Two Porous Polyoxometalate-Resorcin[4]arene-Based

1 day ago - Isostructural 1 and 2 feature porous supramolecular architectures and represent an unusual example of a successful combination of ...
0 downloads 0 Views 5MB Size
Article pubs.acs.org/crystal

Cite This: Cryst. Growth Des. XXXX, XXX, XXX−XXX

Two Porous Polyoxometalate-Resorcin[4]arene-Based Supramolecular Complexes: Selective Adsorption of Organic Dyes and Electrochemical Properties Published as part of a Crystal Growth and Design virtual special issue on Crystalline Functional Materials in Honor of Professor Xin-Tao Wu Qiu-Yi Zhai,† Juan Su,‡ Ting-Ting Guo,† Jin Yang,† Jian-Fang Ma,*,† and Jie-Sheng Chen*,‡ Crystal Growth & Design Downloaded from pubs.acs.org by UNIV OF SOUTH DAKOTA on 09/14/18. For personal use only.



Key Laboratory of Polyoxometalate Science, Department of Chemistry, Northeast Normal University, Changchun 130024, P. R. China ‡ School of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Shanghai 200240, P. R. China S Supporting Information *

ABSTRACT: Two new inorganic−organic hybrid complexes, namely, {[Co4(L)2(HCOO)2(OH)2][SiO4(W3O9)4]}·6DMF· 5H2O (1) and {[Zn4(L)2(HCOO)2(OH)2][SiO4(W3O9)4]}· 6DMF·5H2O (2), have been solvothermally prepared by using a new methyl imidazole functionalized resorcin[4]arene (L), polyoxometalate (POM), and metal cation. Isostructural 1 and 2 feature porous supramolecular architectures and represent an unusual example of a successful combination of resorcin[4]arenes with polyoxometalates. Remarkably, selective adsorptions for organic dyes were studied by using 1 and 2 as adsorbents. Moreover, their electrocatalytic activities for the oxidation of ascorbic acid and reduction of NaBrO3 were also studied.



groups of the L ligand exhibit flexibility around the −S− groups and strong coordination ability with metal cations (Scheme 1).

INTRODUCTION Polyoxometalates (POMs), a class of nanosized metal-oxide clusters,1−4 have gained considerable attention owing to their potential applications in biochemistry, catalysis, magnetism, and electrochemistry.5−13 The POMs have been widely utilized as inorganic building blocks as well as guest counterions for constructing inorganic−organic hybrid complexes.14−20 By introducing POMs units, these hybrid materials acquired the advantages of both inorganic POMs and organic components, making them more potential applications in chemical separation and electrocatalysis.21−25 In this regard, functionalized calixarenes are good organic component candidates for assembly of the POMs-based metal−organic hybrid compounds.26 Calixarenes have received much more attention for their special structures and potential applications in biological medicine, catalysis, and molecular recognition etc.27−29 The calixarenes can be well modified on the bodies and rims with various substituents,30 thus yielding a great diversity of ligands.31−35 Thus far some inorganic−organic hybrid complexes with fascinating structures and properties have been synthesized by applying the calixarene-based ligands.36−40 In this facet, we designed a series of functional resorcin[4]arenes.41−46 By employing these resorcin[4]arenes as ligands, some coordination compounds, featuring fascinating structures and properties, have been successfully achieved.41−46 As a continuation of our work, we designed a new methyl imidazole functionalized resorcin[4]arene ligand (L). The imidazole © XXXX American Chemical Society

Scheme 1. Synthetic Procedure for L

By assembly of L, POMs, and metal cations, two new isostructural inorganic−organic hybrid complexes were obtained, namely, {[Co4(L)2(HCOO)2(OH)2][SiO4(W3O9)4]}·6DMF· 5H2O (1) and {[Zn4(L)2(HCOO)2(OH)2][SiO4(W3O9)4]}· 6DMF·5H2O (2). Remarkably, 1 and 2 show porous architectures and can be used for selective adsorptions of organic dyes as adsorbents. Moreover, they exhibit electrocatalytic activities for reduction of NaBrO3 and oxidation of ascorbic acid (AA). Received: June 11, 2018 Revised: August 27, 2018

A

DOI: 10.1021/acs.cgd.8b00891 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design



Article

EXPERIMENTAL SECTION

Table 1. Crystallographic Data and Structural Refinements for 1 and 2

Materials and Methods. Reagents were achieved commercially. Fourier transform infrared (FT-IR) spectrum was determined on a Mattson Alpha Centauri spectrometer. A Rigaku Dmax 2000 X-ray diffractometer with graphite monochromated CuKα radiation (λ = 0.154 nm) was applied to record powder X-ray diffraction (PXRD) patterns. UV−vis absorption spectrum was recorded on a TU-1900 UV/vis spectrophotometer. Elemental analytical data were determined on a PerkinElmer 2400CHN elemental analyzer. N2 adsorption− desorption experiment was carried out on an adsorption equipment (V-Sorb 2800S). Electrochemical behavior was determined with a CHI660b electrochemical workstation. Thermogravimetric curve was measured on a Q600 Simultaneous differential scanning calorimetry thermogravimetric analysis (DSC-TGA) analyzer under nitrogen gas. Synthesis of Functionalized Resorcin[4]arene (L). The precursor of (I) was prepared according to the known procedure.41,47,48 L was synthesized by the following method. The sample of (I) (0.96 g, 1 mmol), K2CO3 (2.21 g, 16 mmol), 1-methyl-1H-imidazole-2-thiol (0.91 g, 8 mmol), KI (0.1 g, 0.6 mmol), and DMF (250 mL) were placed in a three-necked flask (500 mL), and stirred at 80 °C for 12 h under N2. Then the resulting product was cooled and filtrated. The solvent was evaporated to afford a pale purple solid. The solid was further washed with water to afford white solid of L in a ca. 88% yield. IR data (KBr, cm−1): 3381(w), 2970(w), 2940(w), 1591(w), 1510(w), 1462(s), 1377(w), 1339(w), 1279(w), 1249(m), 1210(w), 1147(w), 1097(m), 1053(w), 1015(m), 975(s), 929(m), 749(m), 685(m), 640(w), 582(w), 497(w). Synthesis of {[Co4(L)2(HCOO)2(OH)2][SiO4(W3O9)4]}·6DMF· 5H2O (1). The ligand L (0.011 g, 0.01 mmol), H4[SiO4(W3O9)4]· xH2O (0.014 g, 0.005 mmol), CoCl2·6H2O (0.010 g, 0.04 mmol), EtOH (4 mL), and DMF (4 mL) were put into a Teflon reactor, and then heated at 110 °C for 72 h. Deep blue crystals of 1 were gained in a ca. 33% yield. Anal. calcd for C132H160O73N22S8Co4SiW12 (Mr = 5949.31): C, 26.65; H, 2.71; N, 5.18. Found: C, 27.12; H, 2.93; N, 5.69. IR data (KBr, cm−1): 3420(w), 3120(w), 2936(w), 1943(w), 1660(s), 1564(s), 1527(w), 1467(m), 1386(m), 1363(m), 1285(m), 1255(m), 1213(m), 1149(m), 1093(m), 1054(w), 1011(m), 968(s), 921(2), 884(w), 802(s), 688(w), 534(w), 482(w). Synthesis of {[Zn4(L)2(HCOO)2(OH)2][SiO4(W3O9)4]}·6DMF· 5H2O (2). A mixture of L (0.011 g, 0.01 mmol), H4[SiO4(W3O9)4]· xH2O (0.014 g, 0.005 mmol), Zn(NO3)·6H2O (0.012 g, 0.04 mmol), 3.5 M HNO3 (two drops), and DMF (8 mL) was placed in a Teflon reactor, and then heated at 110 °C for 72 h. Crystals were obtained in a ca. 42% yield. Anal. calcd for C132H160O73N22S8Zn4SiW12 (Mr = 5975.07): C, 26.53; H, 2.70; N, 5.16. Found: C, 27.23; H, 3.09; N, 5.92. IR data (KBr, cm−1): 3565(w), 3121(m), 2934(m), 2860(w), 1942(w), 1659(s), 1578(s), 1530(w), 1470(s), 1386(m), 1339(w), 1286(m), 1255(m), 1213(m), 1150(s), 1093(s), 1054(w), 1011(m), 968(s), 921(s), 885(m), 803(s), 688(w), 534(w). X-ray Crystallography. Crystallographic data for 1 and 2 were collected on an Oxford Diffraction Gemini R CCD with graphitemonochromated MoKα radiation (λ = 0.71073 Å) at 293 K. Absorption correction was carried out with a multiscan technique. The structures were solved by direct methods with SHELXS-97 and refined on F2 by full-matrix least-squares employing SHELXTL-97 within WINGX.49−51 Non-hydrogen atom was refined anisotropically. Hydrogen atom was generated geometrically (hydrogen atoms of O6 and C29 in 1 and O6 and C29 in 2 were not added to their models). The SQUEEZE function was used in refinements because of disordered solvents.52 Crystallographic data are given in Table 1. Selected bond lengths and bond angles are listed in Tables S1 and S2.

compound

1

2

formula Mr crystal system space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Z Dcalc (g cm−3) F(0 0 0) Rint GOF on F2 R1a [I > 2σ(I)] wR2b (all data)

C132H160N22O73S8SiW12Co4 5949.31 orthorhombic

C132H160N22O73S8SiW12Zn4 5975.07 orthorhombic

Cmca 37.329(2) 23.909(2) 21.4190(15) 90 90 90 19116(3) 8 4.134 22624 0.0497 0.979 0.0565

Cmca 37.332(5) 24.010(5) 21.376(5) 90.000(5) 90.000(5) 90.000(5) 19160(7) 8 4.143 22720 0.0478 1.078 0.0671

0.1685

0.1723

a

R1 = Σ∥Fo| − |Fc∥/Σ|Fo|. bwR2 = {Σ[w(Fo2 − Fc2)2]/Σw(Fo2)2]}1/2.

1 will be described as an example. The asymmetric unit of 1 consists of half a L, one Co(II) cation, half a coordinated HCOO− anion from the degradation of DMF, half a coordinated hydroxy, and a quarter of [α-SiW12O40]4− counteranion. Each Co(II) cation is coordinated with two nitrogen atoms of the same L ligand and two oxygen atoms of one hyxdroxy group and one HCOO− anion, with Co−N distances of 2.024 and 2.035 Å, and Co−O distances of 1.924 and 1.955 Å, respectively (Figure 1a). Two Co(II) cations are bridged by both HCOO− and OH− moieties to produce a Co(II) dimer with the Co(II)···Co(II) distance of 3.27 Å. The large [α-SiW12O40]4− anion is free and balances the negative charge as a counteranion. Interestingly, each [Co2(L)(HCOO)(OH)]2+ cation interacts with four [α-SiW12O40]4− anions (Figure 1b), and each [α-SiW12O40]4− anion is surrounded by eight [Co2(L)(HCOO)(OH)]2+ cations via the weak C−H···O hydrogen bonds formed by methyl H atoms of L and oxygen atoms of [α-SiW12O40]4− (Table S3), resulting in a supramolecular layer (Figure 1c,d). These supramolecular layers are further stacked together to produce a porous architecture with a channel dimension of ca. 7.2 Å × 7.2 Å along the a axis (Figure 1e). For 1 and 2, the total solvent-accessible volume are ca. 6249.9 and 6248.9 Å3, calculated by the PLATON,52 which accommodate ca. 32.7% and 32.6% of the unit cell volumes (19116.0 and 19160.0 Å3), respectively. Selective Adsorption of Organic Dyes. Currently, environmental pollution issues have received considerable attention.53 In this facet, water pollution is one of the most significant problems that influence human health seriously.53 Thus, the efficient removal of organic dye molecules from polluted water is highly desirable. In view of the porous architectures of 1 and 2, their adsorption properties for organic dyes thus were studied. In the adsorption process, the sizes of organic dyes are usually considered as one of the most important factors.54 In this work, different sizes of organic dyes, including methylene blue (MB), neutral red (NR), acid leather orange I (AO) and rhodamine 6G (R6G), were chosen to investigate the dye-capture ability of



RESULTS AND DISCUSSION Structural Description of 1 and 2. The formulas of 1 and 2 were established by diffuse electron density, TGA, as well as elemental analysis (Figure S1). Crystallographic diffraction shows that 1 and 2 are isostructural and crystallize in the same orthorhombic Cmca space group. Therefore, the structure of B

DOI: 10.1021/acs.cgd.8b00891 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 1. (a) Coordination environment of Co(II) cation in 1. (b) Hydrogen-bonding interaction mode of each [Co2(L)(HCOO)(OH)]2+ cation with four [α-SiW12O40]4− anions. (c) Hydrogen-bonding interaction mode of each [α-SiW12O40]4− anion with eight [Co2(L)(HCOO)(OH)]2+ cations. (d) Hydrogen-bonding supramolecular layer formed by [α-SiW12O40]4− anions and [Co2(L)(HCOO)(OH)]2+ cations. (e) View of the porous architecture stacked by neighboring supramolecular layers along the a axis. Symmetry code: #1 x, −y − 1, −z + 1.

1 and 2 (Scheme 2). The adsorption processes are typical as follows. The fresh crystalline samples of 1 or 2 (10 mg) were put into aqueous solutions of MB, NR, AO, and R6G (5.0 × 10−5 M), respectively. The contents of the organic dyes in aqueous solutions were determined by UV−vis absorption spectra. The concentrations of MB and NR in aqueous solution decreased drastically with adsorption time; meanwhile, the concentration of AO reduced moderately (Figure 2). However, the concentration of R6G was nearly a constant with time. After 48 h of adsorption, ca. 87.2% and 67.3% of MB, 92.0% and 80.0% of NR, and 47.9% and 39.2% of AO were taken by 1 and 2, respectively. Nevertheless, for R6G, no obvious solution colors or absorbance intensity changes were found for 1 and 2 (Figure 2d,h). The result indicates that the organic dye

Scheme 2. Organic Dyes Employed in This Work

adsorptions of 1 and 2 are highly selective. To further demonstrate the selective adsorption capacity, the samples of C

DOI: 10.1021/acs.cgd.8b00891 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 2. UV−vis absorption spectra of MB, NR, AO, and R6G adsorbed by 1 (a−d) and 2 (e−h) at the given time intervals, respectively. The photographs reveal color changes of the solutions and the samples before and after 48 h of dye adsorption.

Typically, the dye loaded samples of MB@1 and NR@1 (10 mg) were soaked into DMF (3 mL), respectively, and the contents of the organic dyes in DMF solution were monitored with UV/vis absorption spectra. As shown in Figure 4, the concentrations of the organic dyes in DMF were enhanced rapidly, and the solvents turned blue and yellow for MB and NR, respectively. Importantly, the samples can be reused for the next sorption− desorption cycles. The recycled adsorption and release

1 and 2 were soaked in aqueous solution of the mixed MB and R6G as an example. As illustrated in Figure 3, the solution colors of the mixed organic dyes changed from pale blue to royal purple. The adsorption peaks of MB declined significantly, while the absorbance intensity of R6G only decreased slightly. Moreover, the release of organic dye was also studied in this work. Herein, we took 1 with loaded MB or NR for example. D

DOI: 10.1021/acs.cgd.8b00891 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

framework channels and the sizes of organic dyes are critical factors for the dye accessibility. For NR and MB, they feature relatively short molecular dimensions and can be easily encapsulated into the channels of the supramolecular architectures. However, both 1 and 2 exhibit less adsorbing capacities for AO because of its relatively large molecular size dimension. In particular, the molecular sizes for R6G are too large to access the channels of 1 and 2. In addition, the anionic organic dyes with the small sizes, such as nuclear fast red (NFR) and methyl orange (MO), were also used for the adsorption experiments (Figure S2). No obvious adsorption behaviors for NFR were observed by using 1 as the typical adsorbent (Figure S3); meanwhile the concentrations of MO were only slightly reduced with increasing time. The result demonstrates that the charges of the organic dyes may also be a factor that influences the adsorption ability. Electrochemical Properties. 1 and 2 are insoluble in most organic solvents and water, and therefore their electrocatalytic properties were also studied.55 Electrochemical characters of 1-CPE and 2-CPE were determined in the mixed solutions of Na2SO4 (0.5 M) and H2SO4 (0.1 M) with a scan rate of 50 mV/s. For 1-CPE, four pairs of reversible redox peaks were observed, whereas three pairs were detected for 2-CPE (Figure 5a). The average peak potentials are at −0.30 V(I/I′), −0.69 V(II/II′), and −0.88 V(III/III′) for 1-CPE and −0.25 V(I/I′), −0.52 V(II/II′), and −0.78 V(III/III′) for 2-CPE, which may account for three consecutive two-electron redoxes of the central W of the [α-SiW12O40]4−.56 The peak potential at 0.41 V(IV/IV′) for 1-CPE is probably attributed to one electron redox process of Co(II)/Co(III) (Figure 5B).57 Moreover, cyclic voltammograms with different scan rates (50−500 mV/s) for 1-CPE and 2-CPE were also determined in the mixed solutions of Na2SO4 (0.5 M) and H2SO4 (0.1 M). The anodic peak potentials gradually moved in the positive direction. Accordingly, the cathodic ones shifted in a negative direction. Moreover, the scan rates are proportional to the redox peak currents, demonstrating that both CPEs adopt a surface-confined route (Figure 5b).58 In addition, bifunctional electrocatalytic activities were also explored for 1-CPE and 2-CPE. As illustrated in Figure 5c, the reduction peak currents were enhanced with the increasing concentrations of NaBrO3. Accordingly, the oxidation peak currents decreased. The result demonstrates that BrO3− was reduced by the POMs. Moreover, the oxidation of ascorbic acid (AA) was also studied (Figure 5d). With the increasing concentrations of the AA, the anodic peak currents increased substantially, suggesting that both CPEs exhibit catalytic oxidation ability to the AA.

Figure 3. UV−vis absorption spectra of aqueous solutions with the mixed MB and R6G adsorbed by 1 (a) and 2 (b) at the given time intervals, respectively. The photographs reveal color changes of the solutions and the crystalline samples before and after 48 h of dye adsorption.



CONCLUSIONS In summary, two porous POMs-resorcin[4]arene-based supramolecular complexes have been successfully prepared by applying a new thiol-containing resorcin[4]arene ligand under solvothermal conditions. In 1 and 2, each [M2(L)(HCOO)(OH)]2+ (M = Co(II) or Zn(II)) cation interacts with four [α-SiW12O40]4− anions, and every [α-SiW12O40]4− anion is surrounded by eight [M2(L)(HCOO)(OH)]2+ cations through the weak C−H···O hydrogen bonds, thus producing a supramolecular layer. Adjacent supramolecular layers are further stacked to yield a porous architecture. Both 1 and 2 show selective and reversible adsorption capacities for organic dyes. Strikingly, 1 and 2 can be used as the recycled sorption−desorption adsorbents for the organic dyes. Remarkably, they also may be employed as bifunctional electrocatalysts for the oxidation of

Figure 4. UV−vis absorption spectra of DMF solutions with MB@1 or NR@1 after the dye release in DMF at the given time intervals.

experiments were carried out four times. The adsorption efficiency was slightly reduced in the fourth cycle. The dye adsorption results indicate that the dye adsorption capacities of 1 and 2 are highly size-dependent. The diameters of E

DOI: 10.1021/acs.cgd.8b00891 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 5. Cyclic voltammograms for 1-CPE and 2-CPE in mixed solutions of Na2SO4 (0.5 M) and H2SO4 (0.1 M): (a) at 50 mV/s, (b) at different scan rates from 50 to 500 mV/s, (c) with different concentrations of NaBrO3 at 50 mV/s, and (d) with different concentrations of AA at 50 mV/s.

AA and reduction of NaBrO3. This work presents an unusual example for assembly of the functional inorganic−organic hybrid complexes by combination of resorcin[4]arenes with POMs and metal cations.



Figures, TG, IR spectra, PXRD, and tables (PDF) Accession Codes

CCDC 1848499−1848500 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/ acs.cgd.8b00891. F

DOI: 10.1021/acs.cgd.8b00891 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design



Article

Metal-Organic Framework and Polyoxometalates. J. Am. Chem. Soc. 2009, 131, 1883−1888. (15) Liu, B.; Yang, J.; Yang, G. C.; Ma, J.-F. Four New ThreeDimensional Polyoxometalate-Based Metal-Organic Frameworks Constructed from [Mo6O18(O3AsPh)2]4‑ Polyoxoanions and Copper(I)Organic Fragments: Syntheses, Structures, Electrochemistry, and Photocatalysis Properties. Inorg. Chem. 2013, 52, 84−94. (16) Liu, H.-Y.; Wu, H.; Yang, J.; Liu, Y.-Y.; Liu, B.; Liu, Y.-Y.; Ma, J.-F. pH-Dependent Assembly of 1D to 3D Octamolybdate Hybrid Materials Based on a New Flexible Bis-[(pyridyl)-benzimidazole] Ligand. Cryst. Growth Des. 2011, 11, 2920−2927. (17) Du, D. Y.; Qin, J. S.; Li, S. L.; Su, Z. M.; Lan, Y. Q. Recent Advances in Porous Polyoxometalate-Based Metal-Organic Framework Materials. Chem. Soc. Rev. 2014, 43, 4615−4632. (18) Song, Y. F.; Tsunashima, R. Recent Advances on Polyoxometalate-Based Molecular and Composite Materials. Chem. Soc. Rev. 2012, 41, 7384−7402. (19) Zheng, S. T.; Zhang, J.; Yang, G. Y. Designed Synthesis of POMOrganic Frameworks from {Ni6PW9} Building Blocks under Hydrothermal Conditions. Angew. Chem., Int. Ed. 2008, 47, 3909−3913. (20) Li, X. X.; Wang, Y. X.; Wang, R. H.; Cui, C. Y.; Tian, C. B.; Yang, G. Y. Designed Assembly of Heterometallic Cluster Organic Frameworks Based on Anderson-Type Polyoxometalate Clusters. Angew. Chem., Int. Ed. 2016, 55, 6462−6466. (21) Jin, L.; Fang, Y.; Hu, P.; Zhai, Y.; Wang, E.; Dong, S. Polyoxometalate-Based Inorganic-Organic Hybrid Film Structure with Reversible Electroswitchable Fluorescence Property. Chem. Commun. 2012, 48, 2101−2103. (22) Moussawi, M. A.; Leclerc-Laronze, N.; Floquet, S.; Abramov, P. A.; Sokolov, M. N.; Cordier, S.; Ponchel, A.; Monflier, E.; Bricout, H.; Landy, D.; Haouas, M.; Marrot, J.; Cadot, E. Polyoxometalate, Cationic Cluster, and gamma-Cyclodextrin: From Primary Interactions to Supramolecular Hybrid Materials. J. Am. Chem. Soc. 2017, 139, 12793−12803. (23) Yazigi, F.-J.; Wilson, C.; Long, D.-L.; Forgan, R. S. Synthetic Considerations in the Self-Assembly of Coordination Polymers of Pyridine-Functionalized Hybrid Mn-Anderson Polyoxometalates. Cryst. Growth Des. 2017, 17, 4739−4748. (24) Li, X.-X.; Ma, X.; Zheng, W.-X.; Qi, Y.-J.; Zheng, S.-T.; Yang, G.Y. Composite Hybrid Cluster Built from the Integration of Polyoxometalate and a Metal Halide Cluster: Synthetic Strategy, Structure, and Properties. Inorg. Chem. 2016, 55, 8257−8259. (25) Zhao, J.-W.; Li, Y.-Z.; Chen, L.-J.; Yang, G.-Y. Research Progress on Polyoxometalate-Based Transition-Metal-Rare-Earth Heterometallic Derived Materials: Synthetic Strategies, Structural Overview and Functional Applications. Chem. Commun. 2016, 52, 4418−4445. (26) Lu, B.-B.; Yang, J.; Che, G.-B.; Pei, W.-Y.; Ma, J.-F. Highly Stable Copper(I)-Based Metal-Organic Framework Assembled with Resorcin[4]arene and Polyoxometalate for Efficient Heterogeneous Catalysis of Azide-Alkyne “Click” Reaction. ACS Appl. Mater. Interfaces 2018, 10, 2628−2636. (27) Kim, H. J.; Lee, M. H.; Mutihac, L.; Vicens, J.; Kim, J. S. HostGuest Sensing by Calixarenes on The Surfaces. Chem. Soc. Rev. 2012, 41, 1173−1190. (28) Liu, M.; Liao, W.; Hu, C.; Du, S.; Zhang, H. Calixarene-Based Nanoscale Coordination Cages. Angew. Chem., Int. Ed. 2012, 51, 1585− 1588. (29) La Manna, P.; Talotta, C.; Floresta, G.; De Rosa, M.; Soriente, A.; Rescifina, A.; Gaeta, C.; Neri, P. Mild Friedel-Crafts Reactions inside a Hexameric Resorcinarene Capsule: C-Cl Bond Activation through Hydrogen Bonding to Bridging Water Molecules. Angew. Chem., Int. Ed. 2018, 57, 5423−5428. (30) Kobayashi, K.; Yamanaka, M. Self-Assembled Capsules Based on Tetrafunctionalized Calix[4]resorcinarene Cavitands. Chem. Soc. Rev. 2015, 44, 449−466. (31) Galan, A.; Ballester, P. Stabilization of Reactive Species by Supramolecular Encapsulation. Chem. Soc. Rev. 2016, 45, 1720−1737. (32) Korom, S.; Martin, E.; Serapian, S. A.; Bo, C.; Ballester, P. Molecular Motion and Conformational Interconversion of IrI·COD

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (J.-F.M.). *E-mail: [email protected] (J.-S.C.). ORCID

Jian-Fang Ma: 0000-0002-4059-8348 Jie-Sheng Chen: 0000-0003-1233-7746 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China for support (Grant No. 21471029).



DEDICATION Dedicated to Professor Xin-Tao Wu on the occasion of his 80th birthday.



REFERENCES

(1) Long, D. L.; Burkholder, E.; Cronin, L. Polyoxometalate Clusters, Nanostructures and Materials: from Self Assembly to Designer Materials and Devices. Chem. Soc. Rev. 2007, 36, 105−121. (2) Miras, H. N.; Yan, J.; Long, D. L.; Cronin, L. Engineering Polyoxometalates with Emergent Properties. Chem. Soc. Rev. 2012, 41, 7403−7430. (3) Long, D. L.; Tsunashima, R.; Cronin, L. Polyoxometalates: Building Blocks for Functional Nanoscale Systems. Angew. Chem., Int. Ed. 2010, 49, 1736−1758. (4) Proust, A.; Matt, B.; Villanneau, R.; Guillemot, G.; Gouzerh, P.; Izzet, G. Functionalization and Post-Functionalization: A Step towards Polyoxometalate-Based Materials. Chem. Soc. Rev. 2012, 41, 7605− 7622. (5) Proust, A.; Thouvenot, R.; Gouzerh, P. Functionalization of Polyoxometalates: towards Advanced Applications in Catalysis and Materials Science. Chem. Commun. 2008, 1837−1852. (6) Kortz, U.; Müller, A.; van Slageren, J.; Schnack, J.; Dalal, N. S.; Dressel, M. Polyoxometalates: Fascinating Structures, Unique Magnetic Properties. Coord. Chem. Rev. 2009, 253, 2315−2327. (7) Stracke, J. J.; Finke, R. G. Distinguishing Homogeneous from Heterogeneous Water Oxidation Catalysis when Beginning with Polyoxometalates. ACS Catal. 2014, 4, 909−933. (8) Xu, J.; Li, X.; Li, X.; Li, B.; Wu, L.; Li, W.; Xie, X.; Xue, R. Supramolecular Copolymerization of Short Peptides and Polyoxometalates: toward the Fabrication of Underwater Adhesives. Biomacromolecules 2017, 18, 3524−3530. (9) Clemente-Juan, J. M.; Coronado, E.; Gaita-Arino, A. Magnetic Polyoxometalates: from Molecular Magnetism to Molecular Spintronics and Quantum Computing. Chem. Soc. Rev. 2012, 41, 7464− 7478. (10) Wang, S.-S.; Yang, G.-Y. Recent Advances in PolyoxometalateCatalyzed Reactions. Chem. Rev. 2015, 115, 4893−4962. (11) Wang, Y. J.; Zhou, Y. Y.; Hao, H. G.; Song, M.; Zhang, N.; Yao, S.; Yan, J. H.; Zhang, Z. M.; Lu, T. B. Capped Polyoxometalate Pillars between Metal-Organic Layers for Transferring a Supramolecular Structure into a Covalent 3D Framework. Inorg. Chem. 2018, 57, 1342− 1349. (12) Sha, J.-Q.; Li, M.-T.; Yang, X.-Y.; Sheng, N.; Li, J.-S.; Zhu, M.-L.; Liu, G.-D.; Jiang, J. New Route toward POM[6]Catenane Members for Lithium-Ion Batteries. Cryst. Growth Des. 2017, 17, 3775−3782. (13) Dong, J.; Hu, J.; Chi, Y.; Lin, Z.; Zou, B.; Yang, S.; Hill, C. L.; Hu, C. A Polyoxoniobate−Polyoxovanadate Double-Anion Catalyst for Simultaneous Oxidative and Hydrolytic Decontamination of Chemical Warfare Agent Simulants. Angew. Chem., Int. Ed. 2017, 56, 4473−4477. (14) Sun, C.-Y.; Liu, S.-X.; Liang, D.-D.; Shao, K.-Z.; Ren, Y.-H.; Su, Z.-M. Highly Stable Crystalline Catalysts Based on a Microporous G

DOI: 10.1021/acs.cgd.8b00891 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Included in Rebek’s Self-Folding Octaamide Cavitand. J. Am. Chem. Soc. 2016, 138, 2273−2279. (33) Coletta, M.; McLellan, R.; Murphy, P.; Leube, B. T.; Sanz, S.; Clowes, R.; Gagnon, K. J.; Teat, S. J.; Cooper, A. I.; Paterson, M. J.; Brechin, E. K.; Dalgarno, S. J. Bis-Calix[4]arenes: From Ligand Design to the Directed Assembly of a Metal-Organic Trigonal Antiprism. Chem. - Eur. J. 2016, 22, 8791−8795. (34) Ovsyannikov, A.; Solovieva, S.; Antipin, I.; Ferlay, S. Coordination Polymers Based on Calixarene Derivatives: Structures and Properties. Coord. Chem. Rev. 2017, 352, 151−186. (35) Nakamura, M.; Kishimoto, K.; Kobori, Y.; Abe, T.; Yoza, K.; Kobayashi, K. Self-Assembled Molecular Gear: A 4:1 Complex of Rh(III)Cl Tetraarylporphyrin and Tetra(p-pyridyl)cavitand. J. Am. Chem. Soc. 2016, 138, 12564−12577. (36) Zhang, S. T.; Yang, J.; Wu, H.; Liu, Y. Y.; Ma, J.-F. Systematic Investigation of High-Sensitivity Luminescent Sensing for Polyoxometalates and Iron(III) by MOFs Assembled with a New Resorcin[4]arene-Functionalized Tetracarboxylate. Chem. - Eur. J. 2015, 21, 15806−15819. (37) Zhao, S. S.; Yang, J.; Liu, Y. Y.; Ma, J.-F. Fluorescent Aromatic Tag-Functionalized MOFs for Highly Selective Sensing of Metal Ions and Small Organic Molecules. Inorg. Chem. 2016, 55, 2261−2273. (38) Imamura, T.; Maehara, T.; Sekiya, R.; Haino, T. Frozen Dissymmetric Cavities in Resorcinarene-Based Coordination Capsules. Chem. - Eur. J. 2016, 22, 3250−3254. (39) Mendez-Arroyo, J.; d’Aquino, A. I.; Chinen, A. B.; Manraj, Y. D.; Mirkin, C. A. Reversible and Selective Encapsulation of Dextromethorphan and beta-Estradiol Using an Asymmetric Molecular Capsule Assembled via the Weak-Link Approach. J. Am. Chem. Soc. 2017, 139, 1368−1371. (40) Chen, L.; Chen, Q. H.; Wu, M. Y.; Jiang, F. L.; Hong, M. C. Controllable Coordination-Driven Self-Assembly: From Discrete Metallocages to Infinite Cage-Based Frameworks. Acc. Chem. Res. 2015, 48, 201−210. (41) Pei, W. Y.; Xu, G.; Yang, J.; Wu, H.; Chen, B.; Zhou, W.; Ma, J.-F. Versatile Assembly of Metal-Coordinated Calix[4]resorcinarene Cavitands and Cages through Ancillary Linker Tuning. J. Am. Chem. Soc. 2017, 139, 7648−7656. (42) Han, X.; Jin, S.-S.; Ma, J.-F.; Yang, J.; Mak, T. C. W. Silver(I) Cage and Infinite Chain Stabilized by Bowl-Shaped Resorcin[4]arene Tetraethynide Ligands. Cryst. Growth Des. 2016, 16, 3811−3817. (43) Chen, C.; Ma, J.-F.; Liu, B.; Yang, J.; Liu, Y.-Y. Two Unusual 3D Copper(II) Coordination Polymers Constructed by p-Sulfonated Calixarenes and Bis(triazolyl) Ligands. Cryst. Growth Des. 2011, 11, 4491−4497. (44) Lv, L. L.; Yang, J.; Zhang, H. M.; Liu, Y. Y.; Ma, J.-F. Metal-Ion Exchange, Small-Molecule Sensing, Selective Dye Adsorption, and Reversible Iodine Uptake of Three Coordination Polymers Constructed by A New Resorcin[4]arene-Based Tetracarboxylate. Inorg. Chem. 2015, 54, 1744−1755. (45) Lu, B.-B.; Jiang, W.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. Resorcin[4]arene-Based Microporous Metal-Organic Framework as An Efficient Catalyst for CO2 Cycloaddition with Epoxides and Highly Selective Luminescent Sensing of Cr2O72−. ACS Appl. Mater. Interfaces 2017, 9, 39441−39449. (46) Lu, B.-B.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. A PolyoxovanadateResorcin[4]arene-Based Porous Metal-Organic Framework as an Efficient Multifunctional Catalyst for the Cycloaddition of CO2 with Epoxides and the Selective Oxidation of Sulfides. Inorg. Chem. 2017, 56, 11710−11720. (47) Tunstad, L. M.; Tucker, J. A.; Dalcanale, E.; Weiser, J.; Bryant, J. A.; Sherman, J. C.; Helgeson, R. C.; Knobler, C. B.; Cram, D. J. HostGuest Cmplexation. 48. Octol Building Blocks for Cavitands and Carcerands. J. Org. Chem. 1989, 54, 1305−1312. (48) Boerrigter, H.; Verboom, W.; Reinhoudt, D. N. Novel Resorcinarene Cavitand-Based CMP(O) Cation Ligands: Synthesis and Extraction Properties. J. Org. Chem. 1997, 62, 7148−7155. (49) Sheldrick, G. M. SHELXL-97, Programs for X-ray Crystal Structure Solution; University of Göttingen: Göttingen, Germany, 1997.

(50) Sheldrick, G. M. SHELXL-97, Programs for X-ray Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 1997. (51) Farrugia, L. J. WINGX: A Windows Program for Crystal Structure Analysis; University of Glasgow: Glasgow, UK, 1988. (52) Spek, A. L. Single-Crystal Structure Validation with The Program PLATON. J. Appl. Crystallogr. 2003, 36, 7−13. (53) He, Y. C.; Yang, J.; Kan, W. Q.; Zhang, H. M.; Liu, Y. Y.; Ma, J.-F. A New Microporous Anionic Metal-Organic Framework as A Platform for Highly Selective Adsorption and Separation of Organic Dyes. J. Mater. Chem. A 2015, 3, 1675−1681. (54) Zhao, X.; Bu, X.; Wu, T.; Zheng, S. T.; Wang, L.; Feng, P. Selective Anion Exchange With Nanogated Isoreticular Positive MetalOrganic Frameworks. Nat. Commun. 2013, 4, 2344. (55) Dolbecq, A.; Mialane, P.; Keita, B.; Nadjo, L. PolyoxometalateBased Materials for Efficient Solar and Visible Light Harvesting: Application to The Photocatalytic Degradation of Azo Dyes. J. Mater. Chem. 2012, 22, 24509−24521. (56) Sadakane, M.; Steckhan, E. Electrochemical Properties of Polyoxometalates as Electrocatalysts. Chem. Rev. 1998, 98, 219−238. (57) Gao, G.-G.; Xu, L.; Wang, W.-J.; Qu, X.-S.; Liu, H.; Yang, Y.-Y. Cobalt(II)/Nickel(II)-Centered Keggin-Type Heteropolymolybdates: Syntheses, Crystal Structures, Magnetic and Electrochemical Properties. Inorg. Chem. 2008, 47, 2325−2333. (58) Antonio, M. R.; Chiang, M.-H. Stabilization of Plutonium(III) in The Preyssler Polyoxometalate. Inorg. Chem. 2008, 47, 8278−8285.

H

DOI: 10.1021/acs.cgd.8b00891 Cryst. Growth Des. XXXX, XXX, XXX−XXX