Efficient Catalytic Performance for Acylation-Nazarov Cyclization

Jul 31, 2018 - Synopsis. The efficient catalytic performance for acylation-Nazarov cyclization is rendered by an unusual postsynthetic oxidation strat...
0 downloads 0 Views 4MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

Efficient Catalytic Performance for Acylation-Nazarov Cyclization Based on an Unusual Postsynthetic Oxidization Strategy in a Fe(II)MOF Zhichao Shao,† Mengjia Liu,† Jian Dang,† Chao Huang,‡ Wenjuan Xu,† Jie Wu,*,† and Hongwei Hou*,†

Downloaded via DURHAM UNIV on July 31, 2018 at 23:27:48 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



The College of Chemistry and Molecular Engineering, Zhengzhou University, Zhengzhou, Henan 450001, People’s Republic of China ‡ Center for Advanced Materials Research, Zhongyuan University of Technology, Zhengzhou 450007, People’s Republic of China S Supporting Information *

ABSTRACT: A rare example of SC-SC triggered by Cu 2+ heterogeneous oxidation was demonstrated in a Fe(II)-based MOF {[FeII3(L)2(H2O)6]·3H2O}n (1), which occurred a slow conversion into an oxidized MOF 2 ({[FeIII3(L)2(H2O)6]·3(OH)}n) with retention of single crystallinity. The FeII → FeIII progress of the reaction oxidation was proved by single crystal XRD, PXRD, XPS, 57 Fe Mössbauer spectroscopy, and UV−vis. We used 1 and 2 as catalysts to catalyze the tandem Nazarov cyclization, and the results show that acylation products were only harvested when 1 was a catalyst, while the oxidized transformer 2 lead mainly to the formation of cyclization products under identical conditions. Through the test of different substrates, 2 can be a good heterogeneous catalyst candidate for the formation of cyclopentenone[b] benzenes. This work provides a new way to design efficient and hard-synthesized heterogeneous catalyst materials.



INTRODUCTION

As functional carriers, the different chemical valences of center metals are responsible for the discrepancy in material performance.35,36 This feature may be most salient in the activities of catalyst materials.37 The MOFs-based catalysts have been rapidly developed because of high catalytic activity, recyclable, and preeminent selectivity.38−42 Among them, the efficient catalyst for acylation-Nazarov cyclization to form cyclopentenone[b] has attracted the interests of researchers, since cyclopentenone[b] is an important intermediate for the synthetic natural medication. Encouraged by the above merits, our laboratory has been interested in developing efficient MOF catalysts for tandem Nazarov cyclization, and some porous MOFs materials have been reported for the synthesis of cyclopenta[b] pyrroles by treatment of pyrrole derivatives with α,β-unsaturated carboxylic acids.43,44 However, these catalysts can not play a good catalytic role in the benzene system, because the electron cloud density of benzene is relatively low, and the activity of substrate is not positive. MOF catalysts materials with a high valence state of Lewis acid will be an excellent candidate to catalyze the Nazarov cyclization of the benzene system substrate.

Functional MOFs crystalline materials have recently been of considerable interest due to porous characteristics, intriguing topologies, and potential multifield applications (ferroelectric, gas adsorption, luminescence, magnetism, catalysis, etc.).1−15 In particular, MOFs used for postsynthetic modification (PSM) and molecular reactor can fully highlight the functional feature of MOFs carrier and emphasize the swappable trend of material performance.16−22 In the previous reports about PSM, the modification methods mainly involve modification of linkers,23−25 transformation of metal-containing nodes,26−29 and effect of guest molecular.30,31 Among them, the postsynthetic oxidization (PSO) of metal ions as a simple and direct method has attracted the attention and interest of scholars, which can obtain unattainable highly valence ions MOFs crystalline materials.32,33 However, the strong oxidizing agent (such as oxygen, H2O2, •OH radicals, I2) involves the drastic oxidation process, which easily causes the collapse of structures and the loss of single crystallinity.34 Weak oxidant examples (especially highly charged metal ions) were rarely found. Therefore, exploration of the mild oxidation system is imperative, which can help to maintain good conditions of the MOFs material single crystal and regulate the functional properties. © XXXX American Chemical Society

Received: May 23, 2018

A

DOI: 10.1021/acs.inorgchem.8b01421 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Herein, we report the assembly of a three-dimensional (3D) FeII MOF, namely, {[Fe3L2(H2O)6]·3H2O}n (1) (H3L = 1aminobenzene-3,4,5-tricarboxylic acid). Upon SC-SC heterogeneous oxidation of Cu2+ solution, the FeII MOF 1 transforms to the FeIII framework of {[Fe3L2(H2O)6]·3(OH)}n (2), which is confirmed by 57Fe Mössbauer, XPS, and variable temperature susceptibility. The experimental results indicate that 1 and 2 showed a rare example of SC-SC oxidation behavior of FeII → FeIII driven by free Cu2+ ions. In addition, 1 and 2 are used as catalysts for the formation of cyclopentenone[b] benzenes with 1,3-dimethoxybenzene. The acylation products were obtained when 1 as a catalyst, while 2 can effectively catalyze the synthesis of the cyclization product under the same conditions.



Table 1. Summary of Crystallographic Data for 1 and 2

EXPERIMENTAL SECTION

In this work, the reagents and solvents were used in commercial purchase without further purification. PXRD patterns were collected on a PANalytical X’Pert PRO diffractometer. The element content (C, N, and H) was obtained by a FLASH EA 1112 elemental analyzer. Thermal stability were measured on a Netzsch STA 449C thermal analyzer. Solid ultraviolet diffuse reflection were carried out by JASCO-750 UV−vis spectrophotometer. The IR spectra were obtained on a Bruker Tensor 27 spectrophotometer. 1H NMR spectra were verified with a Bruker Avance 400 spectrometer. Mössbauer data were tested with a Wissel Mössbauer spectrometer with a 57Co source. Preparation of MOFs. Synthesis of {[FeII3L2(H2O)6]·3H2O}n (1). A total of 0.1 mmol FeSO4·7H2O (0.027 g), 0.05 mmol H3L (0.011 g), and 0.1 mmol NaOH (0.004 g) were added into acetonitrile (2 mL) and H2O (3 mL). The system was stirred and sealed in a 10 mL glass vial at 100 °C for 2 day. Single crystals appear as yellow clumps with a yield of 84% based on H3L. The theoretical value of elemental analysis (%) for C18H26Fe3N2O21: C, 28.01; H, 3.34; N, 3.60. The experimental value: C, 28.87; H, 3.21; N, 3.69. IR: 3540m, 1593s, 1391m, 1313m, 1217m, 1091s, 835w, 774w, 660w. Synthesis of {[FeIII3L2(H2O)6]·3(OH)}n (2). The yellow clump crystals were immersed in acetonitrile solution of Cu(NO3)2·3H2O. The system was placed in a 10 mL glass vial at 80 °C for 2 day. Crystal color changed from yellow to brown. The theoretical value of elemental analysis for C18H26Fe3N2O21 (2) is C, 27.91; H, 2.97; N, 3.61. The experimental value is C, 28.47; H, 2.45; N, 3.80. IR: 3424m, 1614m, 1556m, 1461m, 1389s, 1303s, 1107w, 835w, 732w, 655w. Single Crystal X-ray Crystallography. The structures of 1 and 2 were tested on a Bruker D8 VENTURE diffractometer with Mo Kα radiation (λ = 0.71073 Å). The SAINT program was used to control the integration of the diffraction data, polarization effects, and the intensity corrections for the Lorentz.45 We used the SADABS program to perform semiempirical absorption correction46 and solved structures by immediate ways and refined by a full matrix least-squares technique relied on F2 with the SHELXL-1997 software package.47 The hydrogen atoms were generated geometrically and refined isotropically using the riding model. The summary of crystallographic data for 1 and 2 is listed in Table 1. Corresponding bond lengths (Å) and bond angles (deg) are provided in Table S2. General Process of a Tandem Acylation-Nazarov Cyclization. A mixture of 1,3-dimethoxybenzene (1.0 mmol, 1.0 equiv), α,βunsaturated carboxylic acid (4a−d; 1.5 mmol), catalyst (0.1 equiv calculated according to the content of Fe), trifluoroacetic anhydride (TFAA, 4.0 mmol), and 10 mL of 1,2-dichloroethane (DCE) were added in a 25 mL round-bottom flask. Then the system was heated at 90 °C for 5 h and gradually cooled to room temperature. The DCE was removed in vacuo and the grease was extracted by 15 mL of EtOAC. The EtOAC phase was merged and washed with saturated aqueous NaHCO3, dried by Na2SO4, filtered, and evaporated under reduced pressure. The crude residue were refined by column chromatography on silica gel.



compound formula Fw T (K) l (Mo−Ka)/Å crystsyst space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Z Dcalcd (g cm−3) F(000) μ (mm−1) GOF R1 (I > 2σ(I)) wR2 (I > 2σ(I))

1 C18H26Fe3N2O21 773.9 293(2) 0.71073 monoclinic Pn 11.6282(8) 7.6006(5) 14.2257(9) 90 101.488(2) 90 1232.1(4) 2 2.086 788 1.853 1.068 0.0513 0.1546

2 C18H23Fe3N2O21 770.9 293(2) 0.71073 monoclinic Pn 11.675(7) 7.626(4) 14.289(9) 90 101.528(19) 90 1246.5(13) 2 2.054 782 1.831 1.040 0.0587 0.1723

RESULTS AND DISCUSSION Crystal Structure. We conducted single crystal X-ray diffraction analysis of MOF 1, and the result shows that 1 crystallizes in the monoclinal crystal system and Pn space group. In each basic structural unit, there are three Fe2+, two L3−, six coordinated H2O, and three free H 2O. The coordination models of Fe1 and Fe2 display similar octahedron configurations, and the Fe2+ ions are surrounded by six oxygen atoms from two L3− and three H2O moleculars (Figure 1). For Fe3, the ion is also surrounded by six atoms, and the equatorial plane is occupied by four O atoms from carboxyl group and two N from the amino group in the axial sites. The coordination length in 1 is examined, and the results show that Fe−O lengths both are from 2.076 to 2.318 Å and the length of Fe−N is 2.280(7) and 2.191 Å. These values are similar to other reported Fe complexes.48−50 The L3− in 1 is completely deprotonated and attached to three metal ions. The three carboxylic groups acting as μ2η1:η1, μ1-η1:η0, and μ1−η1:η1 modes, respectively. As shown in Figure 1, three Fe2+ ions are connected to each other to form a trinuclear unit [Fe3(CO2)2] by two carboxylate groups. The different trinuclears are interconnected each other via the carboxyl and amino to afford a 3D supramolecular framework. For a better understanding of this supramolecular framework, the structure of 1 is further analyzed. Ligand L3−, trinuclear [Fe3(CO2)2] unit can be considered as 3-connecting nodes and 6-connected nodes, respectively. Then the structure can be simplified to a 3D (3,6)-connected network and corresponding to the Schläfli symbol is 4(6)6(6)8(3)4(3). Study of 1 → 2 Transformation. In order to obtain FeMOF material with high valence state, the single crystals of 1 were dried and immersed in Cu(NO3)2 solution at 80 °C for 2 days. The crystal of 1 is transformed to oxidized complex with a good single crystallinity. In the process, the color of the crystal changes slowly from yellow to brown. Corresponding, the color of the solution became shallow and white precipitate was generated and observed after adding NaCl to the oxidation system, suggesting concentration of Cu2+ decrease and part of them may changed into Cu+, which was proved by ultraviolet B

DOI: 10.1021/acs.inorgchem.8b01421 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. View of 3D network of 1.

Figure 2. Elemental mapping images of 1 and 2.

spectra (Figure S2) and cyclic voltammograms curve (Figure S3). In order to rule out the decrease of Cu2+ coming from deposit or adsorption, EDS spectra of 2 were measured. As shown in Figure 2, the color of Cu is not uniformly distributed in the crystal shape, and the content of Cu is only 0.11%. In order to further prove the result, the atomic absorption experiment of 2 was carried out, the results show that there is 0.18% Cu2+ in 2, which is similar to the result of EDS spectra. Therefore, the decrease of Cu2+ is not coming from deposit or adsorption. Moreover, some important controlled experiments were also conducted: (1) when H2O2 was used as oxidant, similar brown crystals were obtained without good single crystalline state; (2) the trace amounts of oxygen in Cu2+ solution was removed in advance, then MOF 1 was added in the system under N2 protection. The oxidation results were not affected; (3) when the free radicals blocking agent was injected to the system, the yellow crystal 1 was still oxidized to MOF 2. These experimental results show that Cu2+ plays the role of oxidant in the mild oxidation system. The single-crystal X-ray analysis of 2 is similar to 1, and the PXRD pattern showed no obvious structural changes (as shown in Figure 3). According to the final crystallographic parameters, a slight increase of axis was observed in 2, along with an increase of more than 14 Å3 of cell volume. In the IR spectra of 1 and 2 (Figure S12), the peak of 3300m changed,

Figure 3. Transformation of 1 → 2.

which corresponds to the formation of hydroxide. Because the difference of valence may cause changes in magnetic behavior of MOFs material,51,52 the variable temperature magnetic susceptibilities of 1 and 2 were collected at a 1 kOe dc field over 300−2 K. In Figure S5, for 1, the χMT value at 300 K (15.24 cm3 K mol−1) is higher to three spin-only FeII ions, suggesting strong spin−orbit coupling interaction. Upon cooling, the χMT value decreases slowly to 9.60 cm3 K mol−1 C

DOI: 10.1021/acs.inorgchem.8b01421 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. 57Fe Mössbauer spectra and XPS of 1 and 2.

at 30 K, and then the value declines rapidly. This phenomenon may be due to zero-field splitting.31,53 For 2, the χMT is 10.24 cm3 K mol−1 at 300 K, which decreases more quickly than that of 1. The minimum is 2.07 cm3 K mol−1 at 2 K. We infer that the difference in magnetic behavior is mainly attributable to a reduction of the number of electrons, which weakened the antiferromagnetic coupling effect. 57 Fe Mössbauer spectroscopy show that Fe atoms of 1 are divalent, while the valence of Fe changed to III in 2. In Figure 4, for 1, a quadrupole doublet with the isomer shift (δ) and quadrupole splitting (ΔEQ) characteristic of FeII was observed (Table S1). There are only two doublets at 300 K, and doublet was characterized by the quadrupole splitting with ΔEQ = 3.27 and 2.95 mm s−1 and isomer shift δ = 1.11 and 1.07 mm s−1. These values are corresponding to the HS state of FeII (S = 2).31,54 As for 2, the characteristic of ΔEQ = 0.89 mm s−1 and δ = 0.39 mm s−1 is assigned to FeIII. In addition, we also conducted XPS and solid diffuse reflectance spectrometry on 2, and the results both showed that the existence of FeIII was credible (Figure S4). Catalytic Performance of 1 and 2. Before the catalytic experiment, the solvent-resistance properties of 1 and 2 were examined by suspending samples in different organic solvents and pH for 24 h. After this process, we first looked at the appearance of the crystal through optical microscope, and the color and shape of the crystal have not changed significantly. Then, the crystal is filtered and verified for structural integrity by PXRD spectrum. As shown in Figures S7 and S8, the primary frameworks were intact throughout the process. In addition, the thermal analysis were carried out (Figure S6), and the framework of 1 and 2 begins to collapse at 300 °C. The rest of the weight is consistent with the theoretical formation of Fe2O3. As a result, 1 and 2 have good chemical and thermal stability; it cannot be broken down in the process of catalysis. To better verify the catalytic performance of MOF1 and 2 for acylation-Nazarov cyclization reaction, we chose 1,3dimethoxybenzene and 4a as reaction substrates to compare their activities under different reaction conditions (as shown in Table 2). According to the analysis of yield under different conditions, the most suitable reaction solvent is DCE, and the reaction time is 5 hours at reflux. Noticeable, TFAA is indispensable in this system. As shown in Table 1, 1 showed a high specificity of acylation products selectivity (Table 1, entry 8). On the contrary, 2 could effectively catalyze the generation of cyclization productsin yields of 73% (Table 1, entry 9), the

Table 2. Screening of the Reaction Conditions entry 1 2 3 4 5 4 5 6 7 8 9 a

catalysis TFAA 1 2 2 FeCl2 FeCl3 ZnCl2 no 1 2 1 2

yes yes yes yes yes yes yes no no yes yes

solvent

T (°C)

yield (%) of 5ab

yield (%) of 6ab

DCM DCM DCE DCE DCE DCE DCE DCE DCE DCE DCE

60 60 25 90 90 90 90 90 90 90 90

n.o.a 11 n.o.a n.o.a 53 27 5< n.o.a n.o.a n.o.a 73

56 30 38 60 39 61 37 n.o.a n.o.a 90 21

Not observed. bYields are calculated based on 3.

whole tandem reaction was supported. For comparison, the homogeneous catalysts ZnCl2 and FeCl3 as common Lewis acid catalysts were used for this reaction under the same reaction conditions. The target product 5a only were obtained in yields of 27% and 53%, respectively. Thus, heterogeneous catalyst 2 clearly showed superior activity in implementing the tandem acylation-Nazarov cyclization reaction with high selectivity. To further investigate the catalytic performance of 1 and 2 in Nazarov cyclization, other α,β-unsaturated carboxylic acids (4b−d) were introduced as substrates, and the tandem reaction conducted successfully and supported the cyclization of 1,3-dimethoxybenzene. As shown in Table 3, the same as expected, heterogeneous catalyst 2 could effectively catalyzed Nazarov cyclization and 5b−d is produced. The target product of cyclopentenone[b]benzene were got in yields ranging from 52% to 73%. The yields decrease in the order 5a (73%) > 5b (60%) > 5c (56%) > 5d (50%), corresponding to the increase of substrate sizes 4a (3.83 Å) < 4b (3.92 Å) < 4c (4.51 Å) < 4d (4.91 Å). It indicated that the substrate sizes play vital roles in the heterogeneous catalyzed reaction. In addition, 1 shows high selectivity to acylated products with high yield (6a (90%) > 6b (85%) > 6c (81%) > 6d (78%)). We deduced that the high catalytic activities of 1 and 2 were mainly came from their high density Lewis acid sites and the fixed framework channels. The characteristics of the structure have the following advantages: (i) Structural analysis shows that there is a pathway (5.0 × 16.4 Å2) in the framework, such tunnels are favorable for the entry of the reaction substrate. (ii) The D

DOI: 10.1021/acs.inorgchem.8b01421 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Table 3. Catalytic Efficiency for Different Substrates

a

Not observed. bYields are calculated based on 3 and the average of two runs.

Fe center in MOFs catalyst. The coordination interaction with MOF catalyst plays the effect of activation for carbonyl group from α,β-unsaturated acid, and makes carbon atoms more electrophilic. Then, electrophilic carbon atoms is attacked by 1,3-dimethoxybenzene to provide intermediate II. The lost of trifluoroacetic acid unit let III been generated. Subsequently, the electrons rearrangement of III lead to hydrogen atom leaves and give the acylation products IV. Furthermore, the Febased MOFs catalyst activated the unsaturated olefinic bond continuously and generated a pentadienyl cation, followed by 4π conrotatory electrocyclization, generating the benzocyclopentanone V, and Fe-based MOFs was rereleased for participating in the next catalytic cycle. Noteworthy, the frameworks of the two catalysts are similar, but the catalytic effect showed a great difference. The differences can be attributed to that the metal center (Fe2+) in MOF 1 is not strong enough as a lewis acid for activating the unsaturated olefinic bond. The highly charged Fe3+ center in 2 has a strong ability to receive electronically, which can promote 4π conrotatory electrocyclization and make acylation IV can easily be converted.55 In order to better investigate the potential application value of catalysts, the recycling experiment was carried out on behalf of 4a to determine the reusability of the catalyst. Interestingly, the catalyst is easy to be separated by magnetic attraction after the reaction is completed. Compared with centrifugation and filtration methods in previous reports, this characteristic is beneficial for the recovery of the catalyst. The recycling experiment results showed that the catalytic efficiency did not display significant decline after five runs (Figure 6 and Figure S10). Furthermore, we tested the recovered catalyst by PXRD, the spectrum was very similar to that of simulation by single crystal data. The consistency of PXRD patterns before and after the catalysis indicated that the framework was intact and not collapsed (Figure S9), which proved the stability and reusability of the two catalysts after at least five runs, and 1 and

trinuclear [Fe3(CO2)2] units exhibit excellent catalytic activity for the cyclization reaction, because there is coupling synergy between iron ions. At the same time, the trinuclear provides a platform for improving the density of lewis acid sites. (iii) There are free water molecules around the metal center, which are a good leaving group. These groups are conducive to exposing the metal catalytic site to the reaction substrate during the reaction process. Therefore, Fe-based MOFs 1 and 2 can be good candidates as heterogeneous catalyst. In order to prove more accurately that the reaction occurs within the crystals, the control experiment was compared with large crystals of 2 and grinding 2. The similar yield was presented, which eliminates the catalytic activity only come from the surface metal sites. In addition, solid catalysts were removed after 1 h and the conversion has not been improved. A tentative catalytic mechanism of mechanism Fe-based MOFs for Nazarov cyclization is proposed in Figure 5. First, a substitution reaction took place between TFAA and α,βunsaturated acid to give mixed anhydride, which formed I with

Figure 5. Suggested mechanism. E

DOI: 10.1021/acs.inorgchem.8b01421 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 6. Reusability test for catalytic performance of 2.

Notes

2 can be a good heterogeneous catalyst candidate for organic synthesis.

The authors declare no competing financial interest.





CONCLUSIONS A rare example of PSO was demonstrated in a 3D FeII-based MOF(1), and the oxidized MOF (2) retains a good single crystallinity. The progress of the reaction oxidation was proved by single crystal XRD, PXRD, XPS, 57Fe Mössbauer spectroscopy, and UV−vis. Compounds 1 and 2 were used as catalysts for the tandem acylation-Nazarov cyclization, and they show the different catalytic selectivities. At the same time, 1 and 2 show the good stability, high catalytic activity, and recyclability for reuse. These results highlight the superiority of the MOFs materials as heterogeneous catalysts and underscore the feasibility of PSO as an efficient strategy to develop hardsynthesized MOFs materials.



(1) Madrahimov, S.; Gallagher, J.; Zhang, G.; Meinhart, Z.; Miller, J.; Farha, O.; Hupp, J.; Nguyen, S. Gas-Phase Dimerization of Ethylene under Mild Conditions Catalyzed by MOF Materials Containing (bpy)NiII Complexes. ACS Catal. 2015, 5, 6713−6718. (2) Yang, Q.; Xu, Q.; Jiang, H. Metal−organic frameworks meet metal nanoparticles: synergistic effect for enhanced catalysis. Chem. Soc. Rev. 2017, 46, 4774−4808. (3) Shi, Z. Q.; Guo, Z. J.; Zheng, H. G. Two luminescent Zn (II) metal-organic frameworks for exceptionally selective detection of picric acid explosives. Chem. Commun. 2015, 51, 8300−8303. (4) Chen, D. M.; Zhang, N. N.; Liu, C. S.; Du, M. Dual-Emitting Dye@MOF Composite as a Self-Calibrating Sensor for 2,4,6Trinitrophenol. ACS Appl. Mater. Interfaces 2017, 9, 24671−24677. (5) Osadchii, D.; Olivos-Suarez, A.; Li, G.; Nasalevich, M.; Dugulan, I.; Crespo, P.; Hensen, E.; Veber, S.; Fedin, M.; Sankar, G.; Pidko, E.; Gascon, J. Isolated Fe Sites in Metal Organic Frameworks Catalyze the Direct Conversion of Methane to Methanol. ACS Catal. 2018, 8, 5542−5548. (6) Lu, Z. Z.; Zhang, R.; Li, Y. Z.; Zheng, H. G. Solvatochromic behavior of a nanotubular metal−organic framework for sensing small molecules. J. Am. Chem. Soc. 2011, 133, 4172−4174. (7) Han, Y. F.; Jia, W. G.; Yu, W. B.; Jin, G. X. Stepwise formation of organometallic macrocycles, prisms and boxes from Ir, Rh and Rubased half-sandwich units. Chem. Soc. Rev. 2009, 38, 3419−3434. (8) Yao, M. S.; Tang, W. X.; Wang, G. E.; Nath, B.; Xu, G. MOF Thin Film-Coated Metal Oxide Nanowire Array: Significantly Improved Chemiresistor Sensor Performance. Adv. Mater. 2016, 28, 5229−5234. (9) Stavila, V.; Parthasarathi, R.; Davis, R.; Gabaly, F.; Sale, K.; Singh, S.; Allendorf, M. MOF-Based Catalysts for Selective Hydrogenolysis of Carbon−Oxygen Ether Bonds. ACS Catal. 2016, 6, 55− 59. (10) Flaig, R. W.; Popp, T. M.; Fracaroli, A. M.; Reimer, J. A.; Yaghi, O. M. The Chemistry of CO2 Capture in an Amine-Functionalized Metal− Organic Framework under Dry and Humid Conditions. J. Am. Chem. Soc. 2017, 139, 12125−12128. (11) Zhang, W. H.; Ren, Z. G.; Lang, J. P. Rational construction of functional molybdenum (tungsten)−copper−sulfur coordination oligomers and polymers from preformed cluster precursors. Chem. Soc. Rev. 2016, 45, 4995−5019. (12) Du, M.; Wang, L.; Liu, C. S. Divergent kinetic and thermodynamic hydration of a porous Cu (II) coordination polymer with exclusive CO2 sorption selectivity. J. Am. Chem. Soc. 2014, 136, 10906−10909. (13) Yu, C. X.; Shao, Z. C.; Hou, H. W. A functionalized metalorganic framework decorated with O− groups showing excellent

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b01421. Additional experimental details and supporting figures. Crystallographic data for 1 and 2 (PDF). Accession Codes

CCDC 1837090−1837091 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.



REFERENCES

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Jie Wu: 0000-0001-7746-6640 Hongwei Hou: 0000-0003-4762-0920 Funding

This work was financially supported by the National Natural Science Foundation (21371155 and 21671174) and the Natural Science Foundation of Henan province (182300410008). F

DOI: 10.1021/acs.inorgchem.8b01421 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry performance for lead(II) removal from aqueous solution. Chem. Sci. 2017, 8, 7611−7619. (14) Yao, M. S.; Lv, X. J.; Fu, Z. H.; Li, W. H.; Deng, W. H.; Wu, G. D.; Xu, G. Layer-by-Layer Assembled Conductive Metal−Organic Framework Nanofilms for Room-Temperature Chemiresistive Sensing. Angew. Chem., Int. Ed. 2017, 56, 16510−16514. (15) Qin, J.; Ma, B.; Liu, X. F.; Lu, H. L.; Dong, X. Y.; Zang, S. Q.; Hou, H. Aqueous- and vapor-phase detection of nitroaromatic explosives by a water-stable fluorescent microporous MOF directed by an ionic liquid. J. Mater. Chem. A 2015, 3, 12690−12697. (16) Du, Y. R.; Li, X. Q.; Lv, X. J.; Jia, Qi. Highly Sensitive and Selective Sensing of Free Bilirubin Using Metal−Organic Frameworks-Based Energy Transfer Process. ACS Appl. Mater. Interfaces 2017, 9, 30925−30932. (17) Rubin, H. N.; Neufeld, B. H.; Reynolds, M. M. SurfaceAnchored Metal−Organic Framework−Cotton Material for Tunable Antibacterial Copper Delivery. ACS Appl. Mater. Interfaces 2018, 10, 15189−15199. (18) Luis, G. T.; Sabina, R. H.; Inhar, I.; Daniel, M. Spray Drying for Making Covalent Chemistry: Postsynthetic Modification of Metal− Organic Frameworks. J. Am. Chem. Soc. 2017, 139, 897−903. (19) Wu, D. Q.; Shao, D.; Wei, X. Q.; Kempe, D.; Zhang, Y. Z.; Dunbar, K. R.; Wang, X. Y. Reversible On−Off Switching of a SingleMolecule Magnet via a Crystal-to-Crystal Chemical Transformation. J. Am. Chem. Soc. 2017, 139, 11714−11717. (20) Liu, D.; Lang, J. P.; Abrahams, B. F. Highly Efficient Separation of a Solid Mixture of Naphthalene and Anthracene by a Reusable Porous Metal−Organic Framework through a Single-Crystal-toSingle-Crystal Transformation. J. Am. Chem. Soc. 2011, 133, 11042−11045. (21) Chen, D. M.; Zhang, N. N.; Liu, C. S.; Du, M. Pore modulation of metal−organic frameworks towards enhanced hydrothermal stability and acetylene uptake via incorporation of different functional brackets. J. Mater. Chem. A 2017, 5, 4861−4867. (22) Zhang, W. Y.; Lin, Y. J.; Han, Y. F.; Jin, G. X. Facile Separation of Regioisomeric Compounds by a Heteronuclear Organometallic Capsule. J. Am. Chem. Soc. 2016, 138, 10700−10707. (23) Karagiaridi, O.; Bury, W.; Tylianakis, E.; Sarjeant, A. A.; Hupp, J. T.; Farha, O. K. Opening metal−organic frameworks vol. 2: inserting longer pillars into pillared-paddlewheel structures through solvent-assisted linker exchange. Chem. Mater. 2013, 25, 3499−3503. (24) Bury, W.; Jimenez, D. F.; Lalonde, M. B.; Farha, O. K.; Hupp, J. T. Control over Catenation in Pillared Paddlewheel Metal−Organic Framework Materials via Solvent-Assisted Linker Exchange. Chem. Mater. 2013, 25, 739−744. (25) Marshall, R. J.; Griffin, S. L.; Wilson, C.; Forgan, R. S. SingleCrystal to Single-Crystal Mechanical Contraction of Metal−Organic Frameworks through Stereoselective Postsynthetic Bromination. J. Am. Chem. Soc. 2015, 137, 9527−9530. (26) Zhao, J. A.; Mi, L. W.; Hu, J. Y.; Hou, H. W. Cation exchange induced tunable properties of a nanoporous octanuclear Cu(II) wheel with double-helical structure. J. Am. Chem. Soc. 2008, 130, 15222− 15223. (27) Song, X.; Kim, T.; Kim, H.; Kim, D.; Lah, M. S. Post-Synthetic Modifications of Framework Metal Ions in Isostructural MetalOrganic Frameworks: Core-Shell Heterostructures via Selective Transmetalations. Chem. Mater. 2012, 24, 3065−3073. (28) Fu, J. H.; Li, H. J.; Mu, Y. J.; Hou, H. W. Reversible single crystal to single crystal transformation with anion exchange-induced weak Cu2+···I interactions and modification of the structures and properties of MOFs. Chem. Commun. 2011, 47, 5271−5273. (29) Wang, H. R.; Meng, W.; Wu, J.; Ding, J.; Hou, H. W. Crystalline central-metal transformation in metal-organic frameworks. Coord. Chem. Rev. 2016, 307, 130−146. (30) Han, Y.; Xu, H.; Li, H. J.; Hou, H. W.; Batten, S. R. Temperature-Dependent Capture of Water Molecules by SaddleShaped Hexanuclear Carboxylate Cycloclusters in a (3,18)-Connected Metal−Organic Framework. Chem. - Eur. J. 2012, 18, 13954−13958.

(31) Xu, Z. Q.; Meng, W.; Li, H. J.; Hou, H. W.; Fan, Y. T. Guest Molecule Release Triggers Changes in the Catalytic and Magnetic Properties of a FeII-Based 3D Metal−Organic Framework. Inorg. Chem. 2014, 53, 3260−3262. (32) Sun, J.; Dai, F.; Yuan, W.; Bi, W.; Zhao, X.; Sun, W.; Sun, D. Dimerization of a metal complex through thermally induced singlecrystal-to-single-crystal transformation or mechanochemical reaction. Angew. Chem., Int. Ed. 2011, 50, 7061−7064. (33) Ge, J. Y.; Wang, J.; Cheng, J.; Wang, P.; Ma, J.; Liu, Q.; Dong, Y. Oxygen and methanol mediated irreversible coordination polymer structural transformation from a 3D Cu(I)-framework to a 1D Cu(II)-chain. Chem. Commun. 2014, 50, 4434−4437. (34) Huang, C.; Wu, J.; Song, C. J.; Qiao, Y.; Hou, H. W.; Chang, J. B.; Fan, Y. T. Reversible conversion of valence-tautomeric copper metal−organic frameworks dependent single-crystal-to-single-crystal oxidation/reduction: a redox-switchable catalyst for C−H bonds activation reaction. Chem. Commun. 2015, 51, 10353−10356. (35) Huang, Y.; Mu, B.; Schoenecker, P. M.; Carsom, C. G.; Karra, J. R.; Cai, Y.; Walton, K. S. A Porous Flexible Homochiral SrSi2 Array of Single-Stranded Helical Nanotubes Exhibiting Single-Crystal-toSingle-Crystal Oxidation Transformation. Angew. Chem., Int. Ed. 2011, 50, 436−440. (36) He, Q.; Li, X.; Liu, Y.; Yu, Z.; Wang, W.; Su, C. Copper(I) Cuboctahedral Coordination Cages: Host−Guest Dependent Redox Activit. Angew. Chem., Int. Ed. 2009, 48, 6156−6159. (37) Shao, Z. C.; Huang, C.; Han, X.; Hou, H. W.; Fan, Y. T. effect of metal ions on photocatalytic performance based on an isostructural framework. Dalton Trans. 2015, 44, 12832−12838. (38) Giang, H.; Nguyen, T.; Schweitzer, N.; Chang, C.; Drake, T.; Farha, O.; Hupp, J.; Nguyen, S. Vanadium-Node-Functionalized UiO66: A Thermally Stable MOF-Supported Catalyst for the Gas-Phase Oxidative Dehydrogenation of Cyclohexene. ACS Catal. 2014, 4, 2496−2500. (39) Zhu, C.; Xia, Q.; Chen, X.; Liu, Y.; Du, X.; Cui, Y. Chiral Metal-Organic Framework as a Platform for Cooperative Catalysis in Asymmetric Cyanosilylation of Aldehydes. ACS Catal. 2016, 6, 7590− 7596. (40) Majewski, M. B.; Peters, A. W.; Wasielewski, M. R.; Hupp, J. T.; Farha, O. K. Metal−Organic Frameworks as Platform Materials for Solar Fuels Catalysis. ACS Energy Lett. 2018, 3, 598−611. (41) Li, Z. Y.; Peters, A. W.; Platero-Prats, A. E.; Liu, J.; Kung, C.; Hupp, J. T.; Farha, O. K. Fine-Tuning the Activity of Metal−Organic Framework-Supported Cobalt Catalysts for the Oxidative Dehydrogenation of Propane. J. Am. Chem. Soc. 2017, 139, 15251−15258. (42) Shen, K.; Chen, X.; Chen, J.; Li, Y. Development of MOFDerived Carbon-Based Nanomaterials for Efficient Catalysis. ACS Catal. 2016, 6, 5887−5903. (43) Huang, C.; Ding, R.; Song, C.; Wu, J.; Hou, H.; Fan, Y. Template-Induced Diverse Metal−Organic Materials as Catalysts for the Tandem Acylation−Nazarov Cyclization. Chem. - Eur. J. 2014, 20, 16156−16163. (44) Huang, C.; Han, X.; Shao, Z.; Gao, K.; Liu, M.; Wang, Y.; Wu, J.; Hou, H. Solvent-Induced Assembly of Sliver Coordination Polymers (CPs) as Cooperative Catalysts for Synthesizing of Cyclopentenone[b]pyrroles Frameworks. Inorg. Chem. 2017, 56, 4874. (45) SAINT, Program for Data Extraction and Reduction; Bruker AXS, Inc.: Madison, WI, 2001. (46) Sheldrick, G. M. SADABS, Program for Empirical Adsorption Correction of Area Detector Data; University of Göttingen: Germany, 2003. (47) Sheldrick, G. M. A short history of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, A64, 112−122. (48) Wang, K.; Feng, D.; Liu, T.; Su, J.; Yuan, S. A Series of Highly Stable Mesoporous Metalloporphyrin Fe-MOFs. J. Am. Chem. Soc. 2014, 136, 13983−13986. (49) Liu, W.; Li, J.; Ni, Z.; Liu, J.; Tong, M. Incomplete Spin Crossover versus Antiferromagnetic Behavior Exhibited in ThreeG

DOI: 10.1021/acs.inorgchem.8b01421 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Dimensional Porous Fe(II)-Bis(tetrazolate) Frameworks. Cryst. Growth Des. 2012, 12, 1482−1488. (50) Qin, J.; Jia, Y.; Li, H.; Zhao, B.; Wu, D.; Zang, S.; Hou, H.; Fan, Y. Conversion from a Heterochiral 2 + 2 Coaxially Nested DoubleHelical Column to a Cationic Spiral Staircase Stimulated by an Ionic Liquid Anion. Inorg. Chem. 2014, 53, 685−687. (51) Goswami, S.; Adhikary, A.; Jena, H.; Biswas, S.; Konar, S. A 3D Iron(II)-Based MOF with Squashed Cuboctahedral Nanoscopic Cages Showing Spin-Canted Long-Range Antiferromagnetic Ordering. Inorg. Chem. 2013, 52, 12064−12069. (52) Han, Y.; Chilton, N. F.; Li, M.; Hou, H. W.; Batten, S. R. PostSynthetic Monovalent Central-Metal Exchange, Specific I2 Sensing, and Polymerization of a Catalytic [3 × 3] Grid of [CuII5CuI4L6]·(I)2· 13H2O. Chem. - Eur. J. 2013, 19, 6321−6328. (53) Zeng, M. H.; Zhou, Y. L.; Wu, M. C.; Sun, H. L.; Du, M. A unique cobalt(II)-based molecular magnet constructed of hydroxyl/ carboxylate bridges with a 3D pillared-layer motif. Inorg. Chem. 2010, 49, 6436. (54) Garcia, Y.; Bravic, G.; Gieck, C.; Chasseau, D.; Tremel, W.; Gü tlich, P. Crystal Structure, Magnetic Properties, and 57Fe Mo1ssbauer Spectroscopy of the Two-Dimensional Coordination Polymers [M(1,2-bis(1,2,4-triazol-4-yl)ethane) 2(NCS)2] (MII) Fe, Co). Inorg. Chem. 2005, 44, 9723−9730. (55) Liu, M.; Gao, K.; Fan, Y.; Guo, X.; Wu, J. Co-Cluster-Based Metal−Organic Frameworks as Selective Catalysts for Benzene Tandem Acylation−Nazarov Cyclization to Benzocyclopentanone. Chem. - Eur. J. 2018, 24, 1416−1424.

H

DOI: 10.1021/acs.inorgchem.8b01421 Inorg. Chem. XXXX, XXX, XXX−XXX