A Water-Stable Anionic Metal–Organic Framework Constructed from

Feb 20, 2017 - Noticeably, JXNU-4 shows potential as a separation agent for the selective removal of propane and ethane from natural gas with high sel...
4 downloads 19 Views 2MB Size
Article pubs.acs.org/IC

A Water-Stable Anionic Metal−Organic Framework Constructed from Columnar Zinc-Adeninate Units for Highly Selective Light Hydrocarbon Separation and Efficient Separation of Organic Dyes Hui-Fang Ma, Qing-Yan Liu,* Yu-Ling Wang,* and Shun-Gao Yin College of Chemistry and Chemical Engineering, Key Laboratory of Functional Small Organic Molecule of Ministry of Education, Jiangxi Normal University, Nanchang, Jiangxi 330022, P. R. China S Supporting Information *

ABSTRACT: A metal−organic framework (MOF), {(Me2NH2)2[Zn6(μ4-O)(ad)4(BPDC)4]}n (JXNU-4; ad− = adeninate), with an anionic three-dimensional (3D) framework constructed from one-dimensional (1D) columnar [Zn6(ad)4(μ4-O)]n secondary building units (SBUs) and 4,4′-biphenyldicarboxylate (BPDC2−) ligand, was prepared. The anionic 3D framework has 1D square channels with an aperture of about 9.8 Å and exhibits a carboxylate-O-decorated pore environment. The microporous nature of JXNU-4 was established by the N2 adsorption data, which gives Langmuir and Brumauer−Emmett−Teller surface areas of 1800 and 1250 m2 g−1, respectively. Noticeably, JXNU-4 shows potential as a separation agent for the selective removal of propane and ethane from natural gas with high selectivities of 144 for C3H8/CH4 (5:95) and 14.6 for C2H6/CH4 (5:95), respectively. Most importantly, JXNU-4 shows an aqueous-phase adsorption of a positively charged ion of methylene blue selectively over a negatively charged ion of resorufin, which is pertinent to the anionic nature of the framework, and provides a size-exclusive sieving of methylene blue over other positively charged ions of Janus Green B and ethyl violet, which is relevant to its pore structure, enabling the efficient aqueous-phase separation of organic dyes.



INTRODUCTION Metal−organic frameworks (MOFs) are porous hybrid solidstate compounds1 that have attracted much research effort for applications in gas adsorption and separation,2 chemical catalysis,3 proton conduction,4 and sensing.5 Compared to other porous solid materials such as activated carbon atoms and zeolite materials, MOFs possess the merits of structural diversity and adjustable chemical functionality,6 which make them highly promising candidates of porous materials. From a practical point of view, the water stability of the frameworks is of the precondition for their diverse applications such as CO2 capture from the power-plant flue gas7 and removal of organic contaminants.8 Unfortunately, a large amount of the reported MOFs are water-instable except for some examples of imidazolate-based ZIFs,9 MILs based on the trivalent metal ions,10 zirconium(IV)-based MOFs,11 and so on.12 Metal azolate frameworks have been reported showing higher chemical stability due to the formation of strong M−N coordination bonds between the metal ion (M) and azolate N atom.13 Furthermore, some MOFs with special structural motifs also exhibit high water stability such as the series of MOF-74 compounds with one-dimensional (1D) rod-shaped metal carboxylate secondary building units (SBUs).14 © XXXX American Chemical Society

In this contribution, a MOF, {(Me2NH2)2[Zn6(μ4-O)(ad)4(BPDC)4]}n (JXNU-4; ad− = adeninate and BPDC2− = 4,4′-biphenyldicarboxylate), with an anionic three-dimensional (3D) framework constructed from 1D columnar [Zn6(ad)4(μ4O)]n SBUs was synthesized. JXNU-4 shows high stability in aqueous media with a wide pH range, a characteristic that is highly desirable for separation-related applications. The robust 1D columnar [Zn6(ad)4(μ4-O)]n SBU in JXNU-4 presumably accounts for its high water stability, which is similar to the 1D rods consisting of edge-sharing MO6 octahedra in MOF-74.14a The robustness of this structure suggests that the water-stable frameworks can be constructed from the 1D columnar [Zn6(ad)4(μ4-O)]n SBU platform. This anionic MOF with high surface area exhibits a high light hydrocarbon separation selectivity and an aqueous-phase selective separation of organic dyes. To the best of our knowledge, the utilization of MOFs for aqueous-phase adsorption and the removal of organic dyes are much less explored15 because most of them were checked in organic solvents because of the poor moisture stability of the frameworks.16 The synthesis, crystal structure, and selective Received: December 12, 2016

A

DOI: 10.1021/acs.inorgchem.6b03026 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Table 1. Crystallographic Data for JXNU-4a

separation properties of JXNU-4 are presented here. The relationship between the structure and physical properties of JXNU-4 is also discussed.



formula fw temp (K) cryst syst space group Z a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Dcalcd (g cm−3) μ (mm−1) no. of reflns collected no. of indep reflns no. of obsd reflns [I > 2σ(I)] F(000) R(int) R1 [I > 2σ(I)] wR2 [I > 2σ(I)] CCDC

EXPERIMENTAL SECTION

Physical Measurements. IR (KBr pellets) spectra were recorded in the 400−4000 cm−1 range using a PerkinElmer Spectrum One Fourier transform infrared spectrometer. Thermogravimetric analysis (TGA) was carried out on a PerkinElmer Diamond TG/DTA unit at a heating rate of 10 °C min−1 under a nitrogen atmosphere from 30 to 800 °C. Powder X-ray diffraction (PXRD) was performed on a Rigaku Miniflex Π powder diffractometer using Cu Kα radiation (λ = 1.5418 Å). Upon immersion of the crystalline samples in an aqueous solution for 3 days and in an aqueous solution with different pH values (the pH of an aqueous solution was manually adjusted from 4 to 11 by the addition of 0.1 M HCl or NaOH to the distilled H2O) for 24 h, the samples were collected by filtration before PXRD analysis. Gas sorption isotherms were measured using a Micromeritics ASAP2020 gas adsorption instrument up to 1 atm of gas pressure. The highly pure N2 (99.999%), CO2 (99.999%), CH4 (99.999%), C2H6 (99.9%), and C3H8 (99.9%) were used in the sorption experiments. A fresh JXNU-4 sample was washed with N,N-dimethylformamide (DMF) three times and exchanged with chloroform for 5 days to give the chloroformexchanged sample. The chloroform-exchanged sample was degassed at 125 °C for 26 h before gas sorption measurements. UV/vis adsorption spectra were measured at room temperature with a PerkinElmer Lambda 35 UV/vis spectrometer. In a typical dye adsorption experiment, 6 mg (3 μmol) of as-synthesized JXNU-4 was soaked into a 3 mL aqueous solution containing 3 μmol of dye in a 10 mL vial. The upper clear solution of the mixture was taken out for UV/vis absorbance measurement at 0, 1, 3, 6, 17, 24, 29, 47, and 60 h, respectively. After each measurement, the solution was poured back into the original vial to avoid loss of the sample. Chemicals. All chemicals were of reagent grade and were used as commercially obtained. Synthesis of {(Me2NH2)2[Zn6(μ4-O)(ad)4(BPDC)4]}n (JXNU-4). A mixture of Zn(NO3)3·6H2O (30 mg, 0.1 mmol), adenine (6.8 mg, 0.05 mmol), and 4,4′-biphenyldicarboxylic acid (12 mg, 0.05 mmol) in 5 mL of DMF and 1 mL of H2O was introduced into a 25 mL Parr Teflon-lined stainless steel vessel and heated at 120 °C for 3 days. Then the mixture was cooled naturally to form crystals. Crystalline product was filtered, washed with DMF, and dried at ambient temperature. Yield: 18.8 mg. IR spectrum (cm−1, KBr pellets): 3365 (s), 3200 (m), 1655 (m), 1606 (s), 1546 (w), 1469 (m), 1384 (s), 1281 (w), 1214 (m), 1175 (w), 1154 (m), 1103 (w), 1022 (w), 1005 (w), 941 (w), 845 (m), 797 (w), 672 (m), 772 (s), 741 (w), 686 (m), 640 (w), 580 (w), 536 (w), 487 (w), 442 (w). X-ray Crystallography. Single-crystal X-ray diffraction data for JXNU-4 were collected on a Rigaku Oxford SuperNova single-source diffractometer with an EOS detector and Mo Ka radiation (λ = 0.71073 Å). CrysAlisPro (Agilent Technologies) software was used for the collection of frames of data, indexing of the reflections, determination of the lattice constants, absorption correction, and data reduction.17 The structures were solved by direct methods and successive Fourier difference syntheses and refined by a full-matrix least-squares method on F2 (SHELXTL-2014).18 All non-H atoms were refined with anisotropic thermal parameters. Six C atoms and one carboxylate O atom of the BPDC ligand are disordered over two positions. H atoms bonded to C atoms were assigned to calculated positions. The highly disordered dimethylammonium cations could not be located in the difference Fourier map, but they are included in the formula. The unit cell includes a large region of disordered guest solvent molecules that could not be modeled, were treated by the SQUEEZE routine of PLATON,19 and were refined further using the data generated. The R1 values are defined as R1 = ∑||Fo| − |Fc||/∑|Fo| and wR2 = {∑[w(Fo2 − Fc2)2]/∑[w(Fo2)2]}1/2. The details of the crystal parameters, data collection, and refinement are summarized in Table 1, and the selected bond lengths are listed in Table S1.

C80H64N22O17Zn6 1997.75 293(2) tetragonal P4/nnc 4 25.8389(6) 25.8389(6) 21.5389(6) 90 90 90 14380.5(8) 0.923 1.032 81321 9034 5262 4056 0.0578 0.0716 0.1181 1522098

a R1 = ∑||Fo| − |Fc||/∑|Fo| and wR2 = {∑[w(Fo2 − Fc2)2]/ ∑[w(Fo2)2]}1/2.



RESULTS AND DISCUSSION Crystal Structure. The reaction of Zn(NO3)2, adenine, and 4,4′-biphenyldicarboxylic acid in a DMF/H2O solvent afforded JXNU-4, which crystallizes in tetragonal space group P4/nnc (Table 1). One and half ZnII ions, one μ4-O2− (O5) with an occupation of 0.25, one ad−, and one BPDC2− are in the asymmetric unit (Figure 1a). The two crystallographically independent ZnII ions both display tetrahedral coordination geometry. As depicted in Figure 1a, Zn1 atom located at the 2fold axis is coordinated by two imidazolate N atoms of two ad− ligands and two carboxylate O atoms from two BPDC2− ligands, while Zn2 atom is bonded to one imidazolate N atom and one pyrimidine N atom from two ad− ligands and two O atoms from a BPDC2− ligand and a μ4-O. The Zn−O bond distances vary from 1.917(3) to 1.9723(4) Å, and Zn−N bond distances range from 2.020(3) to 2.041(3) Å (Table S1). The ad− ligand binds three Zn ions through its one pyrimidine N atom and two imidazolate N atoms (Scheme S1). As shown in Figure 1b, four μ3-ad− ligands bridge four Zn2 atoms and two Zn1 atoms to generate a Zn6(ad)4 cage, which is different from the less open Zn8(ad)4 octahedral cage, wherein the ad− ligand shows a μ4-coordination mode with its two imidazolate N atoms and two pyrimidine N atoms in bio-MOF-1.20 Analysis of the Zn8(ad)4 unit revealed its close relationship with the Zn6(ad)4 unit but with two ZnII ions incorporated. The two additional ZnII ions each bridge one pair of the uncoordinated pyrimidine N atoms in the Zn6(ad)4 unit (Figure 1b), giving the Zn8(ad)4 unit. Thus, new cages can be obtained by controlling the coordination mode of ad−, indicating the diverse coordination adaptivity of the ad− ligands. The adjacent Zn6(ad)4 cages are linked by μ4-O through the Zn−O bonds to give a 1D infinite columnar [Zn6(ad)4(μ4-O)]n SBU featuring O-centered tetrahedral [Zn4(μ4-O)] units (Figure 1c). Within the 1D columnar [Zn6(ad)4(μ4-O)]n SBU, each Zn6(ad)4 cage is surrounded by eight BPDC2− ligands with two B

DOI: 10.1021/acs.inorgchem.6b03026 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. (a) Coordination environments of ZnII ions. (b) Each Zn6(ad)4 cage surrounded by eight carboxylate ligands. (c) Adjacent [Zn6(ad)4(μ4O)] cages sharing the μ4-O5 atoms to give a 1D columnar SBU.

12 bidentate BPDC2− ligands. Because of the essential differences of the connectivity patterns of the two kinds of 1D columnar zinc adeninate SBUs, one kind of 1D channel is observed in JXNU-4, while bio-MOF-1 has two types of 1D channels with different pore apertures along the c axis.20,21 The resultant 3D framework of the JXNU-4 overall charge is balanced by dimethylammonium cations (Me2NH2+) generated in situ from the decomposition of DMF solvent molecules. As shown in Figure 2, the uncoordinated carboxylate O atoms are aligned on the pore surface. The presence of the double walls and the large columnar [Zn6(ad)4(μ4-O)]n SBUs in this structure prevents interpenetration of the framework. The accessible volume of the solvents and the guest Me2NH2+ cations is 55.8%, as calculated by PLATON.19 The results of thermal analysis show that the chloroform-exchanged JXNU-4 displays the first weight loss (21.97%) from room temperature to 130 °C (Figure S1), which is attributed to the loss of guest solvents. A small gradual mass loss of 2.62% continues subsequently from 140 to 210 °C, which corresponds to the release of Me2NH2+ cations. A fresh JXNU-4 sample was washed with DMF and exchanged with chloroform for 5 days and thermally activated at 125 °C under a dynamic vacuum before gas sorption measurements. The TGA curve of the desolvated JXNU-4 shows a slight weight loss between 70 and 110 °C, which corresponds to the removal of the adsorbed H2O from air (Figure S1). Then a platform is observed between 110 and 360 °C. PXRD of the activated sample is well matched with the as-synthesized one, indicating that the 3D framework is intact after activation (Figure S2). Furthermore, the PXRD patterns of the guest-free JXNU-4 and JXNU-4 after immersion in H2O are consistent with the experimental pattern of the as-synthesized sample (Figure S2). Interestingly, this material can keep its crystalline framework in a wide pH range

monodentate carboxylate groups (Figure 1b and Scheme S1). The BPDC2− ligands link the 1D columnar [Zn6(ad)4(μ4-O)]n SBUs to form an anionic 3D framework of [Zn6(μ4O)(ad)4(BPDC)4]n2n− (Figure 2). The 3D framework with double walls possesses 1D square channels running along the c axis with dimensions of 9.8 × 9.8 Å2, wherein the DMF and H2O molecules are housed, while each Zn8(ad)4 cage of the 1D columnar [Zn8(ad)4(μ4-O)]n SBU in bio-MOF-1 is linked by

Figure 2. Each 1D columnar [Zn6(ad)4(μ4-O)]n SBU linking four [Zn6(ad)4(μ4-O)]n SBUs through BPDC2− ligands to generate a 3D framework showing square channels. C

DOI: 10.1021/acs.inorgchem.6b03026 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. (a) N2 sorption isotherm at 77 K (inset: pore-size distribution analyzed by the NLDFT method). (b) CO2 sorption isotherms at 273 and 293 K for JXNU-4 (inset: Qst as a function of the CO2 uptake). Solid symbols: adsorption. Open symbols: desorption.

Figure 4. (a) CH4, C2H6, and C3H8 sorption isotherms at 273 K (solid symbols, adsorption; open symbols, desorption). (b) Separation factors in binary gas mixtures for JXNU-4.

of 4−11 (Figure S2). To evaluate the permanent porosity of JXNU-4, the N2 adsorption isotherm was measured at 77 K. The N2 sorption of JXNU-4 exhibits a typical reversible type I isotherm and a very sharp uptake at P/P0 < 0.02 with a saturated adsorption amount of 410 cm3 (STP) g−1 (Figure 3a), corroborating the microporous nature of JXNU-4. The Langmuir and Brunauer−Emmett−Teller surface areas calculated from the N2 adsorption data are 1800 and 1250 m2 g−1, respectively. The total pore volume is 0.63 cm3 g−1 calculated from the N2 adsorption isotherm, which is well consistent with the value of 0.67 cm3 g−1 estimated from the crystal data. Analysis of the N2 adsorption data using the nonlocal density functional theory (NLDFT) model confirms a narrow distribution of micropores around 10 Å (Figure 3a, inset). In order to assess the structure−property relationship for JXNU-4, gas adsorption measurements were further undertaken for CO2. JXNU-4 takes up 78.2 and 53.3 cm3 g−1 of CO2 at 273 and 293 K (1 atm), respectively (Figure 3b). The CO2 uptake of 3.49 mmol g−1 for JXNU-4 at 273 K is comparable to those of wellknown anionic frameworks such as bio-MOF-1 (3.41 mmol g−1)21 and CPM-1 (3.15 mmol g−1).22 The isosteric heat of adsorption Qst for CO2 was calculated based on the virial method from fits of its adsorption isotherms at 273 and 293 K,23 which gives a Qst value of 27.84 kJ mol−1 at zero loading (Figure 3b, inset). The Qst value is smaller than that of bioMOF-1 (35 kJ mol−1)21 but similar to those of CPM-13 (28.2 kJ mol−1),24 NOTT-140 (25 kJ mol−1),25 Tripp-1-Co (25.6 kJ mol−1),26 and InOF-1 (29 kJ mol−1).27

Encouraged by the CO2 uptake of JXNU-4, the adsorption of C1−C3 paraffins was measured at 273 and 293 K (Figures 4a and S3). As the experiments revealed, JXNU-4 shows comparatively high adsorption amounts of C3H8 (152.1 cm3 g−1) and C2H6 (120.3 cm3 g−1) at 273 K and 1 atm (Figure 4a). The C3H8 and C2H6 uptakes are much higher than that of the zinc(II) compound constructed from the ad − and 4pyrazolecarboxylate ligands (C3H8, 88.1 cm3 g−1; C2H6, 92.8 cm3 g−1).28 Interestingly, the C3H8 adsorption isotherm was much steeper at low pressure than those for C2H6 and CH4, indicative of the high affinity of JXNU-4 for the larger and highly polarizable molecule of C3H8. In contrast, JXNU-4 was found to adsorb much less CH4 (20.2 cm3 g−1) in the same conditions, suggesting that JXNU-4 has potential as a separation agent for the selective removal of propane and ethane from natural gas. The near-linear adsorption profile for CH4 is indicative of its low affinity for the charged framework of JXNU-4, as expected from its relatively low polarizability. Analysis of the adsorption isotherm data using ideal adsorption solution theory29 confirmed the high gas separation selectivity; namely, the selectivity was ca. 144 for C3H8/CH4 (5:95) and ca. 14.6 for C2H6/CH4 (5:95) at 1 bar (Figure 4b). The C3H8 molecule has a large size and a dipole moment of 0.28 × 10−30 C m,30 while the CH4 molecule is a nondipolar molecule. Moreover, the polarizability of C3H8 [(62.9−63.7) × 10−25 cm3] is 2 times higher than that of CH4 (25.93 × 10−25 cm3),30 which induces stronger interactions between the host framework and C3H8. The C3H8/CH4 selectivity is much higher than D

DOI: 10.1021/acs.inorgchem.6b03026 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 5. UV/vis spectra of solutions of MEB (a) and RS (b) dyes in H2O in the presence of JXNU-4 with time.

process, as evidenced by PXRD studies (Figure S2). In addition, the JB (19.2 × 9.3 Å2) and EV (15.1 × 12.7 Å2) cationic dyes have larger sizes than that of the aperture of the 1D open channels. These two large dye molecules cannot diffuse into the MOF channels, providing a size-exclusive sieving of MEB over JB and EV. Therefore, the material of JXNU-4 is preferred for absorbing the dye MEB, highlighting the efficient separation of organic dyes.

those of some reported MOFs under similar conditions, such as pek-MOF-1 (32),31 gea-MOF-1 (126),32 and Y-ftw-MOF-2 (50).33 Therefore, JXNU-4 can be employed as a hydrocarbon separation agent for the purification of CH4 from condensable hydrocarbons such as C2H6 and C3H8. Additionally, for a comparison with those of other reported compounds, the selectivity of the mole ratio of 50:50 for CO2/CH4 was also calculated (Figure S4), giving a value of 6.5 at 1 bar, which is higher than of those of famous reported MOFs under similar conditions, such as ZIF-100 (5.9),34 MOF-177 (4.4),35 and CNT@Cu3(BTC)2 (5.6).36 As mentioned above, JXNU-4 has an anionic 3D framework of [Zn6(μ4-O)(ad)4(BPDC)4]n2n− with 1D open channels. We were inspired by the structural features that the uncoordinated carboxylate O atoms are lined on the surfaces of the channels (Figure 2), indicating that the present compound with a charged pore surface can potentially be capable of absorbing positively charged molecules selectively over negative ones. To probe the charge of the pore surface within JXNU-4, three cationic organic dyes [methylene blue (MEB; λmax = 665 nm), Janus Green B (JB; λmax = 656 nm), ethyl violet (EV; λmax = 595 nm) and an anionic organic dye [resorufin sodium salt (RS; λmax = 593 nm)] were chosen as diagnostic agents. The asprepared crystalline samples (6 mg) of JXNU-4 were immersed in H2O (3 mL) solutions containing each of the dyes with an initial dye concentration of 1 mM. Then the UV/vis absorbance of the supernatant of the crystal-soaked solutions was monitored at different time intervals. As shown in Figure 5, only the MEB cationic dye was absorbed by JXNU-4, while the JB and EV cationic dyes and RS anionic dye are denied access to the MOF interior (Figures S5−S7). The MEB cationic dye was almost fully absorbed by JXNU-4 within 60 h at room temperature (Figures 5a and S7a). The dye uptake was 0.16 wt %, as determined by UV/vis absorption spectroscopy (Figure S8), which corresponds to 1 molecule of MEB per formula unit of JXNU-4. The TGA trace of the MEB-incorporated JXNU-4 exhibits the first weight loss of 25.88% between 40 and 210 °C, which is attributed to the release of MEB and solvent molecules. Then collapse of the 3D framework occurred at 360 °C (Figure S1). The anionic dye RS (cross dimensions of ca. 11.1 × 4.3 Å2) has a smaller size than that of the cationic dye MEB (13.8 × 5.8 Å2), excluding the possibility of size-based adsorption selectivity. Thus, selective adsorption of the cationic dye MEB over the anionic dye RS can be attributed to the anionic nature of the 3D framework. The crystalline integrity of the framework is maintained throughout the absorption



CONCLUSIONS In conclusion, JXNU-4 with an anionic 3D framework based on the columnar [Zn6(ad)4(μ4-O)]n SBUs has been synthesized and characterized. JXNU-4 shows high stability in aqueous media with a wide pH range, which indicates that the waterstable frameworks can be constructed from the 1D columnar [Zn6(ad)4(μ4-O)]n SBU platform. The 3D framework featuring 1D square channels exhibits a high affinity for larger and highly polarizable molecules of C3H8 and C2H6, revealing its potential for application in gas separation. Moreover, the JXNU-4 framework shows a preference for absorbing the MEB dye over other organic dyes such as JB, EV, and RS in aqueous solution, highlighting the efficient aqueous-phase separation of organic dyes.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.6b03026.



Table of bond lengths, PXRD patterns, TGA curves, adsorption data, and UV/vis spectra (PDF) X-ray structure data in CIF format (CIF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (Q.-Y.L.). Fax: +86-79188336372. *E-mail: [email protected] (Y.-L.W.). Fax: +86-79188336372. ORCID

Qing-Yan Liu: 0000-0003-1991-792X Notes

The authors declare no competing financial interest. E

DOI: 10.1021/acs.inorgchem.6b03026 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



mesostructured MIL-53(Al). J. Colloid Interface Sci. 2013, 405, 157− 163. (9) Park, K. S.; Ni, Z.; Côté, A. P.; Choi, J. Y.; Huang, R.; UribeRomo, F. J.; Chae, H. K.; O’Keeffe, M.; Yaghi, O. M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 10186−10191. (10) Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.; Margiolaki, I. A Chromium Terephthalate-Based Solid with Unusually Large Pore Volumes and Surface Area. Science 2005, 309, 2040−2042. (11) (a) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K. P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability. J. Am. Chem. Soc. 2008, 130, 13850−13851. (b) Wang, B.; Lv, X.-L.; Feng, D.; Xie, L.-H.; Zhang, J.; Li, M.; Xie, Y.; Li, J.; Zhou, H.-C. Highly Stable Zr(IV)-Based Metal−Organic Frameworks for the Detection and Removal of Antibiotics and Organic Explosives in Water. J. Am. Chem. Soc. 2016, 138, 6204−6216. (12) Burtch, N. C.; Jasuja, H.; Walton, K. S. Water Stability and Adsorption in Metal−Organic Frameworks. Chem. Rev. 2014, 114, 10575−10612. (13) (a) Colombo, V.; Galli, S.; Choi, H. J.; Han, G. D.; Maspero, A.; Palmisano, G.; Masciocchi, N.; Long, J. R. High thermal and chemical stability in pyrazolate-bridged metal−organic frameworks with exposed metal sites. Chem. Sci. 2011, 2, 1311−1319. (b) Zhang, J.-P.; Zhang, Y.-B.; Lin, J.-B.; Chen, X.-M. Metal Azolate Frameworks: From Crystal Engineering to Functional Materials. Chem. Rev. 2012, 112, 1001− 1033. (14) (a) Rosi, N. L.; Kim, J.; Eddaoudi, M.; Chen, B.; O’Keeffe, M.; Yaghi, O. M. Rod Packings and Metal−Organic Frameworks Constructed from Rod-Shaped Secondary Building Units. J. Am. Chem. Soc. 2005, 127, 1504−1518. (b) Dietzel, P. D. C.; Panella, B.; Hirscher, M.; Blom, R.; Fjellvag, H. Hydrogen adsorption in a nickel based coordination polymer with open metal sites in the cylindrical cavities of the desolvated framework. Chem. Commun. 2006, 959−961. (c) Dietzel, P. D. C.; Johnsen, R. E.; Blom, R.; Fjellvag, H. Structural Changes and Coordinatively Unsaturated Metal Atoms on Dehydration of Honeycomb Analogous Microporous Metal−Organic Frameworks. Chem. - Eur. J. 2008, 14, 2389−2397. (d) Dietzel, P. D. C.; Johnsen, R. E.; Fjellvag, H.; Bordiga, S.; Groppo, E.; Chavan, S.; Blom, R. Adsorption properties and structure of CO2 adsorbed on open coordination sites of metal−organic framework Ni2(dhtp) from gas adsorption, IR spectroscopy and X-ray diffraction. Chem. Commun. 2008, 5125−5127. (15) (a) Tong, M.; Liu, D.; Yang, Q.; Devautour-Vinot, S.; Maurin, G.; Zhong, C. Influence of framework metal ions on the dye capture behavior of MIL-100 (Fe, Cr) MOF type solids. J. Mater. Chem. A 2013, 1, 8534−8537. (b) Wang, B.; Lv, X.-L.; Feng, D.; Xie, L.-H.; Zhang, J.; Li, M.; Xie, Y.; Li, J.-R.; Zhou, H.-C. Highly Stable Zr(IV)Based Metal−Organic Frameworks for the Detection and Removal of Antibiotics and Organic Explosives in Water. J. Am. Chem. Soc. 2016, 138, 6204−6216. (16) (a) Li, P.; Vermeulen, N. A.; Gong, X.; Malliakas, C. D.; Stoddart, J. F.; Hupp, J. T.; Farha, O. K. Design and Synthesis of a Water-Stable Anionic Uranium-Based Metal−Organic Framework (MOF) with Ultra Large Pores. Angew. Chem., Int. Ed. 2016, 55, 10358−10362. (b) Dong, M.-J.; Zhao, M.; Ou, S.; Zou, C.; Wu, C.-D. A Luminescent Dye@MOF Platform: Emission Fingerprint Relationships of Volatile Organic Molecules. Angew. Chem., Int. Ed. 2014, 53, 1575−1579. (c) Masoomi, M. Y.; Morsali, A.; Junk, P. C. Rapid mechanochemical synthesis of two new Cd(II)-based metal−organic frameworks with high removal efficiency of Congo red. CrystEngComm 2015, 17, 686−692. (d) Masoomi, M. Y.; Bagheri, M.; Morsali, A. High efficiency of mechanosynthesized Zn-based metal−organic frameworks in photodegradation of congo red under UV and visible light. RSC Adv. 2016, 6, 13272−13277. (e) Masoomi, M. Y.; Bagheri, M.; Morsali, A. High efficiency of mechanosynthesized Zn-based metal-organic frameworks in photodegradation of congo red under UV and visible light. RSC Adv. 2016, 6, 13272−13277. (f) Fang, Q.-R.;

ACKNOWLEDGMENTS This work was supported by the NNSF of China (Grant 21361011, 21661014, and 21561015), the Young Scientist Training Project of Jiangxi Province (Grant 20153BCB23017), and the NSF of Jiangxi (Grant 20151BAB203002).



REFERENCES

(1) (a) Férey, G. Hybrid porous solids: past, present, future. Chem. Soc. Rev. 2008, 37, 191−214. (b) Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Introduction to Metal−Organic Frameworks. Chem. Rev. 2012, 112, 673−674. (c) Horike, S.; Shimomura, S.; Kitagawa, S. Soft porous crystals. Nat. Chem. 2009, 1, 695−704. (d) Yaghi, O. M.; O’Keeffe, M.; Ockwig, N. W.; Chae, H. K.; Eddaoudi, M.; Kim, J. Reticular synthesis and the design of new materials. Nature 2003, 423, 705−714. (e) Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent Functional Metal− Organic Frameworks. Chem. Rev. 2012, 112, 1126−1162. (2) (a) Li, J.-R.; Sculley, J.; Zhou, H.-C. Metal−Organic Frameworks for Separations. Chem. Rev. 2012, 112, 869−932. (b) He, Y.; Zhou, W.; Qian, G.; Chen, B. Methane storage in metal−organic frameworks. Chem. Soc. Rev. 2014, 43, 5657−5678. (c) Murray, L. J.; Dincă, M.; Long, J. R. Hydrogen storage in metal−organic frameworks. Chem. Soc. Rev. 2009, 38, 1294−1314. (3) (a) Ma, L.; Abney, C.; Lin, W. Enantioselective catalysis with homochiral metal−organic frameworks. Chem. Soc. Rev. 2009, 38, 1248−1256. (b) Liu, J.; Chen, L.; Cui, H.; Zhang, J.; Zhang, L.; Su, C.Y. Applications of metal−organic frameworks in heterogeneous supramolecular catalysis. Chem. Soc. Rev. 2014, 43, 6011−6061. (c) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Metal−organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450−1459. (d) Yoon, M.; Srirambalaji, R.; Kim, K. Homochiral Metal−Organic Frameworks for Asymmetric Heterogeneous Catalysis. Chem. Rev. 2012, 112, 1196−1231. (4) (a) Sadakiyo, M.; Yamada, T.; Kitagawa, H. Rational Designs for Highly Proton-Conductive Metal-Organic Frameworks. J. Am. Chem. Soc. 2009, 131, 9906−9907. (b) Taylor, J. M.; Dawson, K. W.; Shimizu, G. K. H. A Water-Stable Metal−Organic Framework with Highly Acidic Pores for Proton-Conducting Applications. J. Am. Chem. Soc. 2013, 135, 1193−1196. (c) Zhou, L.-J.; Deng, W.-H.; Wang, Y.-L.; Xu, G.; Yin, S.-G.; Liu, Q.-Y. Lanthanide-Potassium-Biphenyl-3,3′disulfonyl-4,4′-dicarboxylate Frameworks: Gas Sorption, Proton Conductivity, and Luminescent Sensing of Metal Ions. Inorg. Chem. 2016, 55, 6271−6277. (5) (a) Wong, K.-L.; Law, G.-L.; Yang, Y.-Y.; Wong, W.-T. A Highly Porous Luminescent Terbium−Organic Framework for Reversible Anion Sensing. Adv. Mater. 2006, 18, 1051−1054. (b) Wu, Z.-L.; Dong, J.; Ni, W.-Y.; Zhang, B.-W.; Cui, J.-Z.; Zhao, B. Unique Chiral Interpenetrating d−f Heterometallic MOFs as Luminescent Sensors. Inorg. Chem. 2015, 54, 5266−5272. (c) Dong, X.-Y.; Wang, R.; Wang, J.-Z.; Zang, S.-Q.; Mak, T. C. W. Highly selective Fe3+ sensing and proton conduction in a water-stable sulfonate−carboxylate Tb− organic-framework. J. Mater. Chem. A 2015, 3, 641−647. (6) (a) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. (b) McDonald, T. M.; D’Alessandro, D. M.; Krishna, R.; Long, J. R. Enhanced carbon dioxide capture upon incorporation of N,N′-dimethylethylenediamine in the metal−organic framework CuBTTri. Chem. Sci. 2011, 2, 2022−2028. (7) 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. (8) (a) Jhung, S. H.; Lee, J.-H.; Yoon, J. W.; Serre, C.; Férey, G.; Chang, J.-S. Microwave Synthesis of Chromium Terephthalate MIL101 and Its Benzene Sorption Ability. Adv. Mater. 2007, 19, 121−124. (b) Cychosz, K. A.; Matzger, A. J. Water Stability of Microporous Coordination Polymers and the Adsorption of Pharmaceuticals from Water. Langmuir 2010, 26, 17198−17202. (c) Zhou, M.; Wu, Y.-N.; Qiao, J.; Zhang, J.; McDonald, A.; Li, G.; Li, F. The removal of bisphenol A from aqueous − solutions by MIL-53(Al) and F

DOI: 10.1021/acs.inorgchem.6b03026 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Zhu, G.-S.; Jin, Z.; Ji, Y.-Y.; Ye, J.-W.; Xue, M.; Yang, H.; Wang, Y.; Qiu, S.-L. Mesoporous Metal−Organic Framework with Rare etb Topology for Hydrogen Storage and Dye Assembly. Angew. Chem., Int. Ed. 2007, 46, 6638−6642. (g) He, H.; Ma, E.; Cui, Y.; Yu, J.; Yang, Y.; Song, T.; Wu, C.-D.; Chen, X.; Chen, B.; Qian, G. Polarized threephoton-pumped laser in a single MOF microcrystal. Nat. Commun. 2016, 7, 11087. (h) Ma, M.; Gross, A.; Zacher, D.; Pinto, A.; Noei, H.; Wang, Y.; Fischer, R. A.; Metzler-Nolte, N. Use of confocal fluorescence microscopy to compare different methods of modifying metal−organic framework (MOF) crystals with dyes. CrystEngComm 2011, 13, 2828−2832. (17) CrysAlisPro; Rigaku Oxford Diffraction: 2015. (18) Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (19) Spek, A. L. PLATON: A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2001. (20) An, J.; Geib, S. J.; Rosi, N. L. Cation-Triggered Drug Release from a Porous Zinc−Adeninate Metal−Organic Framework. J. Am. Chem. Soc. 2009, 131, 8376−8377. (21) An, J.; Rosi, N. L. Tuning MOF CO2 Adsorption Properties via Cation Exchange. J. Am. Chem. Soc. 2010, 132, 5578−5579. (22) Chen, S.; Zhang, J.; Wu, T.; Feng, P.; Bu, X. Multiroute Synthesis of Porous Anionic Frameworks and Size-Tunable Extraframework Organic Cation-Controlled Gas Sorption Properties. J. Am. Chem. Soc. 2009, 131, 16027−16029. (23) Rowsell, J. L. C.; Yaghi, O. M. Effects of Functionalization, Catenation, and Variation of the Metal Oxide and Organic Linking Units on the Low-Pressure Hydrogen Adsorption Properties of Metal−Organic Frameworks. J. Am. Chem. Soc. 2006, 128, 1304−1315. (24) Zhai, Q.-G.; Lin, Q.; Wu, T.; Wang, L.; Zheng, S.-T.; Bu, X.; Feng, P. High CO2 and H2 Uptake in an Anionic Porous Framework with Amino-Decorated Polyhedral Cages. Chem. Mater. 2012, 24, 2624−2626. (25) Tan, C.; Yang, S.; Champness, N. R.; Lin, X.; Blake, A. J.; Lewis, W.; Schröder, M. High capacity gas storage by a 4,8-connected metal− organic polyhedral framework. Chem. Commun. 2011, 47, 4487−4489. (26) Chen, K. J.; Perry, J. J., IV; Scott, H. S.; Yang, Q. Y.; Zaworotko, M. J. Double-walled pyr topology networks from a novel fluoridebridged heptanuclear metal cluster. Chem. Sci. 2015, 6, 4784−4789. (27) Qian, J.; Jiang, F.; Yuan, D.; Wu, M.; Zhang, S.; Zhang, L.; Hong, M. Highly selective carbon dioxide adsorption in a water-stable indium−organic framework material. Chem. Commun. 2012, 48, 9696−9698. (28) Fu, H.-R.; Zhang, J. Structural Transformation and Hysteretic Sorption of Light Hydrocarbons in a Flexible Zn−Pyrazole−Adenine Framework. Chem. - Eur. J. 2015, 21, 5700−5703. (29) Myers, A. L.; Prausnitz, J. M. Thermodynamics of mixed-gas adsorption. AIChE J. 1965, 11, 121−127. (30) (a) Gao, S.; Morris, C. G.; Lu, Z.; Yan, Y.; Godfrey, H. G. W.; Murray, C.; Tang, C. C.; Thomas, K. M.; Yang, S.; Schröder, M. Selective Hysteretic Sorption of Light Hydrocarbons in a Flexible Metal−Organic Framework Material. Chem. Mater. 2016, 28, 2331− 2340. (b) Drago, R. S.; Webster, C. E.; McGilvray, J. M. A multipleprocess equilibrium analysis of silica gel and HZSM-5. J. Am. Chem. Soc. 1998, 120, 538−547. (31) Alezi, D.; Peedikakkal, A. M. P.; Weseliński, Ł. J.; Guillerm, V.; Belmabkhout, Y.; Cairns, A. J.; Chen, Z.; Wojtas, Ł.; Eddaoudi, M. Quest for Highly Connected Metal−Organic Framework Platforms: Rare-Earth Polynuclear Clusters Versatility Meets Net Topology Needs. J. Am. Chem. Soc. 2015, 137, 5421−5430. (32) Guillerm, V.; Weseliński, Ł. J.; Belmabkhout, Y.; Cairns, A. J.; D’Elia, V.; Wojtas, L.; Adil, K.; Eddaoudi, M. Discovery and introduction of a (3,18)-connected net as an ideal blueprint for the design of metal−organic frameworks. Nat. Chem. 2014, 6, 673−680. (33) Luebke, R.; Belmabkhout, Y.; Weseliński, Ł. J.; Cairns, A. J.; Alkordi, M.; Norton, G.; Wojtas, Ł.; Adil, K.; Eddaoudi, M. Versatile rare earth hexanuclear clusters for the design and synthesis of highlyconnected ftw-MOFs. Chem. Sci. 2015, 6, 4095−4102.

(34) Wang, B.; Côté, A. P.; Furukawa, H.; O’Keeffe, M.; Yaghi, O. M. Colossal cages in zeolitic imidazolate frameworks as selective carbon dioxide reservoirs. Nature 2008, 453, 207−211. (35) Saha, D.; Bao, Z.; Jia, F.; Deng, S. Adsorption of CO2, CH4, N2O, and N2 on MOF-5, MOF-177, and Zeolite 5A. Environ. Sci. Technol. 2010, 44, 1820−1826. (36) Xiang, Z.; Peng, X.; Cheng, X.; Li, X. J.; Cao, D. P. CNT@ Cu3(BTC)2 and Metal−Organic Frameworks for Separation of CO2/ CH4 Mixture. J. Phys. Chem. C 2011, 115, 19864−19871.

G

DOI: 10.1021/acs.inorgchem.6b03026 Inorg. Chem. XXXX, XXX, XXX−XXX