 17.052(15) 26.2843(8)
    α (º) 90 90 90
    β (º) 90 90 90
    γ (º) 90 90 120
    V3) 3669.7(9) 2861(4) 2284.12(12)
    Z 8 8 6
    Dc (g·cm-3) 1.423 1.531 1.058
    µ (mm–1) 1.503 1.734 1.602
    Observed 3447/3447 3665/3665 6665/2549
    Rint 0.0321 0.1480 0.0490
    GOF on F2 1.046 1.078 1.013
    Ra (I > 2σ(I)) 0.0753 0.0782 0.0798
    wRb (I > 2σ(I)) 0.2372 0.2305 0.2161
    aR = Σ||Fo| |Fc||/Σ|Fo|, bwR = Σ[w(Fo2 Fc2)2]/Σ[w(Fo2)2]1/2

    Table 2

    Table 2.  Selected Bond Lengths (Å) and Angles (˚) for Complexes 1~3
    DownLoad: CSV
    Complex 1
    Bond Dist. Bond Dist.
    Zn(1)–O(1) 1.942(5) Zn(1)–N(3)#2 2.038(9)
    Zn(1)–N(1)#1 2.029(10) Zn(1)–Cl(1) 2.226(3)
    Angle (°) Angle (°)
    O(1)–Zn(1)–N(1)#1 99.6(4) O(1)–Zn(1)–Cl(1) 118.7(2)
    O(1)–Zn(1)–N(3)#2 109.2(4) N(1)#1–Zn(1)–Cl(1) 111.5(3)
    N(1)#1–Zn(1)–N(3)#2 107.5(3) N(3)#2–Zn(1)–Cl(1) 109.4(2)
    Complex 2
    Bond Dist. Bond Dist.
    Zn(1)–O(1) 1.972(6) Zn(1)–N(3) 2.026(6)
    Zn(1)–O(3)#3 1.987(6) Zn(1)–N(1)#4 2.060(6)
    Angle (°) Angle (°)
    O(1)–Zn(1)–O(3)#3 108.4(3) O(1)–Zn(1)–N(1)#4 117.5(3)
    O(1)–Zn(1)–N(3) 107.8(2) O(3)#3–Zn(1)–N(1)#4 98.7(2)
    O(3)#3–Zn(1)–N(3) 111.8(2)
    Complex 3
    Bond Dist. Bond Dist.
    Zn(1)–O(1)#1 1.936(8) Zn(1)–N(3)#2 1.964(10)
    Zn(1)–N(1) 1.933(11) Zn(1)–O(1W) 1.987(12)
    Angle (°) Angle (°)
    O(1)#1–Zn(1)–N(1) 118.2(4) O(1)#1–Zn(1)–O(1W) 100.2(5)
    O(1)#1–Zn(1)–N(3)#2 114.4(4) N(1)–Zn(1)–O(1W) 98.3(5)
    N(1)–Zn(1)–N(3)#2 119.5(4) N(3)#2–Zn(1)–O(1W) 99.7(5)
    a Symmetry transformation used to generate the equivalent atoms for 1: #1 –x, –y+1, –z+1; #2 x+1/2, –y+3/2, –z+1; #3 x–1/2, –y+3/2, –z+1.
    For 2: #1 –x+1/2, –y, z–1/2; #2 x–1/2, y, –z+1/2; #3 –x+1/2, –y, z+1/2; #4 x+1/2, y, –z+1/2.
    for 3: #1 xy, –y+1, –z+1/3; #2 y, x–1, –z; #3 xy+1, –y+1, –z+1/3; #4 y+1, x, –z

    Single-crystal X-ray diffractions of complexes 1 and 2 clarify that both of them crystallize in the same orthorhombic space group Pbca, which are isostructural frameworks. In their asymmetric units, there are only one independent metal left and one btca ligand. Zn(1) shows a distorted ZnN2OCl (in compound 1) or ZnN2O2 (in compound 2) tetrahedral geometry with two nitrogen atoms, originating from triazole rings, one oxygen atom from a monodentate carboxylate of btca ligand, a Cl for 1 and deprotonated acetic acid for 2 coordinated with a terminal site (Fig. 1a and 1c). The coordination mode of btca ligand in 1 and 2 is μ3-η1: η0: η1: η0: η1. For the interactions between metal sites and ligands described above, two basic units are formed: every two Zn lefts and two linkers form a [Zn2(btca)2] unit possessing a 14-membered ring; every four Zn lefts and four linkers form a [Zn4(btca)4] unit with a 24-membered ring. The neighboring rings are arranged in an ···ABAB··· order along a wavy curve along the c axis because both 1 and 2 exhibit overall wave-like layers (Fig. 1b and 1d). It is notable that the coordinated acetic acid molecules in complex 2 are arranged in "up" and "down" order along the c axis distributing on both sides of the layer. The layers in structures of 1 and 2 can be reduced as a 3-connected uninodal net (denoted: fes) with the point symbol of {4·82} by reducing each btca organic linker and metal left as a 3-connected node, respectively (Fig. 1e).

    Figure 1

    Figure 1.  (a) Asymmetric unit of 1; (b) Perspective view of the two-dimensional sheet of 1; (c) Coordination environment of 2; (d) Two-dimensional layer in 1; (e) Two-dimensional fes topologic layer in 1 and 2. Hydrogen atoms are omitted for clarity

    In order to extend the structure of btca framework, two urea derivatives e-urea and DMPU were employed to induce structural evolution from achiral layer to chiral threedimensional network. Single-crystal X-ray diffraction of complex 3 reveals that it crystallizes in the trigonal space group P3221. In the structure of 3, the asymmetric unit still consisted of a single independent zinc left, one btca ligand and one solvent water molecule (Fig. 2a). As the same as in 1 and 2, Zn atom in 3 also shows a distorted ZnN2O2 tetrahedral geometry with two nitrogen atoms originating from triazole rings, two oxygen atoms from one monodentate carboxylate of btca ligand and a terminal water molecule. Interestingly, due to the involvement of urea derivatives in the reaction, the coordination angle and geometric configuration of inorganic metal ion and organic linker are quite different from those of 1 and 2, which resulted in the helical sub-structure 3. As shown in Fig. 2b, both P- and M-helixes can be observed in the porous framework. To clarify, the P- and M-helices are unseparated and uneven. These helical sub-structures in 3 intertwine to result in a three-dimensional chiral framework (Fig. 2c). From the viewpoint of topology, the resulting network can be also reduced as a 3-connecteduninodal net (denoted: eta) with the point symbol of {83} by considering both btca ligand and Zn left as a 3-connected node (Fig. 2d).

    Figure 2

    Figure 2.  (a) Coordination environment around the Zn atoms in 3; (b) Helical chain (P- and M-helixes) in 3; (c) Perspective view of the three-dimensional framework of 3; (d) Three-dimensional eta topologic network in 3. Hydrogen atoms are omitted for clarity

    The purity of 3 can be checked by the powder X-ray diffraction patterns (PXRD). As illustrated in Fig. 3, the curve shape of 3 can exactly match with that of the simulated one, which confirms that complex 3 is very pure. The afterabsorbed PXRD patterns are also in good agreement with the as-synthesized samples and the measured single crystals, proving the structure of 3 remains stable in the process of gathering iodine.

    Figure 3

    Figure 3.  Characterization of bulk phase of 3 by PXRD

    The thermogravimetric analysis (TGA) of compound 3 was carried out under N2 atmosphere with a heating rate of 10 ℃·min–1 in the range from 30 to 600 ℃. As depicted in Fig. 4, the TGA curves reveal that complex 3 exhibited a good thermal stability. The solvents on the sample surface were gradually removed in proceeding from 30 to 120 ℃. A slight weight loss of 4.33% was observed between 150 and 300 ℃, which could be assigned to the removal of coordinated water molecule (calcd. 6.59%). Over 330 ℃, the sharp weight loss up to 550 ℃ is due to the decomposition of Pbca organic ligands, accompanying the collapse of the whole framework.

    Figure 4

    Figure 4.  TGA curves of compound 3

    135 mg of prepared samples of 3 was immersed in a cyclohexane solution of I2 (0.1 mol/L) under room temperature and kept out of light in order to test the ability of I2 uptake. The color of solution gradually changed from dark red to pale brown, while the color of sample 3 varied from colorlessness to dark brown (Fig. 5a), which can phenolmenally illustrate the adsorption of I2. Based on its value-added, the weight of 3 after loading iodine was increased by about 15.6wt%. As shown in Fig. 5b, the compounds after loading iodine may release the absorbents I2, resulting in a color variation from colorlessness to pale brown, which indicates that I2 sorption is reversible. A little I2-loaded sample of 3 was put into 9 mL ethanol. This process can be also proven by the UV-vis spectroscopy at room temperature (Fig. 5c) to further confirm and quantify the above observation. The absorbance of I2 extracted into ethanol at 238 nm normally increased over time and the dynamic equilibrium of the release and adsorption of I2 was reached within 2 hours.

    Figure 5

    Figure 5.  (a) Photos of the iodine recovery process. (b) Iodine release process of 3 soaked in 9 mL of EtOH. (c) Temporal evolution of UV/vis absorption spectra for the release of I2 from 3

    In summary, unusual chiral MOFs have been successfully synthesized by employing achiral multidentate benzotriazole-5-carboxylate acid ligand with N- and O-donors. This success mainly derives from changing synthetic method that urea derivatives act as structure-directing agent to revolutionize the coordination angle and geometric configuration of inorganic metal ion and organic linker. Interestingly, compound 3 shows excellent performance of enriching iodine. This work could provide a new chance to build chiral framework with achiral organic ligand.


    1. [1]

      Zhang, J. P.; Zhang, Y. B.; Lin, J. B.; Chen, X. M. Metal azolate frameworks: from crystal engineering to functional materials. Chem. Rev. 2012, 112, 1001–1033. doi: 10.1021/cr200139g

    2. [2]

      Pang, J.; Wu, M.; Qin, J. S.; Liu, C.; Lollar, C. T.; Yuan, D.; Hong, M.; Zhou, H. C. Solvent-assisted, thermally triggered structural transformation in flexible mesoporous metal-organic frameworks. Chem. Mater. 2019, 31, 8787–8793. doi: 10.1021/acs.chemmater.9b02582

    3. [3]

      Liu, Q.; Song, Y.; Ma, Y.; Zhou, Y.; Cong, H.; Wang, C.; Wu, J.; Hu, G.; O'Keeffe, M.; Deng, H. Mesoporous cages in chemically robust MOFs created by a large number of vertices with reduced connectivity. J. Am. Chem. Soc. 2019, 141, 488–496. doi: 10.1021/jacs.8b11230

    4. [4]

      Su, J.; He, W.; Li, X. M.; Sun, L.; Wang, H. Y.; Lan, Y. Q.; Ding, M.; Zuo, J. L. High electrical conductivity in a 2D MOF with intrinsic superprotonic conduction and interfacial pseudo-capacitance. Matter 2020, 2, 1–12. doi: 10.1016/j.matt.2019.12.003

    5. [5]

      Luo, Y. H.; Dong, L. Z.; Liu, J.; Li, S. L.; Lan, Y. Q. From molecular metal complex to metal-organic framework: the CO2 reduction photocatalysts with clear and tunable structure. Coordin. Chem. Rev. 2019, 390, 86–126. doi: 10.1016/j.ccr.2019.03.019

    6. [6]

      Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent functional metal-organic frameworks. Chem. Rev. 2012, 112, 1126–1162. doi: 10.1021/cr200101d

    7. [7]

      Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van Duyne, R. P.; Hupp, J. T. Metal-organic framework materials as chemical sensors. Chem. Rev. 2012, 112, 1105–1125. doi: 10.1021/cr200324t

    8. [8]

      Li, J. R.; Sculley, J.; Zhou, H. C. Metal-organic frameworks for separations. Chem. Rev. 2012, 112, 869–932. doi: 10.1021/cr200190s

    9. [9]

      Horcajada, P.; Gref, R.; Baati, T.; Allan, P. K.; Maurin, G.; Couvreur, P.; Ferey, G.; Morris, R. E.; Serre, C. Metal-organic frameworks in biomedicine. Chem. Rev. 2012, 112, 1232–1268. doi: 10.1021/cr200256v

    10. [10]

      Slater, A. G.; Cooper, A. I. Function-led design of new porous materials. Science 2015, 348, 8075–8085. doi: 10.1126/science.aaa8075

    11. [11]

      Remya, V. R.; Kurian, M. Synthesis and catalytic applications of metal-organic frameworks: a review on recent literature. Int. Nano. Lett. 2019, 9, 17–29. doi: 10.1007/s40089-018-0255-1

    12. [12]

      Zhang, Y.; Pang, J.; Li, J.; Yang, X.; Feng, M.; Cai, P.; Zhou, H. C. Visible-light harvesting pyrene-based MOFs as efficient ROS generators. Chem. Sci. 2019, 10, 8455–8460. doi: 10.1039/C9SC03080H

    13. [13]

      Dong, Z.; Sun, Y.; Chu, J.; Zhang, X.; Deng, H. Multivariate metal-organic frameworks for dialing-in the binding and programming the release of drug molecules. J. Am. Chem. Soc. 2017, 139, 14209–14216. doi: 10.1021/jacs.7b07392

    14. [14]

      Zhao, Y.; Deng, D. S.; Ma, L. F.; Ji, B. M.; Wang, L. Y. A new copper-based metal-organic framework as a promising heterogeneous catalyst for chemo- and regio-selective enamination of β-ketoesters. Chem. Commun. 2013, 49, 10299–10301. doi: 10.1039/c3cc45310c

    15. [15]

      Wu, Q. Q.; Wen, Y. H. Hydrothermal syntheses, crystal structures and luminescence properties of Zn(Ⅱ) and Cu(Ⅱ) complexes based on 2, 2΄-((sulfonylbis(4, 1-phenylene))bis(oxy))diacetic acid. Chin. J. Struct. Chem. 2019, 39, 294–300

    16. [16]

      Jiang, Z. Q.; Li, Y. F.; Zhu, X. J.; Lu, J.; Zhang, L.; Wen, T. Ni(Ⅱ)-based coordination polymers for efficient electrocatalytic oxygen evolution reaction. RSC. Adv. 2018, 8, 38562–38565. doi: 10.1039/C8RA07492E

    17. [17]

      Jiang, Z. Q.; Li, Y. F.; Zhu, X. J.; Lu, J.; Wen, T.; Zhang, L. Ni(Ⅱ)-doped anionic metal-organic framework nanowire arrays for enhancing the oxygen evolution reaction. Chem. Commun. 2019, 55, 4023–4026. doi: 10.1039/C9CC00009G

    18. [18]

      Jiang, Z. Q.; Chen, X. L.; Lu, J.; Li, Y. F.; Wen, T.; Zhang, L. Ultrathin Ni(Ⅱ)-based coordination polymer nanosheets as a co-catalyst for promoting photocatalytic H2-production. Chem. Commun. 2019, 55, 6499–6502. doi: 10.1039/C9CC02680K

    19. [19]

      Jiang, Z. Q.; Wang, F.; Zhang, J. Rational design of zeolitic tetrazolate frameworks with carboxylate ligands for rapid accumulation of iodine. Chin. J. Appl. Chem. 2017, 34, 1072–1078.

    20. [20]

      Seetharaj, R.; Vandana, P. V.; Arya, P.; Mathew, S. Dependence of solvents, pH, molar ratio and temperature in tuning metal organic framework architecture. Arab. J. Chem. 2019, 12, 295–315. doi: 10.1016/j.arabjc.2016.01.003

    21. [21]

      Kundu, T.; Wahiduzzaman, M.; Shah, B. B.; Maurin, G.; Zhao, D. Solvent-induced control over breathing behavior in flexible metal-organic frameworks for natural-gas delivery. Angew. Chem. Int. Ed. 2019, 58, 8073–8077. doi: 10.1002/anie.201902738

    22. [22]

      Li, L. N.; Wang, S. Y.; Chen, T. L.; Sun, Z. H.; Luo, J.; Hong, M. C. Solvent-dependent formation of Cd(Ⅱ) coordination polymers based on a C2-symmetric tricarboxylate linker. Cryst. Growth. Des. 2012, 12, 4109–4115 doi: 10.1021/cg300617h

    23. [23]

      Pedireddi, V. R.; Varughese, S. Solvent-dependent coordination polymers: cobalt complexes of 3, 5-dinitrobenzoic acid and 3, 5-dinitro-4-methylbenzoic acid with 4, 4΄-bipyrdine. Inorg. Chem. 2004, 43, 450–457. doi: 10.1021/ic0349499

    24. [24]

      Köppen, M.; Meyer, V.; Ångström, J.; Inge, A. K.; Stock, N. Solvent-dependent formation of three new bi-metal-organic frameworks using a tetracarboxylic acid. Cryst. Growth Des. 2018, 18, 4060–4067. doi: 10.1021/acs.cgd.8b00439

    25. [25]

      Jiang, Z. Q.; An, Y.; Zhu, X.; Tian, C.; Bai, J.; Li, Y. F. Solvent-dependent synthesis from layer to microporous pillared-layer framework for selective sorption of gas light hydrocarbons. Z. Anorg. Allg. Chem. 2015, 641, 2599–2603. doi: 10.1002/zaac.201500268

    26. [26]

      Tu, B.; Pang, Q.; Xu, H.; Li, X.; Wang, Y.; Ma, Z.; Weng, L.; Li, Q. Reversible redox activity in multicomponent metal organic frameworks constructed from trinuclear copper pyrazolate building blocks. J. Am. Chem. Soc. 2017, 139, 7998–8007. doi: 10.1021/jacs.7b03578

    27. [27]

      Stock, N.; Biswas, S. Synthesis of metal-organic frameworks (MOFs): routes to various MOF topologies, morphologies, and composites. Chem. Rev. 2012, 112, 933–969. doi: 10.1021/cr200304e

    28. [28]

      Zhang, J.; Chen, S.; Bu, X. Multiple functions of ionic liquids in the synthesis of three-dimensional low-connectivity homochiral and achiral frameworks. Angew. Chem. Int. Ed. 2008, 47, 5434–5437. doi: 10.1002/anie.200801838

    29. [29]

      Zhang, J.; Bu, J. T.; Chen, S.; Wu, T.; Zheng, S.; Chen, Y.; Nieto, R. A.; Feng, P.; Bu, X. Urothermal synthesis of crystalline porous materials. Angew. Chem. Int. Ed. 2010, 49, 8876–8879. doi: 10.1002/anie.201003900

    30. [30]

      Yoon, M.; Srirambalaji, R.; Kim, K. Homochiral metal-organic frameworks for asymmetric heterogeneous catalysis. Chem. Rev. 2012, 112, 1196–1231. doi: 10.1021/cr2003147

    31. [31]

      Zhang, J.; Bu, X. Chiralization of diamond nets: stretchable helices and chiral and achiral nets with nearly identical unit cells. Angew. Chem. Int. Ed. 2007, 46, 6115–6118. doi: 10.1002/anie.200701374

    32. [32]

      Zingiryan, A.; Zhang, J.; Bu, X. Cooperative self-assembly of chiral l-malate and achiral succinate in the formation of a three-dimensional homochiral framework. Inorg. Chem. 2008, 47, 8607–8609. doi: 10.1021/ic801404p

    33. [33]

      Zhang, J.; Chen, S.; Bu, X. Nucleotide-catalyzed conversion of racemic zeolite-type zincophosphate into enantio enriched crystals. Angew. Chem. Int. Ed. 2009, 48, 6049–6051. doi: 10.1002/anie.200903001

    34. [34]

      Zhao, X.; Wong, M.; Mao, C.; Trieu, T. X.; Zhang, J.; Feng, P.; Bu, X. Size-selective crystallization of homochiral camphorate metal-organic frameworks for lanthanide separation. J. Am. Chem. Soc. 2014, 136, 12572–12575. doi: 10.1021/ja5067306

    35. [35]

      Jin, J.; Zhao, X.; Feng, P.; Bu, X. A cooperative pillar-template strategy as generalized synthetic method for flexible 3-D homochiral porous frameworks. Angew. Chem. Int. Ed. 2018, 57, 3737–3741. doi: 10.1002/anie.201801116

    36. [36]

      Zhang, J.; Chen, S.; Wu, T.; Feng, P.; Bu, X. Homochiral crystallization of microporous framework materials from achiral precursors by chiral catalysis. J. Am. Chem. Soc. 2008, 130, 12882–12883. doi: 10.1021/ja805272j

    37. [37]

      Zhang, J.; Chen, S.; Nieto, R. A.; Wu, T.; Feng, P.; Bu, X. A tale of three carboxylates: cooperative asymmetric crystallization of a three-dimensional microporous framework from achiral precursors. Angew. Chem. Int. Ed. 2010, 49, 1267–1270. doi: 10.1002/anie.200906248

    38. [38]

      Morris, R. E.; Bu, X. Induction of chiral porous solids containing only achiral building blocks. Nat. Chem. 2010, 2, 353–361. doi: 10.1038/nchem.628

    39. [39]

      Kang, Y.; Chen, S.; Wang, F.; Zhang, J.; Bu, X. Induction in urothermal synthesis of chiral porous materials from achiral precursors. Chem. Commun. 2011, 47, 4950–4952. doi: 10.1039/c1cc10486a

    40. [40]

      Xiao, J.; Liu, B. Y.; Wei, G.; Huang, X. C. Solvent induced diverse dimensional coordination assemblies of cupric benzotriazole-5-carboxylate: syntheses, crystal structures, and magnetic properties. Inorg. Chem. 2011, 50, 11032–11038. doi: 10.1021/ic201571n

    41. [41]

      Lu, W. G.; Jiang, L.; Lu, T. B. Lanthanide contraction and temperature-dependent structures of lanthanide coordination polymers with imidazole-4, 5-dicarboxylate and oxalate. Cryst. Growth Des. 2010, 10, 10, 4310–4318.

    42. [42]

      Zhang, W. X.; Xue, W.; Zheng, Y. Z.; Chen, X. M. Two spin-competing manganese(Ⅱ) coordination polymers exhibiting unusual multi-step magnetization jumps. Chem. Commun. 2009, 45, 3804–3806.

    43. [43]

      Guo, Z.; Li, X.; Gao, S.; Li, Y.; Cao, R. A new three-dimensional supramolecular network, [Cd(Hbic)2(H2O)]·(4, 4΄-bpy)·H2O (H2bic = 1-H-benzimidazole-5-carboxylic acid; 4, 4΄-bpy = 4, 4΄-bipyridine): synthesis, crystal structure and luminescence property. J. Mol. Struct. 2007, 846, 123–127. doi: 10.1016/j.molstruc.2007.01.036

  • Figure 1  (a) Asymmetric unit of 1; (b) Perspective view of the two-dimensional sheet of 1; (c) Coordination environment of 2; (d) Two-dimensional layer in 1; (e) Two-dimensional fes topologic layer in 1 and 2. Hydrogen atoms are omitted for clarity

    Figure 2  (a) Coordination environment around the Zn atoms in 3; (b) Helical chain (P- and M-helixes) in 3; (c) Perspective view of the three-dimensional framework of 3; (d) Three-dimensional eta topologic network in 3. Hydrogen atoms are omitted for clarity

    Figure 3  Characterization of bulk phase of 3 by PXRD

    Figure 4  TGA curves of compound 3

    Figure 5  (a) Photos of the iodine recovery process. (b) Iodine release process of 3 soaked in 9 mL of EtOH. (c) Temporal evolution of UV/vis absorption spectra for the release of I2 from 3

    Table 1.  Crystal Data and Structure Refinement for Complexes 1~3

    Complex 1 2 3
    Formula C13H18ClN5O3Zn C11H12N4O4Zn C7H3N3O3Zn
    Fw 393.14 329.62 242.49
    Crystal system Orthorhombic Orthorhombic Trigonal
    Space group Pbca Pbca P3221
    a (Å) 16.2563(9) 10.189(8) 10.0172(3)
    b (Å) 9.870(2) 16.464(14) 10.0172(3)
    c (Å) 22.871(2) 17.052(15) 26.2843(8)
    α (º) 90 90 90
    β (º) 90 90 90
    γ (º) 90 90 120
    V3) 3669.7(9) 2861(4) 2284.12(12)
    Z 8 8 6
    Dc (g·cm-3) 1.423 1.531 1.058
    µ (mm–1) 1.503 1.734 1.602
    Observed 3447/3447 3665/3665 6665/2549
    Rint 0.0321 0.1480 0.0490
    GOF on F2 1.046 1.078 1.013
    Ra (I > 2σ(I)) 0.0753 0.0782 0.0798
    wRb (I > 2σ(I)) 0.2372 0.2305 0.2161
    aR = Σ||Fo| |Fc||/Σ|Fo|, bwR = Σ[w(Fo2 Fc2)2]/Σ[w(Fo2)2]1/2
    下载: 导出CSV

    Table 2.  Selected Bond Lengths (Å) and Angles (˚) for Complexes 1~3

    Complex 1
    Bond Dist. Bond Dist.
    Zn(1)–O(1) 1.942(5) Zn(1)–N(3)#2 2.038(9)
    Zn(1)–N(1)#1 2.029(10) Zn(1)–Cl(1) 2.226(3)
    Angle (°) Angle (°)
    O(1)–Zn(1)–N(1)#1 99.6(4) O(1)–Zn(1)–Cl(1) 118.7(2)
    O(1)–Zn(1)–N(3)#2 109.2(4) N(1)#1–Zn(1)–Cl(1) 111.5(3)
    N(1)#1–Zn(1)–N(3)#2 107.5(3) N(3)#2–Zn(1)–Cl(1) 109.4(2)
    Complex 2
    Bond Dist. Bond Dist.
    Zn(1)–O(1) 1.972(6) Zn(1)–N(3) 2.026(6)
    Zn(1)–O(3)#3 1.987(6) Zn(1)–N(1)#4 2.060(6)
    Angle (°) Angle (°)
    O(1)–Zn(1)–O(3)#3 108.4(3) O(1)–Zn(1)–N(1)#4 117.5(3)
    O(1)–Zn(1)–N(3) 107.8(2) O(3)#3–Zn(1)–N(1)#4 98.7(2)
    O(3)#3–Zn(1)–N(3) 111.8(2)
    Complex 3
    Bond Dist. Bond Dist.
    Zn(1)–O(1)#1 1.936(8) Zn(1)–N(3)#2 1.964(10)
    Zn(1)–N(1) 1.933(11) Zn(1)–O(1W) 1.987(12)
    Angle (°) Angle (°)
    O(1)#1–Zn(1)–N(1) 118.2(4) O(1)#1–Zn(1)–O(1W) 100.2(5)
    O(1)#1–Zn(1)–N(3)#2 114.4(4) N(1)–Zn(1)–O(1W) 98.3(5)
    N(1)–Zn(1)–N(3)#2 119.5(4) N(3)#2–Zn(1)–O(1W) 99.7(5)
    a Symmetry transformation used to generate the equivalent atoms for 1: #1 –x, –y+1, –z+1; #2 x+1/2, –y+3/2, –z+1; #3 x–1/2, –y+3/2, –z+1.
    For 2: #1 –x+1/2, –y, z–1/2; #2 x–1/2, y, –z+1/2; #3 –x+1/2, –y, z+1/2; #4 x+1/2, y, –z+1/2.
    for 3: #1 xy, –y+1, –z+1/3; #2 y, x–1, –z; #3 xy+1, –y+1, –z+1/3; #4 y+1, x, –z
    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  473
  • HTML全文浏览量:  6
文章相关
  • 发布日期:  2021-02-01
  • 收稿日期:  2020-03-05
  • 接受日期:  2020-05-12
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