A Versatile Microporous Zinc(II) Metal–Organic ... - ACS Publications

Apr 11, 2018 - Tianjin Normal University, Tianjin 300387, People,s Republic of ... concerns in daily life.1 Thus far, gas adsorption separation ... so...
0 downloads 0 Views 3MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

A Versatile Microporous Zinc(II) Metal−Organic Framework for Selective Gas Adsorption, Cooperative Catalysis, and Luminescent Sensing Hongming He,*,† De-Yu Zhang,† Feng Guo,§ and Fuxing Sun*,‡ †

Key Laboratory of Inorganic−Organic Hybrid Functional Material Chemistry of the Ministry of Education, College of Chemistry, Tianjin Normal University, Tianjin 300387, People’s Republic of China ‡ State Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, People’s Republic of China § School of Chemistry and Chemical Engineering, Yangtze Normal University, Chongqing 408100, People’s Republic of China S Supporting Information *

ABSTRACT: A versatile microporous zinc(II) MOF (1) with plentiful Lewis basic sites and open metal sites synchronously has been synthesized successfully. The resultant microporous material displays excellent selectivity adsorption for small gases. The IAST selectivity values for CO2/CH4, C2H6/ CH4, CO2/N2, and C3H8/CH4 are 7, 21, 38, and 125 at 101 kPa and 298 K, respectively. 1 exhibits remarkably heterogeneous catalytic performance for acid−base one-pot reactions. Furthermore, 1 exhibits recyclable selective detectability for TNP. The results illustrate that 1 is a versatile material for selective gas adsorption, cooperative catalysis, and luminescent sensing.



INTRODUCTION The primary component of greenhouse gas is carbon dioxide (CO2) in the atmosphere. CO2 has attracted a great deal of research interest from chemists and materials scientists recently, as it raises a number of energy crises and environmental concerns in daily life.1 Thus far, gas adsorption separation technology is the most effective method of dealing with CO2.2 Natural gas, as a major fossil fuel, can be considered to be a preferable alternative energy fuel, and it contains abundant methane (CH4) and varying amounts of C2H6 and C3H8. Notably, CH4 is not only a prevalent and clean fuel but also an available C1 source in industry for multiple C1 and C2 chemicals.3 A great deal of effort has been focused on exploring novel porous materials to store CO2 and purify CH4 from natural gas. Currently, a gradually increasing interest has been focused on the development of heterogeneous catalysis for onepot and sequential reactions, which is mainly ascribed to their high efficiency, energy efficiency, and reduction of waste.4 Significantly, considerable attention has been given to acid− base one-pot catalysis reactions, which have been widely used in common organic synthesis. Due to the acid−base neutralization reaction between the acid and base sites, it is a remarkable challenge to prepare efficiently site isolating acid−base one-pot heterogeneous catalysts. Meanwhile, nitroaromatic compounds (NACs) are the most common explosives and contaminants and are broadly used in the military, agriculture, and industry. It is found that NACs trigger a series of environmental and safety © XXXX American Chemical Society

problems all over the world. Currently, a potential method to efficiently detect NACs for homeland security and environmental protection vis ia luminescent materials. Hence, it is a crucial challenge to develop and explore multifunctional porous materials for the above goals. Metal−organic frameworks (namely MOFs)5 have attracted tremendous interest in the past 12 years due to their diverse architectures and applications in heterogeneous catalysis,6 gas sorption,7 sensing,8 enzyme immobilization,9 and optical apparatus.10 Because of the relationship between structures and characteristics in MOFs, MOFs are able to be designed and tuned through rationalization of bridging linkers and metal units. Therefore, we speculate that if the frameworks possess both Lewis basic sites (LBSs) and unsaturated open metal sites (OMSs) and have luminescent properties, they can form multifunctional materials for gas separation, catalysis for acid− base sequential reactions, and luminescent sensing for NACs. Recently, 1,3,5-triazine functional groups have been imported successfully into MOFs and can serve as efficient LBSs for many applications.11 In view of the aforementioned discussion of the documented studies, we selected a semirigid organic ligand with abundant N-rich functional groups, 2,4-bis(3,5dicarboxyphenylamino)-6-ol triazine (H4BDPO).11d Under solvothermal conditions, the organic ligand can coordinate Received: April 11, 2018

A

DOI: 10.1021/acs.inorgchem.8b00938 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry with ZnII to form a rare three-dimensional (3D) ZnII paddlewheel-based microporous MOF with the formula [Zn24(BDPO)12(DMF)12]·6DMF·52H2O (1; DMF = N,Ndimethylformamide). Thanks to ZnII having a closed-shell d10 electron configurations without any luminescence quenching effects, 1 has luminescent properties as for the free organic ligand H4BDPO. As we expected, the obtained MOF illustrated remarkably selective adsorption (CO2 over N2 and CH4 and CH4 over C2H6 and C3H8), had excellent performance as a heterogeneous catalyst for acid−base sequential reactions, and exhibited highly selective sensing for TNP (as shown in Scheme 1).

atoms were directly located and anisotropically refined. Due to the fact that the solvent molecules were highly disordered, the SQUEEZE method of PLATON was used to delete them.14 The solvents could be determined from TGA and elemental analysis results. The crystallographic results are given in Table 1, and selected bond angles and lengths are summarized in Table S2.

Table 1. Crystallographic Data for 1 empirical formula formula wt cryst syst space group a = b = c (Å) α = β = γ (deg) V (Å3) Z ρcalcd (g cm−3) μ (mm−1) Nref F(000) R(int) goodness of fit on F2 R1, wR2 (I > 2σ(I)) R1, wR2 (all data)

Scheme 1. Versatile Microporous MOF for Gas Adsorption, Cooperative Catalysis, and Luminescent Detection

C282H338O178N78Zn24 9222 cubic Im3̅m 27.9409(14) 90.00 21813.3(19) 24 1.109 1.361 2014 7248.0 0.0345 1.087 0.0879, 0.2583 0.0895, 0.2610





RESULTS AND DISCUSSION Structure Description. 2,4-Bis(3,5-dicarboxyphenylamino)-6-ol triazine (H4BDPO) and Zn(NO3)2·6H2O were both dissolved in DMF and H2O at 120 °C for 3 days, generating colorless block-shaped crystals. The single-crystal data illustrated that 1 belongs to the cubic space group Im3̅m space group (Figure S1 in the Supporting Information). Figure 1 EXPERIMENTAL SECTION

Materials and Methods. Chemicals and reagents were purchased commercially and used without any purification. Room-temperature powder X-ray diffraction (PXRD) data were collected on a D-Max 2550 diffractometer with Cu Kα radiation at 40 kV and 200 mA with the 2θ values from 4 to 40°. Thermogravimetric analyses (TGA) were carried out with a TGA Q500 analyzer under an N2 flow to 800 °C. A Vario MICRO elemental analyzer was used to obtain elemental analyses of C, H, and N. Fourier-transform infrared spectra (FT-IR) were directly recorded on a nano-FTIR instrument from 4000 to 400 cm−1. All gas measurements were measured with a Micrometrics ASAP 2020 apparatus. UV−vis absorption spectra were obtained with a Mapada V-1200 spectrophotometer. All luminescence data were obtained on a Cary Eclipse fluorescence spectrophotometer at room temperature. The gas chromatographic results were directly obtained with an Agilent 6890 instrument. Synthesis of [Zn24(BDPO)12(DMF)12]·6DMF·52H2O (1). Zn(NO3)2·6H2O (0.030 g, 0.0001 mol) and H4BDPO (0.020 g, 0.000044 mol) were both dissolved in DMF (5 mL), H2O (0.75 mL), and HNO3 aqueous solution (2 M, 0.2 mL) and placed into a 18 mL autoclave and heated at 120 °C for 3 days. Crystals were formed and washed with fresh DMF. The yield was 76% based on H4BDPO. Anal. Calcd for C282H338O178N78Zn24: C, 36.89; H, 3.72; N, 11.98. Found: C, 36.70; H, 3.67; N, 11.84. Representational FT-IR data (KBr pellet, cm−1): 3339 (br), 2931 (s), 2867 (s), 1661 (s), 1595 (s), 1556 (s), 1531 (s), 1375 (s), 1254 (s), 1085 (s), 896 (s), 806 (s), 775 (s), 734 (s), 667 (s), 639 (s), 585 (s), 489 (s). Single-Crystal X-ray Crystallography. Room-temperature single-crystal X-ray diffraction measurements for 1 were collected on APEXII Quazar CCD instrument with Cu Kα radiation. The beam absorption effects were corrected using the SADABS program.12 The final structure of 1 was solved successfully by direct methods and refined on F2 values by the SHELXTL program.13 Non-hydrogen

Figure 1. Illustration for the construction of 1.

shows that the H4BDPO ligand contains two terminal isophthalate moieties, which can be assembled with ZnII paddle-wheel motifs to fabricate the typical MOP-1 structure. Fascinatingly, one axial site of the paddle-wheel motif coordinates to a nitrogen atom from the ligand to increase the net connectivity to 5. 1 has three different microporous cages in the overall structure. The terminal isophthalate moieties can be assembled with classical paddle-wheel motifs to form the typical MOP-1 about 14.9 Å. The medium-sized cages (12.1 Å) are built from eight units and four linkers. Other small cages (8.9 Å) are formed by six units and six linkers. B

DOI: 10.1021/acs.inorgchem.8b00938 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

12.7 Å were found by the DFT method, which agrees with the single crystal results (Figure 2, insert). The activated crystals were further measured for some other small gases. The low-pressure uptake for H2 is reversible, and the maximum adsorption amounts are about 1.76 wt % (198 cm3 g−1) and 1.23 wt % (138 cm3 g−1) at 77 and 87 K under 1 atm, respectively (Figure S8 in the Supporting Information), which are higher than those of PCN-6′ (1.37 wt %)16 and HKUST-1 (1.44 wt %)17 at 77 K. The isosteric heat (Qst) at zero coverage (6.57 kJ mol−1) was obtained from H2 adsorption isotherms at 77 and 87 K by the virial equation (Figure S10 in the Supporting Information). Furthermore, the adsorption amount of O2 at 77 K is about 408 cm3 g−1 (Figure S9 in the Supporting Information), which is similar to that of other highBET MOFs.18 The maximum uptakes of CO2 were about 109 cm3 g−1 (273 K) and 79 cm3 g−1 (298 K) under 1 atm (Figure 3a), which are higher than those of SUN-50 (106 cm3 g−1)19

These different cages are further connected by bridging ligands to generate a 3D structure with multiple pore systems. The resultant 3D network was symbolized as a rare 5-node uninodal topology with the point symbol [46·64] (Figure S2 in the Supporting Information).15 The free volume of 1 was calculated to be about 49.1% after elimination of solvent molecules by the PLATON/VOID method.14b PXRD, FT-IR, and Thermal Analysis. The PXRD data of 1 at room temperature illustrated that the obtained crystalline materials are pure (Figure S3 in the Supporting Information). As shown in Figure S4 in the Supporting Information, TGA curves of 1 under an N2 flow showed a slow weight loss of 24.47% from room temperature to 310 °C due to the solvent molecules in 1 (calculated 24.40%). The main skeleton framework started to decompose with an increase in temperature from 400 °C. The as-synthesized crystals were exchanged solvent molecules with MeOH for 3 days and CH2Cl2 for 3 days, respectively. Then samples were evacuated at 100 °C ofr about 10 h to obtain the activated samples for small gas adsorptions. The activated sample’s PXRD pattern showed that the pristine structure was retained well after the activation process (Figure S3 in the Supporting Information). Figure S5 in the Supporting Information shows that the characteristic peak (1704 cm−1) of −COOH only appeared in the free H4BDPO ligand, confirming that all of the carboxyl groups were fully deprotonated in 1. In addition, the typical peak (1645 cm−1) of CO of DMF molecules disappeared in the activated sample. The results showed that all DMF molecules were removed thoroughly, which was in agreement with the TGA results (Figure S4 in the Supporting Information). In addition, 1 displays excellent stability in different organic solvents and high heat stability still at 300 °C in air (Figures S6 and S7 in the Supporting Information). Gas Sorption Properties. N2 gas sorption measurements (77 K) of the activated sample were carried out to verify the permanent pore character. As indicated in Figure 2, N2 sorption isotherms featured a type I isotherm characteristic of the lowpressure region. The Brunauer−Emmett−Teller surface area and Langmuir surface area are 1021 and 1154 m2 g−1, respectively. To acquire an in-depth insight into the pore structure, the pore width distributions of about 7.3, 10.9, and

Figure 3. Gas sorption isotherms of CO2 (a), N2 (b), CH4 (c), C2H6 (d), and C3H8 (e) at 273 and 298 K, respectively, and (f) the corresponding Qst values of these small gases.

and ZIF-20 (69.8 cm3 g−1)20 at 273 K under 1 atm. The adsorption enthalpy of CO2 is 24.81 kJ mol−1 at 0 loading (Figure 3f and Figure S11 in the Supporting Information). As illustrated in Figure 3b,c, the adsorption and desorption isotherms of N2 and CH4 were obtained at 273 and 298 K. The maximum N2 uptakes at 1 atm are 10 and 6 cm3 g−1 and those of CH4 are 35 and 25 cm3 g−1, respectively. The Qst values of N2 and CH4 are 16.98 and 13.81 kJ mol−1 at zero loading, respectively (Figure 3f and Figures S12 and S13 in the Supporting Information). To study CO2 separation ability, the IAST method was used to predict the theoretical binary gas mixture separation ability from the experimental singlecomponent adsorption data. In addition, the dual-site Langmuir−Frendlich equations can fit the corresponding experimental results (Figure 4a). The predicted results were

Figure 2. N2 sorption isotherms at 77 K. Inset: the pore size distribution via the DFT method and the corresponding microporous framework in 1. C

DOI: 10.1021/acs.inorgchem.8b00938 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

dehyde by an acid-catalyzed process. The following reaction is Knoevenagel condensation by a base-catalyzed process. In a typical experiment, a mixture of dimethoxymethylbenzene (0.001 mol), malononitrile (0.0012 mmol), toluene (4 mL), and 1 (0.1 g) was placed in a glass reactor (20 mL) and stirred at 90 °C for 3 h. A GC equipped with a DB-5HT column was used to monitor the product yield. The reaction results are summarized in Table 2. 1 was an effectively cooperative catalyst Table 2. Catalytic Reaction Using Various Catalystsa entry 1 2 3 4 5 6 7 8

Figure 4. Gas adsorption isotherms of CO2, N2, CH4, C2H6, and C3H8 with DSLF fits (a, c) and binary gas adsorption selectivity by the IAST method (b, d).

catalyst 1 HY zeolite MgO HY zeolite + MgOb HClc TEAd HCl + TEAe no catalyst

conversn of 1 (%)f

yield of 2 (%)f

yield of 3 (%)f

≥99 91 7 83

trace 83 2 52

≥99 8 5 31

≥99 trace trace trace

97 trace trace trace

2 trace trace trace

a

Conditions unless specified otherwise: dimethoxymethylbenzene (0.001 mol), malononitrile (0.0012 mol), toleune (4 mL), catalyst (100 mg), 90 °C, 3 h. bHY zeolite (50 mg) and MgO (50 mg) were both used as catalysts. c0.0001 mol of HCl as a catalyst. d0.0001 mol of triethylamine (TEA) as a catalyst. eHCl (0.05 mmol) and TEA (0.05 mmol) were both used as catalysts. fThe yields were determined by GC with diphenyl as an internal standard.

evaluated from landfill binary gas mixtures (50/50, m/m) at 298 K (Figure 4b). The predicted selectivity values for CO2/ CH4 and CO2/N2 are 7 and 38 under 101 kPa, respectively. The results are distinctly better than those for many previously reported porous MOFs.21 Furthermore, C2H6 and C3H8 adsorption capacities were both studied at 273 and 298 K (Figure 3d,e). The maximum adsorbed uptakes of C2H6 are 80 and 72 cm3 g−1 and of C3H8 are 112 and 86 cm3 g−1, respectively. The Qst values at zero coverage are 29.60 and 54.37 kJ mol−1 for C2H6 and C3H8 via the virial method (Figure 3f and Figures S14 and S15 in the Supporting Information). The dual-site Langmuir−Frendlich (DSLF) equation was applied to match the adsorption data in Figure 4c. The predicted results for C2H6/CH4 and C3H8/CH4 could be calculated from a feed landfill gas (50/50, m/m) at room temperature under 1 atm (Figure 4d). The C2H6/CH4 adsorption selectivity is about 21, higher than those for numerous porous MOFs, including Mg-MOF-74 (11.5), NOTT-101 (12), and FIR-7a (14.6). 22 The C3 H8/CH 4 selectivity value is about 125, which is higher than those for UTSA-35a23 and JLU-Liu18.24 The results illustrated that 1 can be considered to be a promising material not only for separation and storage of CO2 but also for purification of CH4 from natural gas, mainly due to the abundant OMSs and LBSs in 1. Catalysis Properties. Only a few MOFs were applied to a tandem one-pot acid−base deacetalization−Knoevenagel condensation with high yields.25 Thanks to its large amount of OMSs and LBSs, 1 can be used as a bifunctional catalyst for a tandem one-pot acid−base deacetalization−Knoevenagel condensation (Scheme 2). The first reaction process is the deacetalization of dimethoxymethylbenzene to obtain benzal-

for this reaction with a yield of ≥99% after 3 h (Table 2, entry 1). Some control experiments were performed under similar conditions. The traditional HY zeolite (SiO2/Al2O3 = 30) was employed as the acid catalyst for the first reaction (Table 2, entry 2). MgO solid was used as a basic catalyst, which gave almost no products (Table 2, entry 3). When HY zeolite and MgO were both applied as catalysts, the conversion was significantly lower than that of 1 (Table 2, entry 4). In contrast, homogeneous catalysts were used for this reaction. HCl, as an acid catalyst, only effectively catalyzed the first step (Table 2, entry 5). Triethylamine (TEA) was applied as a basic catalyst, which generated almost no products even after 3 h (Table 2, entry 6). As we know, it is easy to neutralize acid and base catalysts in homogeneous systems. When a mixture of HCl and TEA was used as catalyst, the reaction hardly proceeded (Table 2, entry 7). In the absence of catalysts, the reaction also hardly occurred (Table 2, entry 8). The results illustrate that 1 is a bifunctional acid−base catalyst. In addition, ZnII ions cannot be detected in the reaction solution using inductively coupled plasma analysis, further confirming the heterogeneous nature of 1. 1 is easily separated by filtration and can be reused at least for five cycles with similar activity effects (Figure S16a in the Supporting Information). The frameworks of reused samples were confirmed by the PXRD patterns (Figure S16b in the Supporting Information). Furthermore, the luminescent spectra of the H4BDPO ligand and ground sample were both obtained. Figure 5a shows that the free ligand exhibits an emission at 382 nm under excitation at 337 nm. In comparison, the resultant sample shows an emission at 422 nm (λex 292 nm) (Figure 5b), indicating the luminescence of 1 from H4BDPO. To discuss the detectability of NACs, the emission peak of ground powders in DMF with 600 ppm of analytes was checked under 292 nm, considering that the crystal samples were insoluble and had highly

Scheme 2. Illustration of the Deacetalization−Knoevenagel Condensation

D

DOI: 10.1021/acs.inorgchem.8b00938 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

efficiency was confirmed by the Stern−Volmer equation: I0/I = 1 + KSV[Q]. I0 is the original luminescence intensity, I is the luminescence intensity after addition of analytes as quencher, [Q] denotes the molar concentration, and KSV can be determined as the Stern−Volmer constant. When I0/I vs [Q] is almost linear, KSV can be confirmed. The Stern−Volmer plots were almost linear at low concentrations (Figure S18 in the Supporting Information), but the plots significantly deviated from linear at higher concentrations (Figure S19 in the Supporting Information). This may be due to autoadsorption.24 The KSV values are 9.1 × 102 M−1 for NB, 1.0 × 103 M−1 for 1,3-DNB, 1.2 × 103 M−1 for 2,4-DNT, and 1.6 × 104 M−1 for TNP, respectively. Notably, the resulting sample is an efficient sensor for TNP (Table S3 in the Supporting Information).26 Thanks to the three nitro groups, 1 exhibited the most sensitive detection effect for TNP. With increasing −NO2, effective charge transfer can take place from electron rich 1 to more electron deficient analytes, leading to more efficient luminescence quenching.27 In addition, the UV−vis spectrum illustrated that the emission band of 1 has a coverage with the adsorption spectrum in DMF (Figure S20 in the Supporting Information). In fact, 1 is easily recollected from the dispersed mixture and washed with DMF. The detectability of 1 was very well retained even after five cycles (Figure S21a in the Supporting Information). The stability of the recollected crystals can be demonstrated by their PXRD patterns (Figure S21b in the Supporting Information).

Figure 5. Excitation and emission spectra of H4BDPO (a) and 1 (b). (c) Luminescence intensity of 1 with different analytes (600 ppm) in DMF. (d) Emission spectra of 1 (0.1 mg mL−1) by adding NB in DMF.

luminescence intensity in DMF (Figure 5c). Notably, 1 exhibits significant luminescent quenching behavior in nitrobenzene (NB). The emissive response of 1 can be measured by adding NB of various concentrations (Figure 5d). The luminescence intensity sharply decreased to 44% at 1 × 10−3 M and 97% at 1 × 10−2 M, respectively (Figure S17a in the Supporting Information). The luminescence quenching may be mainly due to electron transfer from 1 to the electron-deficient NB.8d,e The quenching behavior is mainly in connection with the electron-withdrawing −NO2 group. Hence, a series of NACs were used to explore the luminescence detectability of 1, including 1,3-dinitrobenzene (1,3-DNB), 2,4-dinitrotoluene (2,4-DNT) and 2,4,6-trinitrophenol (TNP). The alteration of luminescence can be measured at 292 nm by adding analytes (Figure 6a−c), and it can be almost quenched completely (Figure S17a in the Supporting Information). The quenching



CONCLUSION In summary, a microporous MOF (1) was synthesized successfully, which exhibits excellent luminescence properties and possesses many OMSs and LBSs. The resultant material displays high adsorption selectivity for CO2 over CH4 and N2 and for CH4 over C2H6 and C3H8. It also exhibits remarkably heterogeneous catalytic performance with high yields. Furthermore, the resultant sample can detect NACs via luminescence quenching effects. The results display that 1 is a multifunctional porous material for gas separation, cascade catalysis, and luminescence sensing.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b00938. FT-IR, TGA, gas sorption, PXRD, UV−vis spectra, and luminescent spectra (PDF) Accession Codes

CCDC 1829961 contains 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 data_ [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Authors

*E-mail for H.H.: [email protected]. *E-mail for F.S.: [email protected].

Figure 6. Emission spectra of 1 after addition of 1,3-DNB (a), 2,4DNT (b), and TNP (c) in DMF, respectively. (d) Relationships between luminescent intensities of 1 and the corresponding concentrations of NB, 1,3-DNB, 2,4-DNT, and TNP, respectively.

ORCID

Hongming He: 0000-0001-5535-8825 E

DOI: 10.1021/acs.inorgchem.8b00938 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Notes

Sun, F.; Bian, Z.; Gao, L.; Zhu, G. A Porous Metal−Organic Framework Formed by a V-Shaped Ligand and Zn(II) Ion with Highly Selective Sensing for Nitroaromatic Explosives. J. Mater. Chem. A 2015, 3, 16598−16603. (e) He, H.; Sun, F.; Jia, J.; Bian, Z.; Zhao, N.; Qiu, X.; Gao, L.; Zhu, G. Fluorescent Dodecapus in 3D Framework. Cryst. Growth Des. 2014, 14, 4258−4261. (9) (a) He, H.; Han, H.; Shi, H.; Tian, Y.; Sun, F.; Song, Y.; Li, Q.; Zhu, G. Construction of Thermophilic Lipase-Embedded Metal− Organic Frameworks via Biomimetic Mineralization: A Biocatalyst for Ester Hydrolysis and Kinetic Resolution. ACS Appl. Mater. Interfaces 2016, 8, 24517−24524. (b) Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Eubank, T. B. J. F.; Heurtaux, A.; Clayette, P.; Kreuz, C.; Chang, J.-S.; Hwang, Y. K.; Marsaud, V.; Bories, P.-N.; Cynober, L.; Gil, S.; Ferey, G.; Couvreur, P.; Gref, R. Porous Metal−Organic− Framework Nanoscale Carriers as a Potential Platform for Drug Delivery and Imaging. Nat. Mater. 2010, 9, 172−178. (10) (a) Li, Q.; Luo, J.; Wang, L.; Qi, C.; Yang, Y.; Zhang, X.; Qian, J. Two Cage-Based Zinc-Tetracarboxylate Frameworks with White-Light Emission. CrystEngComm 2017, 19, 214−217. (b) He, H.; Sun, F.; Borjigin, T.; Zhao, N.; Zhu, G. Tunable Colors and White-Light Emission Based on a Microporous Luminescent Zn(II)-MOF. Dalton Trans. 2014, 43, 3716−3721. (11) (a) Li, B.; Zhang, Z.; Li, Y.; Yao, K.; Zhu, Y.; Deng, Z.; Yang, F.; Zhou, X.; Li, G.; Wu, H.; Nijem, N.; Chabal, Y. J.; Lai, Z.; Han, Y.; Shi, Z.; Feng, S.; Li, J. Enhanced Binding Affinity, Remarkable Selectivity, and High Capacity of CO2 by Dual Functionalization of a rht−Type Metal−Organic Framework. Angew. Chem., Int. Ed. 2012, 51, 1412− 1415. (b) Luebke, R.; Eubank, J. F.; Cairns, A. J.; Belmabkhout, Y.; Wojtas, L.; Eddaoudi, M. The Unique rht−MOF Platform, Ideal for Pinpointing the Functionalization and CO2 Adsorption Relationship. Chem. Commun. 2012, 48, 1455−1457. (c) Liu, K.; Ma, D.; Li, B.; Li, Y.; Yao, K.; Zhang, Z.; Han, Y.; Shi, Z. High Storage Capacity and Separation Selectivity for C2 Hydrocarbons Over Methane in the Metal−Organic Framework Cu−TDPAT. J. Mater. Chem. A 2014, 2, 15823−15828. (d) He, H.; Sun, Q.; Gao, W.; Perman, J. A.; Sun, F.; Zhu, G.; Aguila, B.; Forrest, K.; Space, B.; Ma, S. A Stable MetalOrganic Framework Featuring Local Buffer Environment for Carbon Dioxide Fixation. Angew. Chem., Int. Ed. 2018, 57, 4657−4662. (12) Sheldrick, G. M. SADABS, Program for Empirical Absorption Correction for Area Detector Data; University of Gottingen: Gottingen, Germany, 1996. (13) Sheldrick, G. M. SHELXTL Version 5.1 Software Reference Manual; Bruker AXS Inc.: Madison, WI, 1997. (14) (a) Spek, A. L. Single-Crystal Structure Validation with the Program PLATON. J. Appl. Crystallogr. 2003, 36, 7−13. (b) Spek, A. L. PLATON, A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2001. (15) (a) Zhang, P.; Li, B.; Zhao, Y.; Meng, X.; Zhang, T. A Novel (3, 36)-Connected and Self-Interpenetrated Metal−Organic Framework with High Thermal Stability and Gas-Sorption Capabilities. Chem. Commun. 2011, 47, 7722−7724. (b) Park, J.; Li, J.-R.; Chen, Y.-P.; Yu, J.; Yakovenko, A. A.; Wang, Z. U.; Sun, L.-B.; Balbuena, P. B.; Zhou, H.-C. A Versatile Metal−Organic Framework for Carbon Dioxide Capture and Cooperative Catalysis. Chem. Commun. 2012, 48, 9995− 9997. (16) Ma, S.; Sun, D.; Ambrogio, M.; Fillinger, J. A.; Parkin, S.; Zhou, H.-C. Framework-Catenation Isomerism in MOFs and Its Impact on Hydrogen Uptake. J. Am. Chem. Soc. 2007, 129, 1858−1859. (17) Lee, J.; Li, J.; Jagiello, J. Gas Sorption Properties of Microporous Metal Organic Frameworks. J. Solid State Chem. 2005, 178, 2527− 2532. (18) (a) Ma, S.; Wang, X. S.; Yuan, D.; Zhou, H. C. A Coordinatively Linked, Doubly Interpenetrated, Yb Metal-Organic Framework Demonstrates High Thermal Stability and Uncommon GasAdsorption Selectivity. Angew. Chem., Int. Ed. 2008, 47, 4130−4133. (b) Cheon, Y. E.; Park, J.; Suh, M. P. Selective Gas Adsorption in a Magnesium-based Metal−Organic Framework. Chem. Commun. 2009, 5436−5438.

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We appreciate the financial support from the Tianjin Science and Technology Fund Project for High Education (2017KJ127), Doctoral Program Foundation of Tianjin Normal University (043135202-XB1702), and the NSFC (Grant No. 21501064).



REFERENCES

(1) (a) Bui, M.; Adjiman, C. S.; Bardow, A.; Anthony, E. J.; Boston, A.; Brown, S.; Fennell, P. S.; Fuss, S.; Galindo, A.; Hackett, L. A.; Hallett, J. P.; Herzog, H. J.; Jackson, G.; Kemper, J.; Krevor, S.; Maitland, G. C.; Matuszewski, M.; Metcalfe, I. S.; Petit, C.; Puxty, G.; Reimer, J.; Reiner, D. M.; Rubin, E. S.; Scott, S. A.; Shah, N.; Smit, B.; Martin Trusler, J. P.; Webley, P.; Wilcox, J.; Mac Dowell, N. Carbon Capture and Storage (CCS): the Way Forward. Energy Environ. Sci. 2018, 11, 1062−1176. (2) (a) He, H.; Song, Y.; Zhang, C.; Sun, F.; Yuan, R.; Bian, Z.; Gao, L.; Zhu, G. A Highly Robust Metal−Organic Framework Based on an Aromatic 12−Carboxyl Ligand with Highly Selective Adsorption of CO2 over CH4. Chem. Commun. 2015, 51, 9463−9466. (b) Duan, X.; Lv, R.; Ji, Z.; Li, B.; Cui, Y.; Yang, Y.; Qian, G. Highly Stable Y(III)− based Metal Organic Framework with Two Molecular Building Block for Selective Adsorption of C2H2 and CO2 over CH4. Inorg. Chem. Front. 2018, 5, 1193−1198. (c) Liu, B.; Zhou, H.−F.; Hou, L.; Wang, Y.−Y. Functionalization of MOFs via a Mixed-Ligand Strategy: Enhanced CO2 Uptake by Pore Surface Modification. Dalton Trans. 2018, 47, 5298−5303. (3) (a) He, Y.; Zhou, W.; Qian, G.; Chen, B. Methane Storage in Metal−Organic Frameworks. Chem. Soc. Rev. 2014, 43, 5657−5678. (4) (a) Lee, J. M.; Na, Y.; Han, H.; Chang, S. Cooperative Multi− catalyst Systems for One-pot Organic Transformations. Chem. Soc. Rev. 2004, 33, 302−312. (b) Wasilke, J. C.; Obrey, S. J.; Baker, R. T.; Bazan, G. C. Concurrent Tandem Catalysis. Chem. Rev. 2005, 105, 1001−1020. (c) Trost, B. M. The Atom Economy−a Search for Synthetic Efficiency. Science 1991, 254, 1471−1477. (5) (a) Jiang, J.; Yaghi, O. M. Brønsted Acidity in Metal−Organic Frameworks. Chem. Rev. 2015, 115, 6966−6997. (b) Zhou, H.-C.; Kitagawa, S. Metal−Organic Frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415−5418. (c) Li, J.; Wang, X.; Zhao, G.; Chen, C.; Chai, Z.; Alsaedi, A.; Hayat, T.; Wang, X. Metal−Organic Framework-based Materials: Superior Adsorbents for the Capture of Toxic and Radioactive Metal Ions. Chem. Soc. Rev. 2018, 47, 2322−2356. (6) (a) Liu, Y.; Moon, S.-Y.; Hupp, J. T.; Farha, O. K. Dual-Function Metal−Organic Framework as a Versatile Catalyst for Detoxifying Chemical Warfare Agent Simulants. ACS Nano 2015, 9, 12358−12364. (b) He, H.; Perman, J. A.; Zhu, G.; Ma, S. Metal-Organic Frameworks for CO2 Chemical Transformations. Small 2016, 12, 6309−6324. (7) (a) Zhang, Y.; Li, B.; Krishna, R.; Wu, Z.; Ma, D.; Shi, Z.; Pham, T.; Forrest, K.; Space, B.; Ma, S. Highly Selective Adsorption of Ethylene over Ethane in a MOF Featuring the Combination of Open Metal Site and pi-Complexation. Chem. Commun. 2015, 51, 2714− 2717. (b) He, H.; Sun, F.; Ma, S.; Zhu, G. Reticular Synthesis of a Series of HKUST-like MOFs with Carbon Dioxide Capture and Separation. Inorg. Chem. 2016, 55, 9071−9076. (8) (a) Fu, H.-F.; Yan, L.-B.; Wu, N.-T.; Ma, L.-F.; Zang, S.-Q. DualEmission MOF⊃Dye Sensor for Ratiometric Fluorescence Recognition of RDX and Detection of a Broad Class of Nitro-Compounds. J. Mater. Chem. A 2018, 6, 9183−9191. (b) He, H.; Song, Y.; Sun, F.; Zhao, N.; Zhu, G. Sorption Properties and Nitroaromatic Explosives Sensing Based on Two Isostructural Metal−Organic Frameworks. Cryst. Growth Des. 2015, 15, 2033−2038. (c) Zheng, H.-Y.; Lian, X.; Qin, S.-J.; Yan, B. Lanthanide Hybrids of Covalently-Coordination Cooperative Post-Functionalized Metal−Organic Frameworks for Luminescence Tuning and Highly-Selectively Sensing of Tetrahydrofuran. Dalton Trans. 2018, 47, 6210−6217. (d) He, H.; Song, Y.; F

DOI: 10.1021/acs.inorgchem.8b00938 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (19) Prasad, T. K.; Hong, D. H.; Suh, M. P. High Gas Sorption and Metal−Ion Exchange of Microporous Metal−Organic Frameworks with Incorporated Imide Groups. Chem. - Eur. J. 2010, 16, 14043− 14050. (20) Hayashi, H.; Cote, A. P.; Furukawa, H.; O’Keeffe, M.; Yaghi, O. M. Zeolite A Imidazolate Frameworks. Nat. Mater. 2007, 6, 501−506. (21) (a) Li, R.-J.; Li, M.; Zhou, X.-P.; Li, D.; O’Keeffe, M. A Highly Stable MOF with a Rod SBU and a Tetracarboxylate Linker: Unusual Topology and CO2 Adsorption Behaviour under Ambient Conditions. Chem. Commun. 2014, 50, 4047−4049. (b) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal−Organic Frameworks. Chem. Rev. 2012, 112, 724−781. (22) (a) Si, X.-L.; Jiao, C.-L.; Li, F.; Zhang, J.; Wang, S.; Wang, S.; Liu, S.; Li, Z.-B.; Sun, L.-X.; Xu, F.; Gabelica, Z.; Schick, C. High and Selective CO2 Uptake, H2 Storage and Methanol Sensing on the Amine-Decorated 12-Connected MOF CAU-1. Energy Environ. Sci. 2011, 4, 4522−4527. (b) He, Y.-P.; Tan, Y.-X.; Zhang, J. Tuning a Layer to a Pillared-Layer Metal−Organic Framework for Adsorption and Separation of Light Hydrocarbons. Chem. Commun. 2013, 49, 11323−11325. (23) He, Y.; Zhang, Z.; Xiang, S.; Fronczek, F. R.; Krishna, R.; Chen, B. A Robust Doubly Interpenetrated Metal−Organic Framework Constructed from a Novel Aromatic Tricarboxylate for Highly Selective Separation of Small Hydrocarbons. Chem. Commun. 2012, 48, 6493−6495. (24) Yao, S.; Wang, D.; Cao, Y.; Li, G.; Huo, Q.; Liu, Y. Two Stable 3D Porous Metal−Organic Frameworks with High Performance for Gas Adsorption and Separation. J. Mater. Chem. A 2015, 3, 16627− 16632. (25) (a) Pan, Y.; Yuan, B.; Li, Y.; He, D. Multifunctional Catalysis by Pd@MIL-101: One-Step Synthesis of Methyl Isobutyl Ketone Over Palladium Nanoparticles Deposited on a Metal−Organic Framework. Chem. Commun. 2010, 46, 2280−2282. (b) He, H.; Sun, F.; Aguila, B.; Perman, J. A.; Ma, S.; Zhu, G. A Bifunctional Metal−Organic Framework Featuring the Combination of Open Metal Sites and Lewis Basic Sites for Selective Gas Adsorption and Heterogeneous Cascade Catalysis. J. Mater. Chem. A 2016, 4, 15240−15246. (c) Li, B.; Zhang, Y.; Ma, D.; Li, L.; Li, G.; Shi, Z.; Feng, S. A Strategy toward Constructing a Bifunctionalized MOF Catalyst: Post-Synthetic Modification of MOFs on Organic Ligands and Coordinatively Unsaturated Metal Sites. Chem. Commun. 2012, 48, 6151−6153. (d) Song, F.; Wang, C.; Lin, W. A Chiral Metal−Organic Framework for Sequential Asymmetric Catalysis. Chem. Commun. 2011, 47, 8256− 8258. (e) Zhang, Y.; Wang, Y.; Liu, L.; Wei, N.; Gao, M.-L.; Zhao, D.; Han, Z.-B. Robust Bifunctional Lanthanide Cluster Based Metal− Organic Frameworks (MOFs) for Tandem Deacetalization−Knoevenagel Reaction. Inorg. Chem. 2018, 57, 2193−2198. (26) (a) Zhang, C.; Sun, L.; Yan, Y.; Li, J.; Song, X.; Liu, Y.; Liang, Z. A Luminescent Cadmium Metal−Organic Framework for Sensing of Nitroaromatic Explosives. Dalton Trans. 2015, 44, 230−236. (b) Xie, W.; Zhang, S.-R.; Du, D.-Y.; Qin, J.-S.; Bao, S.-J.; Li, J.; Su, Z.-M.; He, W.-W.; Fu, Q.; Lan, Y.-Q. Stable Luminescent Metal−Organic Frameworks as Dual-Functional Materials to Encapsulate Ln3+ Ions for White-Light Emission and To Detect Nitroaromatic Explosives. Inorg. Chem. 2015, 54, 3290−3296. (c) Hu, X.-L.; Liu, F.-H.; Qin, C.; Shao, K.-Z.; Su, Z.-M. A 2D Bilayered Metal−Organic Framework as a Fluorescent Sensor for Highly Selective Sensing of Nitro Explosives. Dalton Trans. 2015, 44, 7822−7827. (d) Sanda, S.; Parshamoni, S.; Biswas, S.; Konar, S. Highly Selective Detection of Palladium and Picric Acid by a Luminescent MOF: A Dual Functional Fluorescent Sensor. Chem. Commun. 2015, 51, 6576−6579. (e) Zhou, E.-L.; Huang, P.; Qin, C.; Shao, K.-Z.; Su, Z.-M. A Stable Luminescent Anionic Porous Metal−Organic Framework for Moderate Adsorption of CO2 and Selective Detection of Nitro Explosives. J. Mater. Chem. A 2015, 3, 7224−7228. (f) He, H.; Chen, S.-H.; Zhang, D.-Y.; Yang, E.C.; Zhao, X.-J. A Luminescent Metal−Organic Framework as an Ideal Chemosensor for Nitroaromatic Compounds. RSC Adv. 2017, 7, 38871−38876. (g) He, H.; Chen, S.-H.; Zhang, D.-Y.; Hao, R.; Zhang,

C.; Yang, E.-C.; Zhao, X.-J. A Micrometer-Sized Europium(III)− Organic Framework for Selective Sensing of the Cr2O72− Anion and Picric Acid in Water Systems. Dalton Trans. 2017, 46, 13502−13509. (27) (a) Nagarkar, S. S.; Joarder, B.; Chaudhari, A. K.; Mukherjee, S.; Ghost, S. K. Highly Selective Detection of Nitro Explosives by a Luminescent Metal−Organic Framework. Angew. Chem., Int. Ed. 2013, 52, 2881−2885. (b) Sohn, H.; Sailor, M. J.; Magde, D.; Trogler, W. C. Detection of Nitroaromatic Explosives Based on Photoluminescent Polymers Containing Metalloles. J. Am. Chem. Soc. 2003, 125, 3821− 3830.

G

DOI: 10.1021/acs.inorgchem.8b00938 Inorg. Chem. XXXX, XXX, XXX−XXX