Next Article in Journal
Assessment of the Porous Structure and Surface Chemistry of Activated Biocarbons Used for Methylene Blue Adsorption
Previous Article in Journal
Chemical Composition, Antioxidant Activity, and Antibacterial Activity of Black Poplar Buds’ Hydroalcoholic Macerates from Dobrogea Area
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Metalation of a Hierarchical Self-Assembly Consisting of π-Stacked Cubes through Single-Crystal-to-Single-Crystal Transformation

1
CAS Key Laboratory of Design and Assembly of Functional Nanostructures, and Fujian Provincial Key Laboratory of Nanomaterials, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China
2
Fujian Science & Technology Innovation Laboratory for Optoelectronic Information of China, Fuzhou 350108, China
3
Xiamen Key Laboratory of Rare Earth Photoelectric Functional Materials, Xiamen Institute of Rare Earth Materials, Haixi Institutes, Chinese Academy of Sciences, Xiamen 361021, China
4
University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(13), 4923; https://doi.org/10.3390/molecules28134923
Submission received: 2 June 2023 / Revised: 17 June 2023 / Accepted: 20 June 2023 / Published: 22 June 2023

Abstract

:
Single-crystal-to-single-crystal metalation of organic ligands represents a novel method to prepare metal–organic complexes, but remains challenging. Herein, a hierarchical self-assembly {(H12L8)·([N(C2H5)4]+)3·(ClO4)15·(H2O)32} (1) (L = tris(2-benzimidazolylmethyl) amine) consisting of π-stacked cubes which are assembled from eight partially protonated L ligands is obtained. By soaking the crystals of compound 1 in the aqueous solution of Co(SCN)2, the ligands coordinate with Co2+ ions stoichiometrically and ClO4 exchange with SCN via single-crystal-to-single-crystal transformation, leading to {([CoSCNL]+)8·([NC8H20]+)3·(SCN)11·(H2O)13} (2).

1. Introduction

Metal–organic complexes have been widely used in a lot of fields including luminescence [1,2,3,4,5,6,7,8], electrical conductivity [9,10,11,12,13], magnetism [14,15,16,17,18,19], catalysis [20,21,22,23,24,25], and so on. In these metal–organic-complex-relevant applications, the metal ions usually act as the functional centers. For example, metal ions can be the luminescent centers in luminescent materials [26], paramagnetic centers in magnetic materials [27], and the active centers in catalysts [28]. Therefore, the metal ions are usually critical to the functionalities of metal–organic complexes.
Metal–organic complexes are traditionally prepared homogeneously by dissolving inorganic metal salts and ligands in solution, and crystals of metal–organic complexes are harvested by crystallization. The crystalline morphology of a metal–organic complex prepared by this strategy is difficult to control. Metalation also occurs heterogeneously via single-crystal-to-single-crystal transformation [29,30,31,32,33]. In general, the metal–organic complexes prepared by this method preserve the parent morphology [30,31]. Therefore, the crystal structures of the produced metal–organic complexes can be easily predicted. Recently, such a post-synthetic metalation method has been used to mount metal active sites onto the open chelating sites of metal–organic frameworks (MOFs) to endow the framework with catalytic activity [32,33]. Specifically, MOFs with open chelating sites were firstly synthesized using linkers with hard carboxylate donors and a soft chelating moiety. Single-crystal-to-single-crystal metalation of the open chelating sites was then achieved under solvothermal conditions [32,33]. We envisage that this metalation method can be used to achieve hierarchical self-assembly bearing open chelating sites, which has not been reached.
Recently, the tripodal ligands tris(2-benzimidazolylmethyl) amine (L) and tris(2-naphthimidazolylmethyl) amine have attracted our attention because of their chelating sites formed by four nitrogen atoms [34,35,36,37]. The benzimidazolylmethyl arms enable L to self-assemble into π-stacked polyhedrons with open chelating sites through π···π interactions. These open chelating sites provide the possibility of single-crystal-to-single-crystal metalation. In this work, we report the single-crystal-to-single-crystal metalation of a hierarchical self-assembly {(H12L8)·([N(C2H5)4]+)3·(ClO4)15·(H2O)32} (1) consisting of π-stacked cubes. Upon immersing the crystals of compound 1 into the aqueous solution of Co(SCN)2, Co2+ ions were mounted onto the chelating sites of L stoichiometrically and ClO4 exchanged with SCN leading to the crystals of {([CoSCNL]+)8·([NC8H20]+)3·(SCN)11·(H2O)13} (2).

2. Results and Discussion

2.1. Structural Characterization of the Hierarchical Self-Assembly of Compound 1

The self-assembly of iron(II) perchlorate hydrate (Fe(ClO4)2·6H2O) with L in MeOH with traces of tetraethylammonium hydroxide ([N(C2H5)4]OH) affords yellow cubic crystals of {(H12L8)·([N(C2H5)4]+)3·(ClO4)15·(H2O)32} (1) (Figure S1). The formula was determined by a combination of single-crystal X-ray crystallography (Table S1) and TG analysis (Figure S2). Compound 1 crystallizes in the cubic space group Fm 3 - c. The asymmetric unit contains 1/3 partially protonated L, 1/8 [N(C2H5)4]+, 5/8 ClO4, and disordered H2O molecules. Surprisingly, the chelating site of L (Figure 1a) was not occupied by FeII during the self-assembly process which may be attributed to the fact that FeII is too soft to coordinate with the chelating site of L.
The crystal structure of compound 1 can be described as a hierarchical self-assembly consisting of π-stacked cubes, and the hierarchical structural complexity is best appreciated by describing it in a bottom-up fashion. As shown in Figure 1, eight partially protonated L ligands associate together through π···π interactions forming a homochiral π-stacked cube, and the cube [H12L8]12+ carries twelve protons to compensate the negative charges from ClO4. Each partially protonated L locates on the vertices of the cube associating with its three neighbors (Figure 1b). The cubes with opposite handedness are connected by [N(C2H5)4]+ and ClO4 forming a three-dimensional (3D) hierarchical self-assembly with [N(C2H5)4]+ and ClO4 positioned at the interstitial regions between the cubes. Due to the simple cubic packing (SC) of the π-stacked cubes in lattice, there exists one-dimensional (1D) channels filled with disordered H2O molecules (Figure 1c).

2.2. Single-Crystal-to-Single-Crystal Metalation of Compound 1

Since we have obtained the supramolecular self-assembly with open chelating sites, we then try to verify the possibility of mounting metal ions onto the chelating sites via single-crystal-to-single-crystal transformation. Soaking in the aqueous solution of Co(SCN)2 (0.15 mol/L), the crystals gradually change from yellow to purple (Figure 2). The photographs showing the color change in the single crystal of compound 1 at different time intervals during the metalation process are shown in Figure S3. After being soaked for ten days, an intact crystal was picked out for structural determination. Single-crystal X-ray crystallography confirmed the successful metalation, and compound 1 transitioned to a similar hierarchical self-assembly {([CoSCNL]+)8·([NC8H20]+)3·(SCN)11·(H2O)13} (2), which consists of π-stacked cubes of {[CoSCNL]+}8. The formula was also determined by a combination of single-crystal X-ray crystallography (Table S1) and TG analysis (Figure S4). A mixture of L (40.8 mg, 0.1 mmol), Co(SCN)2 (35.0 mg, 0.2 mmol), and tetraethylammonium hydroxide ([N(C2H5)4]OH) (0.06 mL) in 5 mL methanol was stirred at room temperature for 5 min. The resulting solution was placed in a beaker undisturbed at room temperature for volatile crystallization; however, compound 2 was not obtained.
The asymmetric unit of compound 2 contains 1/3 [CoSCNL]+, 1/8 [N(C2H5)4]+, 11/24 SCN, and disordered H2O molecules. The π-stacked cube structure is preserved after metalation, and the size of the cube remains almost unchanged. The mounted Co2+ ion coordinates with four N atoms from the same L ligand and one SCN ion, exhibiting a trigonal-bipyramidal geometry (Figure 3a). All the coordinated SCN in the [CoSCNL]+ cations point inward to the center of the cube (Figure 3b). The crystal crystallizes in the same space group Fm 3 - c before and after metalation, but the crystallographic axes significantly shrink from 39.96 to 38.66 Å (Table S1, Figure 4). The unit cell contraction is cross-checked by the significant shift to higher angles of the powder X-ray diffraction (PXRD) peaks of compound 2 compared to those of compound 1. For example, the peaks indexed to (220), (400), and (622) shift from 6.26°, 8.84°, and 15.28° in compound 1 to 6.44°, 9.04°, and 15.58° in compound 2, respectively (Figure 5a). However, the counter anions ClO4 at the interstitial regions between cubes in compound 1 were replaced by SCN in compound 2 (Figure 3c) which can also be revealed by the IR spectrum change (Figure 5b) from compound 1 to compound 2 [38,39]. The absorbances at 1020 and 620 cm−1 characteristic of ClO4 become unobvious, and the absorbance at 2225 cm−1 characteristic of SCN appears. The side lengths of a cubic cage with the central N atom of the ligand L as the vertex are slightly expand after the single-crystal transformation (from 10.415 to 10.618 Å), respectively. We speculate that the anion exchange from tetrahedral ClO4 to linear SCN is the major reason for the observed significant unit cell contraction (from 63,829 to 57,762 Å3) (Figure 4).
To clarify whether the metalation of the ligand Lis complete, a temperature-dependent magnetization study was performed on compound 2. The χmT value of 19.41 cm3 K mol−1 per π-stacked cube is higher than the spin-only value of eight isolated high-spin CoII ions (15 cm3 mol−1 K) (Figure 5c) [40,41], suggesting all the L ligands are coordinated with CoII ions. Upon cooling, the χmT value keeps almost constant until 100 K, and then begins to decrease, reaching a value of 11.81 cm3 mol−1 K at 2 K, implying very weak antiferromagnetic coupling between CoII ions. In the range of 2−300K, the magnetic susceptibility data follows the Curie−Weiss law, providing θ = −4.42 K and C = 19.47 cm3 K mol−1, confirming the dominant weak antiferromagnetic interaction between the CoII ions. To provide further insight into the magnetism of compound 2, the field-dependent magnetizations were measured. As shown in Figure 5d, the magnetization increases slowly with increasing field and reaches a value of 16.24 at 50 K Oe without obvious hysteresis, which is consistent with the weak antiferromagnetic couplings between CoII ions [42,43].

3. Experimental

3.1. Materials and Physical Measurements

The ligand tris(2-benzimidazolylmethyl) amine (L) was synthesized according to the procedure reported in the literature [44], and all the other reagents were commercially obtained and used without further purification. Scanning electron microscopy (SEM; Hitachi SU1510, Chiyoda-ku, Tokyo, Japan) analysis was carried out on a Hitachi SU1510 SEM. Powder X-ray diffraction (PXRD; Miniflex 600, Akishima, Rigaku, Tokyo, Japan) patterns were performed on a Rigaku Miniflex 600 diffractometer with Cu-Kα radiation using flat plate geometry. Thermogravimetric analysis (TGA/DSC 1, Mettler Toledo, Zurich, Switzerland) was performed on a Mettler Toledo TGA/DSC 1 system with a heating rate of 10 K/min under an argon atmosphere. Fourier-transform infrared (FTIR; Nicolet iS 50, Thermo Fisher, Waltham, MA, USA) spectra were recorded in the range of 500–4000 cm−1 on a Thermo Nicolet is50 FT-IR spectrometer at room temperature. Magnetic measurements (MPMS-5S SQUID, Quantum Design, San Diego, CA, USA) were performed on a MPMS-5S SQUID magnetometer under an external field of 1000 Oe.

3.2. Synthesis of Compounds 1 and 2

Synthesis of {(H12L8)·([N(C2H5)4]+)3·(ClO4)15·(H2O)32} (1): A mixture of L (40.8 mg, 0.1 mmol), Fe(ClO4)2·6H2O (76.4 mg, 0.3 mmol), and tetraethylammonium hydroxide ([N(C2H5)4]OH) (0.06 mL) in 5 mL methanol was stirred at room temperature for 5 min. The resulting solution was placed in a beaker undisturbed at room temperature. Yellow cubic crystals of compound 1 were obtained within seven days. Yield: 80% based on L ligand.
Synthesis of {([CoSCNL]+)8·([NC8H20]+)3·(SCN)11·(H2O)13} (2). Upon immersing the crystals of compound 1 in the aqueous of Co(SCN)2 (0.15 mol/L) for 10 days, the crystals of compound 2 were harvested as purple crystals.

3.3. Crystallography

Single-crystal X-ray data were harvested on a Bruker D8 Venture diffractometer with Mo-Kα radiation at 200 K. Structures were solved using a direct method and refined by the full-matrix least-squares technique on F2 with the SHELXTL 2014 program [45]. All the H atoms were geometrically generated and refined using a riding model. The X-ray crystallographic coordinates for structures reported in this article have been deposited at the Cambridge Crystallographic Data Centre (CCDC) under deposition numbers 2266971–2266972. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif (accessed on 2 June 2023). Detail crystallographic data are listed in Table S1.

4. Conclusions

In conclusion, we have synthesized a hierarchical self-assembly consisting of π-stacked cubes with open chelating sites. Learning from the single-crystal-to-single-crystal metalation of MOFs, metal ions were successfully mounted onto the chelating sites via single-crystal-to-single-crystal transformation. This strategy may be used for metalation of other supramolecular frameworks with open coordination sites, thus providing a new method of synthesizing metal–organic complexes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28134923/s1, Table S1: Crystallographic data of the synthetic compounds (1 and 2), Figure S1: Scanning electron microscope (SEM) image of a single crystal of compound 1; Figure S2: TGA for compound 1; Figure S3. The photographs showing the color change in the single crystal of compound 1 at different time intervals during the metalation process. Figure S4: TGA for compound 2.

Author Contributions

B.W. performed experiments and wrote the paper. Z.-A.N. directed the compound characterization and data analysis. J.L. and Z.-X.L. collected data. Z.Z., W.W. and G.-L.L. contributed to interpreting the data. Y.-G.H. conceived the project. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Found (NNSF) of China (92261109 and 21901242), the NSF of Fujian Province (2020J05080), the NSF of Xiamen (3502Z20206080), the Fujian Science and Technology Innovation Laboratory for Optoelectronic Information of China (2021ZR110), the Recruitment Program of Global Youth Experts, and the Youth Innovation Promotion Association.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data related to this study are presented in this publication.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples are available from the author.

References

  1. Wen, H.-R.; Dong, P.-P.; Liu, S.-J.; Liao, J.-S.; Lianga, F.-Y.; Liu, C.-M. 3d–4f heterometallic trinuclear complexes derived from amine-phenol tripodal ligands exhibiting magnetic and luminescent properties. Dalton Trans. 2017, 46, 1153–1162. [Google Scholar] [CrossRef] [PubMed]
  2. Huang, W.; Pan, F.F.; Liu, Y.; Huang, S.D.; Li, Y.J.; Yong, J.; Li, Y.; Kirillov, A.M.; Wu, D.Y. An efficient blue-emissive metal−organic framework (MOF) for lanthanide-encapsulated multicolor and stimuli-responsive luminescence. Inorg. Chem. 2017, 56, 6362–6637. [Google Scholar] [CrossRef] [PubMed]
  3. Connell, T.U.; Earl, S.K.; Ng, C.; Roberts, A.; Davis, T.J.; White, J.M.; Polyzos, A.; Gómez, D.E. Luminescence of a transition metal complex inside a metamaterial nanocavity. Small 2017, 13, 1700692. [Google Scholar] [CrossRef] [Green Version]
  4. Ferrari, F.; Braun, J.; Anson, C.E.; Wilts, B.D.; Moatsou, D.; Bizzarri, C. Cyan-emitting Cu(I) complexes and their luminescent metallopolymers. Molecules 2021, 26, 2567. [Google Scholar] [CrossRef] [PubMed]
  5. Ali, M.A.; Liu, X.F.; Xu, B.B.; Li, Y.; Mohamed, M.A.; Yue, Y.Z.; Qiu, J.R. Vitrification and luminescence properties of metal–organic complexes. ACS Mater. Lett. 2022, 4, 2613–2621. [Google Scholar] [CrossRef]
  6. Herr, P.; Kerzig, C.; Larsen, C.B.; Häussinger, D.; Wenger, O.S. Manganese(I) complexes with metal-to-ligand charge transfer luminescence and photoreactivity. Nat. Chem. 2021, 13, 956–962. [Google Scholar] [CrossRef]
  7. Lee, L.C.C.; Lo, K.K.W. Luminescent and photofunctional transition metal complexes: From molecular design to diagnostic and therapeutic applications. J. Am. Chem. Soc. 2022, 144, 14420–14440. [Google Scholar] [CrossRef]
  8. Saito, D.; Ogawa, T.; Yoshida, M.; Takayama, J.; Hiura, S.; Murayama, A.; Kobayashi, A.; Kato, M. Intense red-blue luminescence based on superfine control of metal–metal interactions for self-assembled platinum (II) complexes. Angew. Chem. Int. Ed. 2020, 59, 18723–18730. [Google Scholar] [CrossRef]
  9. Kitagawa, Y.; Tada, H.; Era, I.; Fujii, T.; Ikenaga, K.; Nakano, M. Theoretical study on the difference in electron conductivity of a one-dimensional penta-nickel (II) complex between anti-ferromagnetic and ferromagnetic states—Possibility of molecular switch with open-shell molecules. Molecules 2019, 24, 1956. [Google Scholar] [CrossRef] [Green Version]
  10. Rubio-Giménez, V.; Tatay, S.; Martí-Gastaldo, C. Electrical conductivity and magnetic bistability in metal–organic frameworks and coordination polymers: Charge transport and spin crossover at the nanoscale. Chem. Soc. Rev. 2020, 49, 5601–5638. [Google Scholar] [CrossRef]
  11. Chen, G.; Gee, L.B.; Xu, W.Q.; Zhu, Y.B.; Lezama-Pacheco, J.S.; Huang, Z.H.; Li, Z.Q.; Babicz, J.T., Jr.; Choudhury, S.; Chang, T.-H.; et al. Valence-dependent electrical conductivity in a 3D tetrahydroxyquinone-based metal−organic framework. J. Am. Chem. Soc. 2020, 142, 21243–21248. [Google Scholar] [CrossRef]
  12. Hutchins, K.M.; Rupasinghe, T.P.; Ditzler, L.R.; Swenson, D.C.; Sander, J.R.G.; Baltrusaitis, J.; Tivanski, A.V.; MacGillivray, L.R. Nanocrystals of a metal−organic complex exhibit remarkably high conductivity that increases in a single-crystal-to-single-crystal transformation. J. Am. Chem. Soc. 2014, 136, 6778–6781. [Google Scholar] [CrossRef]
  13. Liu, J.; Lu, Z.-X.; Wu, F.-F.; Wang, B.; Cao, X.-L.; Wang, W.; Zhuo, Z.; Li, Q.-H.; Huang, Y.-G. A chiral SrSi2 (srs) superstructure constructed by a dual interaction system showing isotropic electrical conductivity. Chin. Chem. Lett. 2022, 1001, 108100. [Google Scholar] [CrossRef]
  14. Miao, Q.; Liu, W.-M.; Kock, T.; Blok, A.; Timmer, M.; Overhand, M.; Ubbink, M. A double-armed, hydrophilic transition metal complex as a paramagnetic NMR probe. Angew. Chem. Int. Ed. 2019, 58, 13093–13100. [Google Scholar] [CrossRef] [Green Version]
  15. Lu, Z.X.; Zhuo, Z.; Wang, W.; Huang, Y.-G.; Hong, M.C. {Gd44Ni22}: A gigantic 3d–4f wheel-like nanoscale cluster with a large magnetocaloric effect. Inorg. Chem. Front. 2023, 10, 979–983. [Google Scholar] [CrossRef]
  16. Terada, N.; Mamiya, H. High-efficiency magnetic refrigeration using holmium. Nat. Commun. 2021, 12, 1212. [Google Scholar] [CrossRef]
  17. Liu, S.-J.; Han, S.-D.; Zhao, J.-P.; Xu, J.-L.; Bu, X.-H. In-situ synthesis of molecular magnetorefrigerant materials. Coordin. Chem. Rev. 2019, 394, 39–52. [Google Scholar] [CrossRef]
  18. Bahjat, E.R.; Liu, J.W.; Béreau, V.; Duhayon, C.; Horino, Y.; Suzuki, T.; Coolen, L.; Sutter, J.P. Concomitant emergence of circularly polarized luminescence and single-molecule magnet behavior in chiral-at-metal Dy complex. Inorg. Chem. Front. 2020, 7, 4527–4534. [Google Scholar]
  19. Matsuoka, R.; Yoshimoto, T.; Kitagawa, Y.; Kusamoto, T. Structural and magnetic studies on Nickel (II) and Cobalt (II) complexes with polychlorinated diphenyl(4-pyridyl) methyl radical ligands. Molecules 2021, 26, 5596. [Google Scholar] [CrossRef]
  20. Sawano, T.; Lin, Z.K.; Boures, D.; An, B.; Wang, C.; Lin, W.B. Metal–organic frameworks stabilize mono (phosphine)–metal complexes for broad-scope catalytic reactions. J. Am. Chem. Soc. 2016, 138, 9783–9786. [Google Scholar] [CrossRef]
  21. Zhao, X.; Li, J.; Jian, H.; Lu, M.; Wang, M. Two novel schiff base manganese complexes as bifunctional electrocatalysts for CO2 reduction and water oxidation. Molecules 2023, 28, 1074. [Google Scholar] [CrossRef] [PubMed]
  22. Takeda, H.; Cometto, C.; Ishitani, O.; Robert, M. Electrons, photons, protons and earth-abundant metal complexes for molecular catalysis of CO2 reduction. ACS Catal. 2017, 7, 70–88. [Google Scholar] [CrossRef]
  23. Zhang, L.; Chen, Z.; Liu, Z.P.; Bu, J.; Ma, W.X.; Yan, C.; Bai, R.; Lin, J.; Zhang, Q.Y.; Liu, J.Z.; et al. Efficient electrocatalytic acetylene semihydrogenation by electron–rich metal sites in N–heterocyclic carbene metal complexes. Nat. Commun. 2021, 12, 6574. [Google Scholar] [CrossRef] [PubMed]
  24. Lu, Y.L.; Song, J.Q.; Qin, Y.H.; Guo, J.; Huang, Y.H.; Zhang, X.D.; Pan, M.; Su, C.Y. A redox-active supramolecular Fe4L6 cage based on organic vertices with acid-base-dependent charge tunability for dehydrogenation catalysis. J. Am. Chem. Soc. 2022, 144, 8778–8788. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, Y.Z.; Shi, H.T.; Lee, C.-S.; Yiu, S.-M.; Man, W.-M.; Lau, T.-C. Room temperature aerobic peroxidation of organic substrates catalyzed by cobalt (Ш) alkylperoxo complexes. J. Am. Chem. Soc. 2021, 143, 14445–14450. [Google Scholar] [CrossRef]
  26. Cuerva, C.; Cano, M.; Lodeiro, C. Advanced functional luminescent metallomesogens: The key role of the metal center. Chem. Rev. 2021, 121, 12966–13010. [Google Scholar] [CrossRef]
  27. Nakano, M.; Oshio, H. Magnetic anisotropies in paramagnetic polynuclear metal complexes. Chem. Soc. Rev. 2011, 40, 3239–3248. [Google Scholar] [CrossRef]
  28. Kang, Y.-S.; Lu, Y.; Chen, K.; Zhao, Y.; Wang, P.; Sun, W.-Y. Metal–organic frameworks with catalytic centers: From synthesis to catalytic application. Coord. Chem. Rev. 2019, 378, 262–280. [Google Scholar] [CrossRef]
  29. Grossert, K.; Sicking, W.; Zellermann, E.; Schmuck, C. Hierarchical self-assembly of a small monomer with two orthogonal binding sites: From discrete hexagonal containers to a stimuli-responsive supramolecular gel. Supramol. Chem. 2018, 30, 395–403. [Google Scholar] [CrossRef]
  30. Evans, J.D.; Sumby, C.J.; Doonan, C.J. Post-synthetic metalation of metal–organic frameworks. J. Chem. Soc. Rev. 2014, 43, 5933–5951. [Google Scholar] [CrossRef]
  31. Gonzalez, M.I.; Turkiewicz, A.B.; Darago, L.E.; Oktawiec, J.; Bustillo, K.; Grandjean, F.; Long, G.J.; Long, J.R. Confinement of atomically defined metalhalide sheets in a metal–organic framework. Nature 2020, 577, 64–68. [Google Scholar] [CrossRef]
  32. Valvekens, P.; Blochb, E.D.; Long, J.R.; Ameloot, R.; De Vos, D.E. Counteranion effects on the catalytic activity of copper salts immobilized on the 2,2′-bipyridine-functionalized metal–organic framework MOF-253. Catal. Today 2014, 246, 55. [Google Scholar] [CrossRef] [Green Version]
  33. Gonzalez, M.I.; Bloch, E.D.; Mason, J.A.; Teat, S.J.; Long, J.R. Single-crystal-to-single-crystal metalation of a metal–organic framework: A route toward structurally well-defined catalysts. Inorg. Chem. 2015, 54, 2995–3005. [Google Scholar] [CrossRef]
  34. Li, S.; Li, G.-L.; Wang, W.; Liu, Y.; Cao, Z.-M.; Cao, X.-L. A 2D metal-organic framework interpenetrated by a 2D supramolecular framework assembled by CH/π interactions. Inorg. Chem. Commun. 2021, 130, 108705. [Google Scholar] [CrossRef]
  35. Li, L.-L.; Huang, M.; Chen, T.; Xu, X.-F.; Zhuo, Z.; Wang, W.; Huang, Y.-G. A porous π-stacked self-assembly of cup-shaped palladium complex for iodine capture. Molecules 2023, 28, 2881. [Google Scholar] [CrossRef]
  36. Chen, T.; Li, S.; Wang, Z.-B.; Wu, Z.-Y.; Huang, M.; Wang, W.; Zhuo, Z.; Huang, Y.-G. Iodine uptake enhanced electrical conductivity by a metal-organic framework bearing nanotube array of π-stacked columns. J. Mol. Struct. 2023, 1289, 135858. [Google Scholar] [CrossRef]
  37. Li, G.-L.; Zhuo, Z.; Wang, B.; Cao, X.-L.; Su, H.-F.; Wang, W.; Huang, Y.-G.; Hong, M.-C. Constructing π-stacked supramolecular cage based hierarchical self-assemblies via π···π stacking and hydrogen bonding. J. Am. Chem. Soc. 2021, 143, 10920–10929. [Google Scholar] [CrossRef]
  38. Erkabaev, A.M.; Yaroslavtseva, T.V.; Popov, S.E.; Bushkova, O.V. IR spectroscopic and quantum-chemical investigation of perchlorate anion solvation in acetonitrile. Russ. J. Phys. Chem. 2015, 89, 76–81. [Google Scholar] [CrossRef]
  39. Akhtar, M.; Alotaibi, M.A.; Alharthi, A.I.; Zierkiewicz, W.; Tahir, M.N.; Mazhar, M.; Isab, A.A.; Monim-ul-Mehboob, M.; Ahmad, S. Spectroscopic and DFT studies of zinc (II) complexes of diamines and thiocyanate; crystal structure of (cis-1,2-diaminocyclohexane)bis(thiocyanato-κN)zinc (II). J. Mol. Struct. 2017, 1128, 455–461. [Google Scholar] [CrossRef]
  40. Yang, E.C.; Hendrickson, D.N.; Wernsdorfer, W.; Nakano, M.; Zakharov, L.N.; Sommer, R.D.; Rheingold, A.L.; Ledezma-Gairaud, M.; Christou, G. Cobalt single-molecule magnet. J. Appl. Phys. 2002, 91, 7382–7384. [Google Scholar] [CrossRef]
  41. Ostrovsky, S.M.; Falk, K.; Pelikan, J.; Brown, D.A.; Tomkowicz, Z.; Haase, W. Orbital angular momentum contribution to the magneto-optical behavior of a binuclear cobalt (II) complex. Inorg. Chem. 2006, 2, 688–694. [Google Scholar] [CrossRef] [PubMed]
  42. Guo, L.R.; Zhu, F.; Chen, Y.; Li, Y.Z.; Zheng, L.M. Layered copper compounds based on 4-(3-bromothienyl) phosphonate (BTP): Weak ferromagnetism observed in [Cu2(4,4′-bpy)0.5(BTP)2]·H2O. Dalton Trans. 2009, 40, 8548–8554. [Google Scholar] [CrossRef] [PubMed]
  43. Azuma, M.; Kaimori, S.; Takano, M. High-pressure synthesis and magnetic properties of layered double perovskites Ln2CuMO6 (Ln = La, Pr, Nd, and Sm; M = Sn and Zr). Chem. Mater. 1998, 10, 3124–3130. [Google Scholar] [CrossRef]
  44. Zhang, S.R.; Du, D.Y.; Qin, J.S.; Bao, S.J.; Li, S.L.; He, W.W.; Lan, Y.Q.; Shen, P.; Su, Z.M. A fluorescent sensor for highly selective detection of nitroaromatic explosives based on a 2D, extremely stable, metal–organic framework. Chem. Eur. J. 2014, 20, 3589–3594. [Google Scholar] [CrossRef]
  45. Sheldrick, G.M. SHELXTL-integrated space-group and crystal-structure determination. Acta Crystallogr. 2015, A71, 3–8. [Google Scholar]
Figure 1. Hierarchical structure of the compound 1. (a) The structure of the isolated ligand L. (b) The cubic cage [H12L8]12+ formed by eight partially protonated ligands L (the π···π interactions are indicated by purple dotted lines). (c) The 3D hierarchical self-assembly with [N(C2H5)4]+ and ClO4 positioned at the interstitial regions between cubes. (inset: adjacent cubic cages with opposite chirality are denoted by ∆ and Λ, respectively).
Figure 1. Hierarchical structure of the compound 1. (a) The structure of the isolated ligand L. (b) The cubic cage [H12L8]12+ formed by eight partially protonated ligands L (the π···π interactions are indicated by purple dotted lines). (c) The 3D hierarchical self-assembly with [N(C2H5)4]+ and ClO4 positioned at the interstitial regions between cubes. (inset: adjacent cubic cages with opposite chirality are denoted by ∆ and Λ, respectively).
Molecules 28 04923 g001
Figure 2. The photographs showing the single-crystal-to-single-crystal transformation from compound 1 to compound 2.
Figure 2. The photographs showing the single-crystal-to-single-crystal transformation from compound 1 to compound 2.
Molecules 28 04923 g002
Figure 3. Structure of compound 2. (a) Structure of [CoSCNL]+. (b) The π-stacked cubes of {[CoSCNL]+}8 in compound 2 (the π···π interactions are indicated by purple dotted lines). (c) The 3D hierarchical structure of compound 2 with [N(C2H5)4]+ and SCN positioned at the interstitial regions between the cubes. (inset: adjacent cubic cages with opposite chirality are denoted by ∆ and Λ, respectively).
Figure 3. Structure of compound 2. (a) Structure of [CoSCNL]+. (b) The π-stacked cubes of {[CoSCNL]+}8 in compound 2 (the π···π interactions are indicated by purple dotted lines). (c) The 3D hierarchical structure of compound 2 with [N(C2H5)4]+ and SCN positioned at the interstitial regions between the cubes. (inset: adjacent cubic cages with opposite chirality are denoted by ∆ and Λ, respectively).
Molecules 28 04923 g003
Figure 4. Schematic diagram of unit cell and unit cell parameters changes of compound 1 and compound 2 before and after metalation through single-crystal-to-single-crystal transformation. (a) The unit cell and unit cell parameters of compound 1. (b) The unit cell and unit cell parameters of compound 2.
Figure 4. Schematic diagram of unit cell and unit cell parameters changes of compound 1 and compound 2 before and after metalation through single-crystal-to-single-crystal transformation. (a) The unit cell and unit cell parameters of compound 1. (b) The unit cell and unit cell parameters of compound 2.
Molecules 28 04923 g004
Figure 5. (a) PXRD patterns of compounds 1 and 2. (b) IR spectra of compounds 1 and 2. (c) The plot of χmTT and χm−1T of compound 2. (d) The plot of MH of compound 2.
Figure 5. (a) PXRD patterns of compounds 1 and 2. (b) IR spectra of compounds 1 and 2. (c) The plot of χmTT and χm−1T of compound 2. (d) The plot of MH of compound 2.
Molecules 28 04923 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, B.; Nan, Z.-A.; Liu, J.; Lu, Z.-X.; Wang, W.; Zhuo, Z.; Li, G.-L.; Huang, Y.-G. Metalation of a Hierarchical Self-Assembly Consisting of π-Stacked Cubes through Single-Crystal-to-Single-Crystal Transformation. Molecules 2023, 28, 4923. https://doi.org/10.3390/molecules28134923

AMA Style

Wang B, Nan Z-A, Liu J, Lu Z-X, Wang W, Zhuo Z, Li G-L, Huang Y-G. Metalation of a Hierarchical Self-Assembly Consisting of π-Stacked Cubes through Single-Crystal-to-Single-Crystal Transformation. Molecules. 2023; 28(13):4923. https://doi.org/10.3390/molecules28134923

Chicago/Turabian Style

Wang, Bin, Zi-Ang Nan, Jin Liu, Zi-Xiu Lu, Wei Wang, Zhu Zhuo, Guo-Ling Li, and You-Gui Huang. 2023. "Metalation of a Hierarchical Self-Assembly Consisting of π-Stacked Cubes through Single-Crystal-to-Single-Crystal Transformation" Molecules 28, no. 13: 4923. https://doi.org/10.3390/molecules28134923

Article Metrics

Back to TopTop