Resorcin[4]arene-Based Microporous Metal–Organic Framework as

Oct 31, 2017 - 1 exhibits a fascinating 3D microporous framework structure, in which the free water molecules and the [(CH3)2NH2]+ cations were locate...
0 downloads 13 Views 4MB Size
Research Article www.acsami.org

Cite This: ACS Appl. Mater. Interfaces XXXX, XXX, XXX-XXX

Resorcin[4]arene-Based Microporous Metal−Organic Framework as an Efficient Catalyst for CO2 Cycloaddition with Epoxides and Highly Selective Luminescent Sensing of Cr2O72− Bing-Bing Lu,† Wei Jiang,‡ Jin Yang,*,† Ying-Ying Liu,† and Jian-Fang Ma*,† †

Key Lab of Polyoxometalate Science, Department of Chemistry, Northeast Normal University, Changchun 130024, China Institute of Petrochemical Technology, Jilin Institute of Chemical Technology, Jilin 132022, P. R. China



S Supporting Information *

ABSTRACT: A stable microporous anionic metal−organic framework (MOF), [(CH3)2NH2]6[Cd3L(H2O)2]·12H2O (1), has been synthesized via solvothermal assembly of a new resorcin[4]arene-functionalized dodecacarboxylic acid (H12L) and Cd(II) cations. The constructed MOF (1) was characterized by single-crystal X-ray diffraction and other physicochemical analyses. 1 exhibits a fascinating 3D microporous framework structure, in which the free water molecules and the [(CH3)2NH2]+ cations were located. Remarkably, the exposed Lewis acid Cd(II) sites of activated 1 make it an efficient heterogeneous catalyst for the cycloaddition of CO2 with epoxides at 1 and 20 atm. Importantly, the activated samples of 1 can be reused at least five circles with excellent catalytic performance. Moreover, the fluorescence detection of Cr2O72− and Fe3+ was studied by using 1 as a potential luminescent sensor. KEYWORDS: metal−organic framework, cycloaddition reaction, carbon dioxide, resorcin[4]arene, luminescent detection



INTRODUCTION

MOFs and cages with elegant structures have been assembled.37−39 On the basis of the above consideration, a new resorcin[4]arene-functionalized dodecacarboxylic acid (H12L) was designed (Scheme 1). For the H12L ligand, the carboxylate groups show the flexibility around the −O− groups and variable coordination modes with metals. Assembly of the H12L ligand with Cd(NO3)2·4H2O produces a new 3D microporous MOF [(CH3)2NH2]6[Cd3L(H2O)2]·12H2O (1). Remarkably, activated 1 can be employed as an efficient heterogeneous and recyclable catalyst for CO2 coupling with epoxides at 1 and 20 atm. Strikingly, 1 was also utilized as a luminescent sensor for the detection of Cr2O72− and Fe3+.

Carbon dioxide (CO2) as the primary greenhouse gas is greatly related to the global warming.1−4 Thereby, CO2 elimination is gradually becoming a hot issue.5−7 Thus far, a variety of methods have been developed to convert CO2 into highly valuable products including carbonates, poly(carbonates), carbamate derivatives, and carboxylic acids.8−11 Among the CO2 conversions, the cycloaddition of CO2 with epoxides to generate cyclic carbonates is particularly interesting and attracts considerable attention.12−14 In terms of CO2 cycloaddition with epoxides, a large number of homo- and heterogeneous catalysts have been exploited, such as alkali-metal salts,15 transitionmetal complexes,16 ionic liquids,17 zeolites,18,19 metal oxides,20 and so forth. In contrast to these mentioned catalysts, metal− organic frameworks (MOFs) as heterogeneous catalysts have some advantages because of their high surface area, tunable pore size, and functionality.21−25 It is well-established that organic linkers play an important role in the assembly of MOF catalysts with tunable structural features.26−30 In this regard, multicarboxylate-based ligands have been extensively employed in constructing stable MOF catalysts.31,32 Over the past several years, we have conducted a systematic study for the functionalization of resorcin[4]arene unit by various tetracarboxylates, octacarboxylates, and so on.33−36 By using the resorcin[4]arene-based ligands, a series of © XXXX American Chemical Society



EXPERIMENTAL SECTION

Materials and Methods. All chemicals were commercially available and were used without further purification. Powder X-ray diffraction (PXRD) patterns were determined on a Rigaku D/MAX2000 X-ray diffractometer with graphite-monochromatized Cu Kα radiation (λ = 0.154 nm). Fourier transform infrared spectra were recorded on a Mattson Alpha Centauri spectrometer. Elemental analysis data for C, H, and N were measured on a EuroVector EA3000 analyzer. 1H NMR spectra were carried out in CDCl3 on a Varian 500 Received: September 18, 2017 Accepted: October 20, 2017

A

DOI: 10.1021/acsami.7b14179 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces Scheme 1. Synthetic Strategy for the H12L Ligand

MHz. Catalytic products for the cycloaddition of CO2 were determined by gas chromatography (GC) equipment with a capillary (30 m long × 0.25 mm i.d., WondaCap 17) and a flame ionization detector (GC-2014C, Shimadzu, Japan). Synthesis of 1,3-Bis[(methoxycarbonyl)methoxy]benzene (A). A mixture of resorcinol (22.2 g, 200 mmol), anhydrous K2CO3 (82.8 g, 600 mmol), methyl chloroacetate (54.3 g, 500 mmol), KI (0.5 g, 3 mmol), and 300 mL of MeCN was stirred and heated at reflux under N2 for 12 h. Excess methyl chloroacetate and acetone were removed under reduced pressure, and the residue was partitioned between chloroform and water. The chloroform phase was extracted three times with water and dried with anhydrous Na2SO4. Yellow oil was produced after removing the chloroform under reduced pressure. Then, diethyl ether was put into the oil, and 1,3-bis[(methoxycarbonyl)methoxy]benzene (A) was achieved in a 76% yield (Scheme 1). Synthesis of Methyl 2-(4-Formylphenoxy)acetate (B). A mixture of 4-hydroxybenzaldehyde (24.4 g, 200 mmol), anhydrous K2CO3 (55.2 g, 400 mmol), methyl chloroacetate (32.6 g, 300 mmol), KI (0.5 g, 3 mmol), and MeCN (300 mL) was placed in a N2 atmosphere and heated at reflux for 12 h. After the removal of the solvent under reduced pressure, the raw material was extracted by chloroform and washed three times with water. The chloroform phase was dried with anhydrous Na2SO4. Then, product B was produced after the removal of the chloroform under reduced pressure in an 85% yield (Scheme 1). Synthesis of H12L. Boron trifluoride diethyl etherate (14.19 g, 100 mmol) was put into a mixture of A (12.7 g, 50 mmol), B (9.7 g, 50 mmol), and anhydrous dichloromethane (100 mL) under ice bath. The formed mixture was stirred for 48 h at ambient temperature. Then, water (100 mL) was added, and the formed layers were separated. After washing the mixture with water three times, the organic phase was dried with anhydrous Na2SO4, and yellow oil was produced after removing the dichloromethane. The precursor of H12L was achieved in an 83% yield after the addition of acetone to the oil. A mixture of the precursor of H12L (13.7 g, 8 mmol), tetrahydrofuran (50 mL), sodium hydroxide (4.8 g, 120 mmol), and water (50 mL) was heated at reflux for 8 h. Tetrahydrofuran was removed, and the pH value was adjusted to 1−2 by using HCl (1.0 mol·L−1). H12L was achieved in a 71% yield after being washed with water. Synthesis of [(CH3)2NH2]6[Cd3L(H2O)2]·12H2O (1). H12L (0.015 g, 0.006 mmol) and Cd(Ac)2·2H2O (0.019 g, 0.08 mmol) were dissolved in a mixture of H2O (2 mL) and dimethylformamide (DMF) (6 mL). Then, the mixture was sealed in a Teflon reactor (15 mL) and heated at 120 °C for 6 days. 1 was achieved in a 41% yield based on H12L.

Anal. Calcd for C88H128N6O50Cd3 (Mr = 2407.16): C, 43.90; H, 5.36; N, 3.49. Found: C, 43.99; H, 5.25; N, 3.32. IR data (KBr, cm−1): 3067 (w), 2925 (w), 2797 (w), 1581 (s), 1494 (s), 1468 (m), 1414 (s), 1281 (s), 1234 (s), 1177 (m), 1101 (w), 1059 (m), 709 (w). X-ray Crystallography. Crystallographic data for 1 were collected on an Oxford Diffraction Gemini R CCD diffractometer with graphitemonochromated Mo Kα radiation (λ = 0.71073 Å) at 293 K. The structure of 1 was solved by direct methods and refined on F2 by fullmatrix least squares using the SHELXS-2014 program within WingX.40−42 All nonhydrogen atoms were refined anisotropically. Hydrogen atoms attached to carbons were generated geometrically. Experimental details of the crystallographic analyses are listed in Table S1. Cycloaddition of CO2 to Epoxides. Prior to the catalytic reaction, the samples were activated by soaking in acetone for 24 h and then dried at 393 K under vacuum for 12 h to generate the activated catalyst 1. All of the catalytic reactions at ambient pressure were carried out in a 10 mL round-bottom flask. Typically, epoxide (5 mmol), tetra-n-butylammonium bromide (TBAB) (0.16 g, 0.5 mmol), and activated 1 (5 mg, 0.002 mmol) were added and then stirred. The catalytic reactions at high pressure were performed in a 25 mL stainless-steel high-pressure reactor. A mixture of epoxide (20 mmol), TBAB (0.32 g, 1 mmol), and activated catalyst 1 (10 mg, 0.004 mmol) was placed in the reactor and purified twice by CO2. Then, the required pressure (20 atm) and temperature (80 °C) were attained. The product yields were determined by GC and further confirmed by 1 H NMR. Measurement of Luminescent Sensing. The luminescent detection for anions and metal cations was conducted by the literature method.43,44 For all measurements of luminescent sensing, 3 mg of sample of 1 was employed. Typically, finely ground sample of 1 (3 mg) was dispersed in aqueous solution (3 mL) of various anions or metal cations. After sonication treatment, the formed suspensions were transferred into a Suprasil cuvette and luminescent detection was conducted.



RESULTS AND DISCUSSION Crystal Structure of [(CH3)2NH2]6[Cd3L(H2O)2]·12H2O (1). Crystal samples of 1 were synthesized by using Cd(Ac)2· 2H2O and H12L under solvothermal conditions. Crystallographic analysis shows that 1 crystallized in the monoclinic space group C2/c. As illustrated in Figure 1a, the asymmetric structure of 1 contains one and a half of Cd(II) ions, half a L12− B

DOI: 10.1021/acsami.7b14179 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces

Table 1. Cycloaddition of CO2 with Epichlorohydrin at Different Conditionsa entry

TBAB (mmol)

1 (mg)

temperature (°C)

time (h)

yield (%)b

1 2 3 4 5

0 0.25 0.5 0.25 0.5

5 5 5 0 5

50 50 50 50 50

4 4 4 4 12

41 79 88 55 99

a Reaction conditions: 1 (5 mg, 0.002 mmol based on Lewis acid sites), epoxide (5 mmol), and CO2 (1 atm). bIsolated yields were calculated by GC.

Table 2. Cycloaddition of CO2 to Epoxides at 1 atma

a

Reaction conditions: 1 (5 mg, 0.002 mmol based on Lewis acid sites), epoxide (5 mmol), TBAB (0.16 g, 0.5 mmol), and CO2 (1 atm). b Isolated yields were calculated by GC, Tp = 50 °C. cIsolated yields were calculated by GC, Tp = 80 °C. dTurnover number (TON) = [mmol (product)]/[mmol (Lewis acid sites)].

Figure 1. (a) Coordination environments of Cd(II) in 1. (b) Coordination mode of L12− cation in 1. (c) View of the 3D microporous framework structure of 1. (d) View of the 3D structure of 1 along the b axis. Symmetry codes: #1 −x, y, −z + 1/2; #2 x, −y + 1, z − 1/2; #3 −x, −y + 1, −z + 1; #4 x, −y + 1, z + 1/2; #5 −x + 1/2, y − 1/2, −z + 3/2; #6 −x, y, −z + 3/2.

Scheme 2. Cycloaddition of CO2 with Epichlorohydrin

Figure 2. Recycling experiments for catalyst 1 at 1 atm.

anion, one coordinated water, and three [(CH3)2NH2]+ cations. The [(CH3)2NH2]+ cation was produced by the decomposition of the DMF solvent and balances the negative charge of the framework. Two Cd(II) cations display different coordination environments. Cd1 is six-coordinated with two coordinated water molecules and four carboxylate oxygen

atoms from two L12− anions in a distorted octahedral geometry, whereas Cd2 is surrounded by eight carboxylate oxygen atoms from four L12− anions. The two L12− anions exhibit two types of coordination modes: one L12− anion bridges 2 Cd(II) cations while the other links 10 Cd(II) cations (Figure 1b). In these fashions, neighboring Cd(II) cations were linked by the L12− C

DOI: 10.1021/acsami.7b14179 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces Table 3. Cycloaddition of CO2 with Epoxides at 20 atma,b

(5 mg) at 50 °C and 1 atm for 4 h, the conversion yield of the cyclic carbonate reached only 41% (entry 1, Table 1; Figure S2a). The conversion yield markedly increased to 79 and 88%, respectively, when the cocatalyst TBAB was added to the reaction system with the contents of 0.25 and 0.5 mmol (entries 2 and 3; Figure S2b,c). By contrast, the conversion yield of 55% was achieved when only TBAB was used as the catalyst (0.25 mmol) (entry 4; Figure S2d), suggesting that 1 plays an important role in the catalytic process. By further prolonging the reaction time to 12 h, the conversion yield was enhanced to 99% (entry 5; Figure S2e). To further investigate the catalytic generality of 1, different epoxide substrates were employed for the CO2 cycloaddition reaction at 1 atm and 50 °C with a 1/TBAB loading of 0.004:1. The epoxides with variable alkyl chains were first chosen as the reaction substrates. As shown in Table 2, the conversion yields markedly decreased from 97% to trace with the increasing alkyl chain length of the epoxides (entries 2−5; Figure S3a−d). Noticeably, the catalytic performance is related to the microporous framework of catalyst 1 to some extent. The decreasing conversion yields were probably attributed to the stereohindrance of the substrates. The transportation of the substrates and products through the microporous framework of catalyst 1 was greatly influenced by the long alkyl chains during the catalytic reaction, thus leading to the much lower conversion yields. For the epoxides with long alkyl chains, the conversion yields were significantly increased with elevated temperatures from 50 to 80 °C (entries 4 and 5; Figure S3e,f). For the substrates cyclohexene oxide and 1,2-epoxyethylbenzene, 42 and 74% conversion yields were achieved (entries 7 and 8; Figure S3h,i), respectively. By contrast, benzyl glycidyl ether and glycidyl phenyl ether were completely converted to the products under the same conditions (entries 9 and 10; Figure S3j,k). The excellent conversions of benzyl glycidyl ether and glycidyl phenyl ether were mainly ascribed to their electron-withdrawing groups in their structures.50−52 Moreover, catalyst 1 can be easily separated by simple filtration and washing with dichloromethane. The separated samples were dried in the air and then reused at least five circles with excellent catalytic performance (Figure 2). The PXRD pattern of catalyst 1 after catalysis corresponds well to the simulated one (Figure S1), indicating that 1 is a potential recyclable heterogeneous catalyst. Cycloaddition of CO2 with Epoxides at 20 atm. The cycloaddition of CO2 with epoxides was also carried out by using activated 1 and TBAB as a cocatalyst at 20 atm. The catalytic reaction was performed with a fixed catalyst 1/ substrate ratio of 1:5000 at 80 °C and 20 atm. As shown in Table 3, the substrate 2-(phenoxymethyl)oxirane was completely converted to the product within 2 h with a much higher TON of 4950 (entry 1; Figures S4a and S5). For epichlorohydrin, the overall conversion reached >99% within 4 h (entry 2; Figures S4b and S6). On the contrary, the conversion yield of the substrate cyclohexene oxide is only up to 80% within 6 h (entry 3; Figures S4c and S7). Although the reaction time was prolonged to 10 h, the conversion yield was not effectively improved. For the remaining substrates, their conversion yields reached more than 90% within 6 h or 8 h (entries 4−10; Figures S4d−j and S8−S14). Importantly, catalyst 1 can be reused at least five circles with high catalytic activity (Figure 3). The PXRD pattern after catalysis at 20 atm

a

Reaction conditions: 1 (10 mg, 0.004 mmol based on Lewis acid sites), epoxide (20 mmol), TBAB (0.32 g, 1 mmol), CO2 (20 atm), and Tp = 80 °C. bIsolated yields were calculated by GC; the catalytic products were confirmed by 1H NMR. cTurnover number (TON) = [mmol (product)]/[mmol (Lewis acid sites)].

Figure 3. Recycling experiments for catalyst 1 at 20 atm.

anions into a fascinating 3D framework with two kinds of open channels (ca. 5.4 Å × 5.4 Å) and (ca. 4.8 Å × 8.2 Å) (Figure 1c,d), in which the [(CH3)2NH2]+ cations and water molecules were located. The solvent-accessible volume of 1 was estimated by the PLATON calculation to be ca. 37.2% of the unit cell volume (10 356.2 Å3).45 Catalytic Cycloaddition of CO2 with Epoxides at 1 atm. The exposed Lewis acid Cd(II) sites of the activated samples encouraged us to evaluate the catalytic performance for the chemical fixation of CO2 with epoxides at 1 and 20 atm.46−48 Prior to the catalytic reaction, the samples of 1 were activated by immersing in acetone for 24 h and then dried at 393 K under vacuum for 12 h to give the activated samples of 1.49 As shown in Figure S1, the PXRD patterns of the activated sample of 1 correspond well to the simulated one. To achieve the optimum reaction condition, epichlorohydrin was chosen as a typical substrate for the cycloaddition of CO2 (Scheme 2). When the catalytic reaction was conducted by using catalyst 1 D

DOI: 10.1021/acsami.7b14179 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces

Scheme 3. Proposed Catalytic Mechanism for the Cycloaddition of CO2 with Epoxides Using 1 and TBAB as a Cocatalysta

a

Symmetry codes:

#1

−x, y, −z + 1/2;

#2

x, −y + 1, z − 1/2;

#3

−x, −y + 1, −z + 1.

Figure 4. Luminescence spectra of H12L and 1 in the solid state.

Figure 5. Luminescence spectra and intensities of 1 in water and aqueous solutions of various metal cations.

resulting in the generation of metal carbonates.57,58 Finally, an intramolecular ring-closure reaction occurs for the metal carbonate intermediate, where the intramolecular carbonate O atom attacks the C−Br carbon to yield a cyclic carbonate.59,60 Luminescence Properties. The fluorescence performance of MOFs with d10 metal cations has received considerable attention because of their widespread applications in sensing, labeling, and electroluminescence.61,62 Herein, the luminescence spectra of 1 and free H12L were determined in the solid state. As illustrated in Figure 4, the emission spectrum of H12L

matches well with the simulated one (Figure S1), indicating the good structural stability in the recycled catalytic experiments. On the basis of our experiments and previously reported works, a possible mechanism for the CO2 cycloaddition to epoxides catalyzed by 1 was proposed (Scheme 3). The Cd(II) centers of activated 1 as Lewis acid sites are involved in coordination with the epoxides.53,54 Then, the Br− ion of TBAB engages in a nucleophilic attack on the epoxide substrate, leading to the ring opening of the epoxide.55,56 The O atom of the ring-opened epoxide further interacted with CO2, thus E

DOI: 10.1021/acsami.7b14179 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces

Figure 6. Luminescence spectra and intensities of 1 in water and aqueous solutions of various concentrations of Fe3+ cations.

Figure 7. Luminescence spectra and intensities of 1 in water and various aqueous solutions of anions.

Figure 8. Luminescence spectra and intensities of 1 in aqueous solutions of various concentrations of Cr2O72−.

(M = Na+, Cd2+, Al3+, Mn2+, Cu2+, Co2+, Cr3+, Ni2+, and Fe3+). The luminescence spectra of the samples with the loaded metal ions were determined in the suspensions. As shown in Figure 5, the loaded metal cation species significantly influenced the luminescence intensities. For example, the luminescence intensities (λem = 450 nm) of 1 increased with the loading of Na+ cations. On the contrary, the luminescence intensities (λem = 450 nm) were obviously reduced when other metal cations were loaded; especially, the Fe3+ cation exhibits the most significant quenching effect on the emission intensity of 1. To further investigate the fluorescence responses of 1 toward Fe3+ cations, the luminescence spectra of 1 in different

exhibited an emission peak at 390 nm (λex = 320 nm), which may arise from the π* → π or π* → n transitions.63,64 The emission peak of 1 was located at 450 nm (λex = 380 nm), which is sharply red-shifted (60 nm) with respect to H12L. This probably results from the coordination of Cd(II) cations with the H12L ligand.65 Sensing of Fe3+ Cations. The selective detection of Fe3+ is greatly important because its content is highly related to human body.66 In view of the intense emission of 1, its potential application for the detection of metal cations was examined in detail. Typically, the finely ground samples of 1 (3 mg) were dispersed in aqueous solutions of AgNO3, Pb(NO3)2, and MClx F

DOI: 10.1021/acsami.7b14179 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces concentrations of Fe3+ were detected. As illustrated in Figure 6, the emission intensities of 1 decline sharply with the increasing Fe3+ concentrations from 0 to 0.01 M; especially, the luminescence intensity almost disappeared at the Fe 3+ concentration of 0.01 M. Ksv is 2.67 × 105, calculated from the Stern−Volmer equation I0/I = 1 + Ksv[M].67 The result suggests that 1 is a potential luminescent sensor for the detection of Fe3+. Sensing of Cr2O72−. Hexavalent chromium (Cr6+) as a type of toxic anion pollutant has received much concern because of its increasing utilization in agriculture and industries.68 Therefore, rapid and efficient detection of trace chromate anions in water is particularly important.69 Herein, the luminescent detection of anions was explored by 1. The powder samples of 1 (3 mg) were immersed in aqueous solutions of KXn (0.01 mol·L−1, X = SO42−, IO3−, CO32−, SCN−, S2O82−, Cr2O72−, Cl−, Br−, and I−), and their luminescent behaviors were determined. As depicted in Figure 7, the luminescence intensities decreased and greatly related to the types of the anion species. Remarkably, the Cr2O72− anions show the most observable quenching effect on the emission of the suspension of 1 (Figure 7). To better understand the luminescence sensitivity of 1 toward Cr2O72− anions, different concentrations (0−0.01 M) of Cr2O72− in aqueous solutions were selected to accomplish the luminescence-sensitive experiments. As depicted in Figure 8, the luminescence intensities decreased gradually with the increasing concentrations of Cr2O72− and were finally completely quenched when the concentration of Cr2O72− reached 5 × 10−3 M. According to the experimental quenching data, the calculated Ksv is 9.19 × 105, indicating that 1 was a potential fluorescent candidate probe for Cr2O72− anions.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (grant nos. 21471029 and 21371030).



(1) Li, J.-R.; Ma, Y.; McCarthy, M. C.; Sculley, J.; Yu, J.; Jeong, H.-K.; Balbuena, P. B.; Zhou, H.-C. Carbon Dioxide Capture-Related Gas Adsorption and Separation in Metal-Organic Frameworks. Coord. Chem. Rev. 2011, 255, 1791−1823. (2) Lee, W. R.; Hwang, S. Y.; Ryu, D. W.; Lim, K. S.; Han, S. S.; Moon, D.; Choi, J.; Hong, C. S. Diamine-Functionalized Metal− Organic Framework: Exceptionally High CO2 Capacities from Ambient Air and Flue Gas, Ultrafast CO2 Uptake Rate, and Adsorption Mechanism. Energy Environ. Sci. 2014, 7, 744−751. (3) Liu, J.; Thallapally, P. K.; McGrail, B. P.; Brown, D. R.; Liu, J. Progress in Adsorption-Based CO2 Capture by Metal−Organic Frameworks. Chem. Soc. Rev. 2012, 41, 2308−2322. (4) 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. (5) Tu, W.; Zhou, Y.; Zou, Z. Photocatalytic Conversion of CO2 into Renewable Hydrocarbon Fuels: State-of-The-Art Accomplishment, Challenges, and Prospects. Adv. Mater. 2014, 26, 4607−4626. (6) McDonald, T. M.; Mason, J. A.; Kong, X.; Bloch, E. D.; Gygi, D.; Dani, A.; Crocellà, V.; Giordanino, F.; Odoh, S. O.; Drisdell, W. S.; Vlaisavljevich, B.; Dzubak, A. L.; Poloni, R.; Schnell, S. K.; Planas, N.; Lee, K.; Pascal, T.; Wan, L. F.; Prendergast, D.; Neaton, J. B.; Smit, B.; Kortright, J. B.; Gagliardi, L.; Bordiga, S.; Reimer, J. A.; Long, J. R. Cooperative Insertion of CO2 in Diamine-Appended Metal-Organic Frameworks. Nature 2015, 519, 303−308. (7) Dey, S.; Bhunia, A.; Breitzke, H.; Groszewicz, P. B.; Buntkowsky, G.; Janiak, C. Two Linkers are Better Than One: Enhancing CO2 Capture and Separation with Porous Covalent Triazine-Based Frameworks from Mixed Nitrile Linkers. J. Mater. Chem. A 2017, 5, 3609−3620. (8) Lu, X.-B.; Darensbourg, D. J. Cobalt Catalysts for The Coupling of CO2 and Epoxides to Provide Polycarbonates and Cyclic Carbonates. Chem. Soc. Rev. 2012, 41, 1462−1484. (9) Zhou, Z.; He, C.; Xiu, J.; Yang, L.; Duan, C. Metal−Organic Polymers Containing Discrete Single-Walled Nanotube as A Heterogeneous Catalyst for The Cycloaddition of Carbon Dioxide to Epoxides. J. Am. Chem. Soc. 2015, 137, 15066−15069. (10) Ai, J.; Min, X.; Gao, C.-Y.; Tian, H.-R.; Dang, S.; Sun, Z.-M. A Copper-Phosphonate Network as A High-Performance Heterogeneous Catalyst for The CO2 Cycloaddition Reactions and Alcoholysis of Epoxides. Dalton Trans. 2017, 46, 6756−6761. (11) Li, J.; Jia, D.; Guo, Z.; Liu, Y.; Lyu, Y.; Zhou, Y.; Wang, J. Imidazolinium Based Porous HypercrossLinked Ionic Polymers for Efficient CO2 Capture and Fixation with Epoxides. Green Chem. 2017, 19, 2675−2686. (12) Chatelet, B.; Joucla, L.; Dutasta, J.-P.; Martinez, A.; Szeto, K. C.; Dufaud, V. Azaphosphatranes as Structurally Tunable Organocatalysts for Carbonate Synthesis from CO2 and Epoxides. J. Am. Chem. Soc. 2013, 135, 5348−5351. (13) Escárcega-Bobadilla, M. V.; Belmonte, M. M.; Martin, E.; Escudero-Adán, E. C.; Kleij, A. W. A Recyclable Trinuclear Bifunctional Catalyst Derived from A Tetraoxo Bis-Zn(salphen) Metalloligand. Chem.Eur. J. 2013, 19, 2641−2648. (14) Chen, J.; Li, H.; Zhong, M.; Yang, Q. Hierarchical Mesoporous Organic Polymer with An Intercalated Metal Complex for The Efficient Synthesis of Cyclic Carbonates from Flue Gas. Green Chem. 2016, 18, 6493−6500. (15) Qu, J.; Cao, C.-Y.; Dou, Z.-F.; Liu, H.; Yu, Y.; Li, P.; Song, W.G. Synthesis of Cyclic Carbonates: Catalysis by an Iron-Based



CONCLUSIONS In summary, a stable microporous MOF (1) has been constructed by the new resorcin[4]arene-functionalized dodecacarboxylic acid and Cd(II) cations. The open Lewis acid Cd(II) sites of activated 1 make it a promising heterogeneous catalyst for the cycloaddition of CO2 with various epoxides. Most importantly, 1 could be easily separated and recycled at least five circles without any noticeable decrease in its catalytic activity. Besides, fluorescent 1 was used to probe Fe3+ and Cr2O72− in aqueous solutions with high selectivity and sensitivity.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.7b14179.



REFERENCES

PXRD patterns, GC spectra, 1H NMR spectra, and tables (PDF) Crystallographic data of C88H102Cd3N6O38·12H2O (CIF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (J.Y.). *E-mail: [email protected]. Fax: +86-431-85098620 (J.F.M). ORCID

Jian-Fang Ma: 0000-0002-4059-8348 G

DOI: 10.1021/acsami.7b14179 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces

Shaped Asymmetric Multicarboxylate and N-Donor Ligands. Cryst. Growth Des. 2013, 13, 4695−4704. (32) Ou, Y.-C.; Gao, X.; Zhou, Y.; Chen, Y.-C.; Wang, L.-F.; Wu, J.Z.; Tong, M.-L. Magnetic Properties and Photoluminescence of Lanthanide Coordination Polymers Constructed with ConformationFlexible Cyclohexane-Tetracarboxylate Ligands. Cryst. Growth Des. 2016, 16, 946−952. (33) 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. (34) 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. (35) 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. (36) Zhang, H.; Yang, J.; Liu, Y.-Y.; Song, S.; Ma, J.-F. A Family of Metal−Organic Frameworks with A New Chair-Conformation Resorcin[4]arene-Based Ligand: Selective Luminescent Sensing of Amine and Aldehyde Vapors, and Solvent-Mediated Structural Transformations. Cryst. Growth Des. 2016, 16, 3244−3255. (37) 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. (38) Hu, Y.-J.; Yang, J.; Liu, Y.-Y.; Song, S.; Ma, J.-F. A Family of Capsule-Based Coordination Polymers Constructed from A New Tetrakis(1,2,4-triazol-ylmethyl)resorcin[4]arene Cavitand and Varied Dicarboxylates for Selective Metal-Ion Exchange and Luminescent Properties. Cryst. Growth Des. 2015, 15, 3822−3831. (39) Dong, Y.-B.; Shi, H.-Y.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. Molecular Dumbbell, Sandwich, and Paddle-Wheel Assembled with Methylresorcin[4]arene Cavitands and Organooxotin Clusters. Cryst. Growth Des. 2015, 15, 1546−1551. (40) Sheldrick, G. M. SHELXS-97, Programs for X-ray Crystal Structure Solution; University of Göttingen: Göttingen, Germany, 1997. (41) Sheldrick, G. M. SHELXL-97, Programs for X-ray Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 1997. (42) Farrugia, L. J. WINGX, A Windows Program for Crystal Structure Analysis; University of Glasgow: Glasgow, UK, 1988. (43) Han, T.-T.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. Rhodamine 6G Loaded Zeolitic Imidazolate Framework-8 (ZIF-8) Nanocomposites for Highly Selective Luminescent Sensing of Fe3+, Cr6+ and Aniline. Microporous Mesoporous Mater. 2016, 228, 275−288. (44) Shi, M.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. Four Coordination Polymers Based on 1,4,8,11-Tetrazacyclotetradecane-N,N′,N″,N‴Tetra-Methylene-Benzoic Acid: Syntheses, Structures, and Selective Luminescence Sensing of Iron(III) Ions, Dichromate Anions, and Nitrobenzene. Dyes Pigm. 2016, 129, 109−120. (45) Spek, A. L. Single-Crystal Structure Validation with the Program PLATON. J. Appl. Crystallogr. 2003, 36, 7−13. (46) Wang, Z.; Wang, L.; Jiang, Y.; Hunger, M.; Huang, J. Cooperativity of Brønsted and Lewis Acid Sites on Zeolite for Glycerol Dehydration. ACS Catal. 2014, 4, 1144−1147. (47) Nakajima, K.; Baba, Y.; Noma, R.; Kitano, M.; Kondo, J. N.; Hayashi, S.; Hara, M. Nb2O5·nH2O as A Heterogeneous Catalyst with Water-Tolerant Lewis Acid Sites. J. Am. Chem. Soc. 2011, 133, 4224− 4227. (48) Kathalikkattil, A. C.; Kim, D.-W.; Tharun, J.; Soek, H.-G.; Roshan, R.; Park, D.-W. Aqueous-Microwave Synthesized Carboxyl Functional Molecular Ribbon Coordination Framework Catalyst for the Synthesis of Cyclic Carbonates from Epoxides and CO2. Green Chem. 2014, 16, 1607−1616.

Composite and the Role of Hydrogen Bonding at the Solid/Liquid Interface. ChemSusChem 2012, 5, 652−655. (16) Ma, R.; He, L.-N.; Zhou, Y.-B. An Efficient and Recyclable Tetraoxo-Coordinated Zinc Catalyst for the Cycloaddition of Epoxides with Carbon Dioxide at Atmospheric Pressure. Green Chem. 2016, 18, 226−231. (17) Wang, X.; Zhou, Y.; Guo, Z.; Chen, G.; Li, J.; Shi, Y.; Liu, Y.; Wang, J. Heterogeneous Conversion of CO2 Into Cyclic Carbonates at Ambient Pressure Catalyzed by Ionothermal-Derived Meso-Macroporous Hierarchical Poly(ionic liquid)s. Chem. Sci. 2015, 6, 6916− 6924. (18) Kuruppathparambil, R. R.; Babu, R.; Jeong, H. M.; Hwang, G.Y.; Jeong, G. S.; Kim, M.-I.; Kim, D.-W.; Park, D.-W. A Solid Solution Zeolitic Imidazolate Framework as A Room Temperature Efficient Catalyst for the Chemical Fixation of CO2. Green Chem. 2016, 18, 6349−6356. (19) Miralda, C. M.; Macias, E. E.; Zhu, M.; Ratnasamy, P.; Carreon, M. A. Zeolitic Imidazole Framework-8 Catalysts in the Conversion of CO2 to Chloropropene Carbonate. ACS Catal. 2012, 2, 180−183. (20) Yamaguchi, K.; Ebitani, K.; Yoshida, T.; Yoshida, H.; Kaneda, K. Mg−Al Mixed Oxides as Highly Active Acid−Base Catalysts for Cycloaddition of Carbon Dioxide to Epoxides. J. Am. Chem. Soc. 1999, 121, 4526−4527. (21) Gao, W.-Y.; Chen, Y.; Niu, Y.; Williams, K.; Cash, L.; Perez, P. J.; Wojtas, L.; Cai, J.; Chen, Y.-S.; Ma, S. Crystal Engineering of An nbo Topology Metal-Organic Framework for Chemical Fixation of CO2 under Ambient Conditions. Angew. Chem., Int. Ed. 2014, 53, 2615−2619. (22) Beyzavi, M. H.; Klet, R. C.; Tussupbayev, S.; Borycz, J.; Vermeulen, N. A.; Cramer, C. J.; Stoddart, J. F.; Hupp, J. T.; Farha, O. K. A Hafnium-Based Metal−Organic Framework as An Efficient and Multifunctional Catalyst for Facile CO2 Fixation and Regioselective and Enantioretentive Epoxide Activation. J. Am. Chem. Soc. 2014, 136, 15861−15864. (23) Feng, D.; Chung, W.-C.; Wei, Z.; Gu, Z.-Y.; Jiang, H.-L.; Chen, Y.-P.; Darensbourg, D. J.; Zhou, H.-C. Construction of Ultrastable Porphyrin Zr Metal−Organic Frameworks through Linker Elimination. J. Am. Chem. Soc. 2013, 135, 17105−17110. (24) Li, P.-Z.; Wang, X.-J.; Liu, J.; Lim, J. S.; Zou, R.; Zhao, Y. A Triazole-Containing Metal−Organic Framework as A Highly Effective and Substrate Size-Dependent Catalyst for CO2 Conversion. J. Am. Chem. Soc. 2016, 138, 2142−2145. (25) Zheng, J.; Wu, M.; Jiang, F.; Su, W.; Hong, M. Stable Porphyrin Zr and Hf Metal−Organic Frameworks Featuring 2.5 nm Cages: High Surface Areas, SCSC Transformations and Catalyses. Chem. Sci. 2015, 6, 3466−3470. (26) Jiang, W.; Yang, J.; Liu, Y.-Y.; Song, S.-Y.; Ma, J.-F. A Stable Porphyrin-Based Porous mog Metal−Organic Framework as An Efficient Solvent-Free Catalyst for C−C Bond Formation. Inorg. Chem. 2017, 56, 3036−3043. (27) Jiang, W.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. Porphyrin-Based MixedValent Ag(I)/Ag(II) and Cu(I)/Cu(II) Networks as Efficient Heterogeneous Catalysts for the Azide−Alkyne “click” Reaction and Promising Oxidation of Ethylbenzene. Chem. Commun. 2016, 52, 1373−1376. (28) Wu, H.; Yang, J.; Su, Z.-M.; Batten, S. R.; Ma, J.-F. An Exceptional 54-Fold Interpenetrated Coordination Polymer with 103srs Network Topology. J. Am. Chem. Soc. 2011, 133, 11406−11409. (29) Schoedel, A.; Li, M.; Li, D.; O’Keeffe, M.; Yaghi, O. M. Structures of Metal−Organic Frameworks with Rod Secondary Building Units. Chem. Rev. 2016, 116, 12466−12535. (30) 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. (31) Yang, W.; Wang, C.; Ma, Q.; Feng, X.; Wang, H.; Jiang, J. SelfAssembled Zn(II) Coordination Complexes Based on Mixed VH

DOI: 10.1021/acsami.7b14179 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces (49) Zou, R.-Q.; Sakurai, H.; Han, S.; Zhong, R.-Q.; Xu, Q. Probing the Lewis Acid Sites and CO Catalytic Oxidation Activity of the Porous Metal−Organic Polymer [Cu(5-methylisophthalate)]. J. Am. Chem. Soc. 2007, 129, 8402−8403. (50) Castro-Osma, J. A.; Lamb, K. J.; North, M. Cr(salophen) Complex Catalyzed Cyclic Carbonate Synthesis at Ambient Temperature and Pressure. ACS Catal. 2016, 6, 5012−5025. (51) Yang, S.; Peng, L.; Huang, P.; Wang, X.; Sun, Y.; Cao, C.; Song, W. Nitrogen, Phosphorus, and Sulfur Co-Doped Hollow Carbon Shell as Superior Metal-Free Catalyst for Selective Oxidation of Aromatic Alkanes. Angew. Chem., Int. Ed. 2016, 55, 4016−4020. (52) Lu, B.-B.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. A Polyoxovanadate− Resorcin[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. (53) Ugale, B.; Dhankhar, S. S.; Nagaraja, C. M. Construction of 3Fold-Interpenetrated Three-Dimensional Metal−Organic Frameworks of Nickel(II) for Highly Efficient Capture and Conversion of Carbon Dioxide. Inorg. Chem. 2016, 55, 9757−9766. (54) Wang, J.-C.; Ding, F.-W.; Ma, J.-P.; Liu, Q.-K.; Cheng, J.-Y.; Dong, Y.-B. Co(II)-MOF: A Highly Efficient Organic Oxidation Catalyst with Open Metal Sites. Inorg. Chem. 2015, 54, 10865−10872. (55) Ema, T.; Miyazaki, Y.; Shimonishi, J.; Maeda, C.; Hasegawa, J.-y. Bifunctional Porphyrin Catalysts for the Synthesis of Cyclic Carbonates from Epoxides and CO2: Structural Optimization and Mechanistic Study. J. Am. Chem. Soc. 2014, 136, 15270−15279. (56) Babu, R.; Roshan, R.; Gim, Y.; Jang, Y. H.; Kurisingal, J. F.; Kim, D. W.; Park, D.-W. Inverse Relationship of Dimensionality and Catalytic Activity in CO2 Transformation: A Systematic Investigation by Comparing Multidimensional Metal−Organic Frameworks. J. Mater. Chem. A 2017, 5, 15961−15969. (57) North, M.; Pasquale, R.; Young, C. Synthesis of Cyclic Carbonates from Epoxides and CO2. Green Chem. 2010, 12, 1514− 1539. (58) Babu, R.; Roshan, R.; Kathalikkattil, A. C.; Kim, D. W.; Park, D.W. Rapid, Microwave-Assisted Synthesis of Cubic, Three-Dimensional, Highly Porous MOF-205 for Room Temperature CO2 Fixation via Cyclic Carbonate Synthesis. ACS Appl. Mater. Interfaces 2016, 8, 33723−33731. (59) Ma, D.; Li, B.; Liu, K.; Zhang, X.; Zou, W.; Yang, Y.; Li, G.; Shi, Z.; Feng, S. Bifunctional MOF Heterogeneous Catalysts Based on the Synergy of Dual Functional Sites for Efficient Conversion of CO2 under Mild and Co-Catalyst Free Conditions. J. Mater. Chem. A 2015, 3, 23136−23142. (60) D’Elia, V.; Dong, H.; Rossini, A. J.; Widdifield, C. M.; Vummaleti, S. V. C.; Minenkov, Y.; Poater, A.; Abou-Hamad, E.; Pelletier, J. D. A.; Cavallo, L.; Emsley, L.; Basset, J.-M. Cooperative Effect of Monopodal Silica-Supported Niobium Complex Pairs Enhancing Catalytic Cyclic Carbonate Production. J. Am. Chem. Soc. 2015, 137, 7728−7739. (61) Rocha, J.; Carlos, L. D.; Paz, F. A. A.; Ananias, D. Luminescent Multifunctional Lanthanides-Based Metal−Organic Frameworks. Chem. Soc. Rev. 2011, 40, 926−940. (62) Wang, Y.; Yang, J.; Liu, Y.-Y.; Ma, J.-F. Controllable Syntheses of Porous Metal-Organic Frameworks: Encapsulation of LnIII Cations for Tunable Luminescence and Small Drug Molecules for Efficient Delivery. Chem.Eur. J. 2013, 19, 14591−14599. (63) Du, P.; Yang, Y.; Yang, J.; Liu, B.-K.; Ma, J.-F. Syntheses, Structures, Photoluminescence, Photocatalysis, and Photoelectronic Effects of 3D Mixed High-Connected Metal−Organic Frameworks Based on Octanuclear and Dodecanuclear Secondary Building Units. Dalton Trans. 2013, 42, 1567−1580. (64) Zhang, L.-P.; Ma, J.-F.; Yang, J.; Pang, Y.-Y.; Ma, J.-C. Series of 2D and 3D Coordination Polymers Based on 1,2,3,4-Benzenetetracarboxylate and N-Donor Ligands: Synthesis, Topological Structures, and Photoluminescent Properties. Inorg. Chem. 2010, 49, 1535− 1550.

(65) Bauer, C. A.; Timofeeva, T. V.; Settersten, T. B.; Patterson, B. D.; Liu, V. H.; Simmons, B. A.; Allendorf, M. D. Influence of Connectivity and Porosity on Ligand-Based Luminescence in Zinc Metal−Organic Frameworks. J. Am. Chem. Soc. 2007, 129, 7136−7144. (66) Wang, J.; Li, Y.; Patel, N. G.; Zhang, G.; Zhou, D.; Pang, Y. A Single Molecular Probe for Multi-Analyte (Cr3+, Al3+ and Fe3+) Detection in Aqueous Medium and Its Biological Application. Chem. Commun. 2014, 50, 12258−12261. (67) Thomas, S. W.; Joly, G. D.; Swager, T. M. Chemical Sensors Based on Amplifying Fluorescent Conjugated Polymers. Chem. Rev. 2007, 107, 1339−1386. (68) Zhang, H.-M.; Yang, J.; Kan, W.-Q.; Liu, Y.-Y.; Ma, J.-F. Photocatalytic Properties and Luminescent Sensing for Cr3+ Cations of Polyoxovanadates-Based Inorganic−Organic Hybrid Compounds with Multiple Lewis Basic Sites. Cryst. Growth Des. 2016, 16, 265−276. (69) You, Y.; Cho, S.; Nam, W. Cyclometalated Iridium(III) Complexes for Phosphorescence Sensing of Biological Metal Ions. Inorg. Chem. 2014, 53, 1804−1815.

I

DOI: 10.1021/acsami.7b14179 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX