Metal–Organic Frameworks with Nanoporous Channels and Vacant

Apr 28, 2014 - Two Copper(II) Metal−Organic Frameworks with Nanoporous. Channels and Vacant Coordination Sites. Yumei Huang,. †,#. Bingguang Zhang...
0 downloads 0 Views 407KB Size
Article pubs.acs.org/crystal

Two Copper(II) Metal−Organic Frameworks with Nanoporous Channels and Vacant Coordination Sites Yumei Huang,†,# Bingguang Zhang,‡,# Jingui Duan,∥ Wenlong Liu,§ Xiaofang Zheng,† Lili Wen,*,† Xiaohuan Ke,† and Dongfeng Li*,† †

Key Laboratory of Pesticide & Chemical Biology of Ministry of Education, College of Chemistry, Central China Normal University, Wuhan, 430079, P. R. China ‡ Key Laboratory of Catalysis and Materials Science of the State Ethnic Affairs Commission & Ministry of Education, South-Central University for Nationalities, Wuhan 430074, P. R. China § College of Chemistry and Chemical Engineering, Yangzhou University, Yangzhou, 225002, P. R. China ∥ State Key Laboratory of Materials-Oriented Chemical Engineering, Nanjing University of Technology, Nanjing, 210009, P. R. China S Supporting Information *

ABSTRACT: Two three-dimensional microporous compounds, Cu6(BTTC)4(H2O)6·xS (1) and [(CH3)2NH2]3[(Cu4Cl)3(BTTC)8]·yS (2, H3BTTC = benzo-(1,2;3,4;5,6)tris (thiophene-2′-carboxylic acid), S represents noncoordinated solvent molecules), have been solvothermally synthesized and characterized, both of which are based upon truncated octahedron subunits and contain uniform nanosized cavities but exhibit different topological frameworks. Complex 1 demonstrates high adsorption enthalpies for H2 and CO2 gas molecules, stemming principally from the presence of the exposed metal Cu(II) sites on the pore surface. In particular, activated complex 1 shows high efficiency for the separation of energy-correlated molecules, including CO2 over N2 and CH4 under ambient conditions.



INTRODUCTION Metal−organic frameworks (MOFs) assembled from metal ions and multitopic organic moieties are a new type of crystalline porous material, which have attracted much interest recently.1 Owing to their uniform but tunable cavities, permanent porosity, and tailorable chemistry, MOFs have being considered as promising candidates in extensive applications, ranging from strategic storage and separation of gases to catalysis, chemical sensing, and drug delivery.2 Among many ongoing endeavors to develop high-performance MOF materials, one of the effective approaches to improve their affinity and gas selectivity is to build structures bearing vacant coordination sites on the pore surfaces.3 The vacant coordination sites serving as charge-dense binding sites could interact more strongly with the absorbed gas molecule, and moreover, favor selectively separate gas mixtures due to their different polarizability and/or quadrupole moment. The triangular carboxylate ligand benzo-(1,2;3,4;5,6)-tris(thiophene-2′-carboxylic acid) (H3BTTC, Scheme 1) was chosen as the organic linker to construct MOFs with high porosity due to its relatively larger size and multiple binding modes. Interestingly, Zhou recently has successfully applied H3BTTC to connect Cd2+ and Fe2+, respectively; the two resulting well-characterized compounds show hierarchically assembled micro- and mesopores and exhibit fascinating gas adsorption behavior and versatile frameworks.4 In our study, © 2014 American Chemical Society

Scheme 1. Benzo-(1,2;3,4;5,6)-tris(thiophene-2′-carboxylic acid) (H3BTTC)

the copper(II) ion was deliberately selected as the metal center based on the consideration that it may facilitate the presence of vacant coordination sites in a Cu2-paddlewheel structural motif. Herein, with H3BTTC with Cu2+, we successfully prepared two three-dimensional (3D) microporous MOFs, Cu6(BTTC)4(H2O)6·xS (1) and [(CH3)2NH2]3[(Cu4Cl)3(BTTC)8]·yS (2, S represents noncoordinated solvent molecules), both of which are based upon truncated octahedron subunits and contain uniform nanosized cavities but display different topological frameworks: tbo for 1 and the for 2. The former demonstrates high adsorption enthalpies for H2 and CO2 gas molecules, stemming principally from the presence of the exposed metal Cu(II) sites on the pore surface. In particular, activated 1 could Received: January 31, 2014 Revised: April 19, 2014 Published: April 28, 2014 2866

dx.doi.org/10.1021/cg500175k | Cryst. Growth Des. 2014, 14, 2866−2872

Crystal Growth & Design

Article

refinements were based on the new data. Crystallographic data have been deposited with the Cambridge Crystallographic Data Centre: CCDC 975087−975088. Crystallographic data and refinement information for complex 1 are provided in Table 1.

selectively adsorb energy-correlated gas molecules efficiently, including CO2 over N2 and CH4 under ambient conditions.



EXPERIMENTAL SECTION

Materials and Instrumentation. All purchased starting materials were used without further purification. Ligand H3BTTC was synthesized according to the reported literature.5 Elemental C, H, and N microanalyses were carried out with a Perkin-Elmer 240 elemental analyzer. IR spectra were recorded on KBr discs on a Bruker Vector 22 spectrophotometer in the 4000−400 cm−1 region. Thermogravimetric analyses were performed on a simultaneous SDT 2960 thermal analyzer under flowing N2 with a heating rate of 10 °C/ min between ambient temperature and 700 °C. The powder X-ray diffraction (PXRD) data were collected on a Siemens D5005 diffractometer with Cu Kα radiation (λ = 1.5418 Å) over the 2θ range of 5−30° at room temperature. Sample Activation. Solvent-exchanged complex 1 was obtained by soaking the as-synthesized sample in methanol for 3 days with methanol refreshing every 8 h. The completely desolvated 1 about 100 mg was afforded by heating the solvent-exchanged bulk at 393 K under a vacuum overnight. A color change from pale green to deep green was observed, which indicates the presence of unsaturated CuII sites. The typical color change upon activation was also detected for other MOFs based on Cu2-paddlewheel subunits.6 Before the gas adsorption measurement, the samples were further activated by using the degassing port in the surface area analyzer for 10 h at 393 K. Sorption Measurements. UHP-grade gases were used throughout the gas sorption measurements. Low-pressure nitrogen (N2), carbon dioxide (CO2), methane (CH4), and hydrogen (H2) sorption experiments (up to 1 bar) were recorded on a Quantachrome IQ2 system. High pressure gas adsorption measurements were evaluated using Rubotherm ISOSORP-HyGpra+V adsorption analyzer over a pressure range of 0−20 bar or 0−100 bar. In the case of high-pressure measurements, the completely evacuated sample mass was monitored until equilibrium was attained at each pressure. Syntheses of Complexes 1 and 2. A mixture of Cu(NO3)2·3H2O (24.7 mg, 0.1 mmol), H3BTTC (12.6 mg, 0.03 mmol), DMF (2 mL), and several drops of HCl (2 mol L−1) was placed in a glass bottle, which was tightly capped and heated at 85 °C for 2 days and then cooled down to room temperature. A green cube-shaped crystal of complex 1 was obtained (yield: 42% based on Cu). IR spectrum (cm−1): 3433w, 3030m, 1711s, 1660m, 1602s, 1413m, 1289s, 1122m, 1065w, 924m, 863m, 773m, 688w, 667m, 579w, 542w. Anal. calcd for the evacuated sample of complex 1 (C10H2CuO5S2): C, 36.42; H, 0.61; found: C, 36.48; H, 0.73. In addition, yellowish-green cube-shaped crystal of complex 2 was isolated as a byproduct (yield: 4% based on Cu). IR spectrum (cm−1): 3429w, 3032m, 1715s, 1665m, 1604s, 1410m, 1282s, 1118m, 1068w, 920m, 861m, 765m, 683w, 662m, 582w, 539w. Elemental analysis and consequent gas adsorption experiments for desolvated sample of 2 were not possible to perform due to the very limited yield. X-ray Crystallography. Both crystal data were collected at 173(2) K. The X-ray diffraction intensity data for compound 1 was measured on a Bruker D8 QUEST PHOTON 100 CMOS detector with TRIUMPH curved crystal monochromator for Mo−Kα radiation (λ = 0.71073 Å). The diffraction data for compound 2 was collected on a Bruker Smart Apex DUO CCD diffractometer with graphitemonochromated Mo−Kα radiation. Raw data for all structures were processed using SAINT,7 and absorption corrections were applied using SADABS.8 The structures were solved by direct methods and refined anisotropically with SHELXTL using full-matrix least-squares procedures based on F2 values.9 For complex 1: The coordination aqua O3 atom lies in the Wyckoff position of i, the occupancy of which is constrained to 1/4, and its hydrogen atoms were not located. For compounds 1 and 2: the guest solvent molecules were chemically featureless to refine using conventional discrete-atom models. To resolve these issues, PLATON/SQUEEZE10 was employed to calculate the diffraction contribution of all guest molecules and, thereby, to yield a set of solvent-free diffraction intensities; subsequent

Table 1. Crystal Data and Refinement Information for Complexes 1 and 2 formula formula weight crystal system space group a /Å b /Å c /Å V /Å3 Z ρcalcd/g cm−3 μ /mm−1 collectedreflections unique reflections R1 [I > 2σ(I)] wR2 (all data)

1

2

C10H2CuO5S2 329.78 cubic Fm3̅m 36.4790(10) 36.4790(10) 36.4790(10) 48543(2) 48 0.541 0.645 21 179 2110 0.0950 0.2510

C40H8ClCu4O16S8 1290.67 cubic Pm3̅m 20.4434(14) 20.4434(14) 20.4434(14) 8544.0(10) 3 0.752 0.935 23 300 1633 0.0221 0.0610



RESULTS AND DISCUSSION Single-crystal X-ray diffraction analysis reveals that complex 1 crystallizes in cubic space group Fm3̅m with a relatively large unit cell dimension of 36.4790 Å. From crystallographic studies and elemental analysis, the formula is determined as Cu6(BTTC)4(H2O)6·xS (S represents noncoordinated solvent molecules). The asymmetric unit of 1 contains one-sixth BTTC3−, onequarter Cu(II) atom, as well as one-quarter coordinated aqua molecule (O3) (Figure S1a, Supporting Information). As shown in Figure 1a, two copper atoms are bridged by four carboxylates to generate a paddle-wheel secondary building unit (SBU) with Cu···Cu separation of 2.662 Å, where both two Cu atoms are five-coordinate in square-pyramidal geometry. It should be noted that all paddle-wheel unsaturated metal sites in 1 are furnished by axial aqua ligands (Cu1−O3: 2.173 Å), which can be easily removed by heating and/or evacuation without destroying the framework structure, thus forming vacant coordination sites. Obviously, each SBU links four BTTC3− moieties, and in turn each BTTC3− connects three SBUs to afford a truncated-octahedron (Figure 1b) with an internal diameter of about 1.0 nm (accounting for the van der Waals radii), where six vertices are completed by the SBUs, and four of the eight faces are crossed by BTTC3− ligands. Further, eight such octahedrons reside at the eight vertices of a cube, thereby resulting in a cuboctahedron via corner-sharing (Figure 1c), which were packed to create a non-interwoven 3D framework with large cavities of 1.2 nm in size. The free solvent molecules reside in the nanosized channels. The effective free void of 1, estimated by PLATON software,10 is ∼78% the crystal volume (37869/48543(2) Å3) upon guest solvents and coordinated water molecules removal, and the calculated density is 0.541 g cm−3. Topology analysis by the Topos 4.0 program11 suggests the 2 2 (6 ·8 ·102)3(63)4 topology symbol, which indicates 1 demonstrates the same topology as the prototypical tbo-type MOF (HKUST-1)12a and other isoreticular MOFs, such as the 2867

dx.doi.org/10.1021/cg500175k | Cryst. Growth Des. 2014, 14, 2866−2872

Crystal Growth & Design

Article

peripheral rings of 3.7°, which is identical to that of H3BTC in HKUST-1. X-ray diffraction study indicates that complex 2 also crystallizes in cubic but in a different space group, Pm3̅m, with a unit cell dimension of 20.4434 Å. From crystallographic studies, the formula is determined as [(CH3)2NH2]3[(Cu4Cl)3(BTTC)8]·yS (S represents noncoordinated solvent molecules). The countercation (CH3)2NH2+ is generated by in situ decomposition of the DMF solvent during the reaction. The asymmetric unit of 2 contains one-sixth BTTC3−, one16th μ4-Cl−, and one-quarter Cu (II) atom. Interestingly, compound 2 consists of a rare chloride-centered square-planar tetracopper cluster Cu4(μ4-Cl) as SBU (Figure 2a). All four Cu atoms in the SBU are five-coordinate in square-pyramidal geometry with a Cu−Cl distance of 2.529 Å. Every square SBU connects eight triangular BTTC3− ligands, and each BTTC3− binds three SBUs to form a truncated octahedral cage with inner sphere diameter of ca. 1.4 nm (Figure 2b), where the six vertices are located by SBUs and all eight faces are spanned by the BTTC3− moieties, different from that in 1. Subsequently, an anionic three-dimensional (3D) framework is formed by simple cubic packing of eight Oh-cages with large cavities of 1.4 nm (Figure 2c), which is occupied by the charge-balancing (CH3)2NH2+ cations and structurally disordered solvent molecules. Notably, the topology of 2 is assigned with the symbol the (Figure 2d), which has been known in several reported MOFs based on the similar square-planar tetranuclear M4Cl (or M4O) units, such as [Mn(DMF)6]3[(Mn4Cl)3(BTT) 8 (H 2 O) 12 ]·42DMF·11H 2 O·20CH 3 OH 13a (BTT = 1 ,3 ,5 -ben z e n e t r i st e t r a zo l a t e ) an d [ ( CH 3 ) 2 NH 2 ] 3 [(Cu4Cl)3(btc)8]·9DMA,13b as well as single-net of PCN-9.13c The total void accessible to guest molecules in compound 2 was estimated to be ∼68% per unit cell (5811/8544 Å3), taking into account the countercations, and the calculated density is 0.752 g cm−3. The PXRD profiles of complex 1 indicate that the framework preserves its crystallinity even though the guest solvents and coordinated water are removed (Figure S4, Supporting Information). To confirm the permanent porosity of compound 1, the methanol-exchanged bulk is activated under high vacuum at 393 K for 12 h to obtain the evacuated framework. The N2 adsorption for 1 at 77 K displays a completely reversible type-I adsorption behavior (Figure 3), characteristic of microporous materials. On the basis of the N2 sorption isotherm, the estimated apparent Brunauer−Emmett− Teller (BET) surface area is ∼1675 m2 g−1 (Langmuir surface ∼1858 m2 g−1), and the total pore volume is determined to be 0.655 cm3 g−1 for 1. The pore size distributions calculated from the N2 isotherms using the nonlocal density functional theory (NLDFT) method indicates the average pore size for 1 is 1.3 nm in diameter, consistent with the crystallographic data (ca. 1.2 nm). Low-pressure H2 sorption isotherms were recorded at 77 and 87 K to evaluate its H2 adsorption performance, which were completely reversible, as shown in Figure 4 inset. A significant H2 uptake of 1.37 and 0.91 wt % (153 and 102 cm3 g−1) is achieved at 77 and 87 K (1 bar) for compound 1 without saturation, respectively, which is quite close to the performance of well-known compounds PCN-6′ (1.37 wt %)12d and HKUST-1(1.44 wt %)14 at 77 K/1 bar. To better understand these observations, the adsorption enthalpy of H2 for activated 1 was calculated by a virial method (Figure S5, Supporting Information).15 At zero loading, indicative of the interaction of

Figure 1. Assembly of complex 1. (a) A Cu2-paddlewheel structural unit. (b) Truncated-octahedral cage. (c) Cuboctahedral cage from the [001] direction. (d) Schematic representation of the 3D network. Cu2(COO)4 SBU, blue; H3BTTC ligand, purple.

mesoMOF-1,12b PCN-6′,12c and single net of PCN-612d (Figure 1d). Although H3BTTC has much larger size than H3BTC (1,3,5-benzene tricarboxylate), H3BTTC in compound 1 is nearly planar with dihedral angles between central and 2868

dx.doi.org/10.1021/cg500175k | Cryst. Growth Des. 2014, 14, 2866−2872

Crystal Growth & Design

Article

Figure 3. N2 adsorption isotherm for complex 1 at 77 K and 0−1 bar, where adsorption and desorption branches are shown with filled and empty shape, respectively. Inset: calculated pore size distributions according to N2 isotherm.

Figure 4. H2 sorption isotherms for complex 1 at 77 K and 0−100 bar, where filled squares represent excess adsorption. (Inset) H2 sorption isotherm for 1 at 77 K (triangles) and 87 K (circles) under 0−1 bar.

H2 with the most energetically favored sites in the framework, Qst was determined to be ∼7.0 kJ mol−1 (Figure 6b), which is higher than those of “benchmark MOFs” with vacant metal cation sites, such as the NOTT series (5.68−6.70 kJ mol−1),16 SNU-21S (6.65 kJ mol−1),17 UMCM-150 (6.3 kJ mol−1),18 MIL-100 (6.3 kJ mol−1)19 and MIL-102 (6.0 kJ mol−1),20 PCN6′ (6.0 kJ mol−1)12d and PCN-46 (6.36 kJ mol−1),21 and socMOF (6.5 kJ mol−1).22 In addition, the value apparently surpasses that of the typical van der Waals-type interactions.23 It should be pointed out that the high enthalpy of H2 adsorption of compound 1 stems principally from the presence of the exposed metal Cu(II) sites on the pore surface. Furthermore, the high-pressure H2 sorption study was evaluated using the gravimetric measurement method up to 100 bar; at 77 K, the excess gravimetric H2 uptake capacity of 1 reaches 40 mg g−1. The value is lower than those of PCN-6′ (4.2 wt %) and PCN-6 (6.7 wt %) at 50 bar24 but close to that of HKUST-1 (3.3 wt %) at 77 bar25 and is superior to the reported MOFs with vacant coordination sites, such as MIL100 (3.28 wt %, 90 bar),19 Ni(dhtp)2 (1.8 wt %, 70 bar),26 and Sm2Zn3(oxdc)6 (1.19 wt %, 34 bar) under 77 K.27 The CO2 isotherms of 1 are collected at 273 and 298 K/1 bar, both of which are fully reversible. Activated 1 takes up a maximum CO2 uptake of 73.6 cm3 g−1 (3.29 mmol g−1, 14.5%) at 273 K and 40.5 cm3 g−1 (1.81 mmol g−1, 7.95%) at 298 K, respectively (Figure 5). The value is near that for the reported

Figure 2. Assembly of complex 2. (a) Chloride-centered square-planar tetracopper structural unit. (b) Truncated-octahedral cage. (c) Cuboctahedral cage from the [001] direction. (d) Schematic representation of the 3D network. Cu4(μ4-Cl) SBU, blue; H3BTTC ligand, purple. 2869

dx.doi.org/10.1021/cg500175k | Cryst. Growth Des. 2014, 14, 2866−2872

Crystal Growth & Design

Article

Figure 5. Sorption isotherms for CO2, CH4, and N2 at 273 and 298 K of complex 1 (adsorption and desorption branches are shown with filled and empty shapes, respectively).

MOFs under the same conditions.28 Meanwhile, the lowpressure CH4 sorption isotherms indicate that 1 has adsorption capacities of 19.76 cm3 g−1 (0.882 mmol g−1, 1.41%) and 12.2 cm3 g−1 (0.545 mmol g−1, 0.87%) at 273 and 298 K. Comparatively, the limited N2 uptake is only 5.36 cm3 g−1 (0.24 mmol g−1, 0.67%) and 3.31 cm3 g−1 (0.15 mmol g−1, 0.41%) at 273 and 298 K/1 bar, respectively. The adsorption capacity of CO2 for 1 is 3.0 and 11.2 times the values of CH4 and N2 at 298 K/1 bar, highlighting that compound 1 is a promising material for the highly selective separation of CO2/ CH4 and CO2/N2 at room temperature. Selective sorption on porous materials from gas mixtures can be calculated from single-component sorption isotherms.29 The ratios of the slopes were used to analyze the sorption selectivity for CO2 versus CH4 and N2. The CO2 versus N2 selectivity was 31:1 at 273 K and 19:1 at 298 K, and the CO2 versus CH4 selectivity was 4.9:1 at 273 K and 4.5:1 at 298 K (Figure S7, Supporting Information). The values are close to those of the ZIF materials28a (ZIF68, CO2/N2: 18.7, CO2/CH4: 5.0; ZIF69, CO2/N2: 19.9, CO2/CH4: 5.1; ZIF70, CO2/N2: 17.3, CO2/ CH4: 5.2; 298 K), MOF-5 (CO2/N2: 17.5, 298 K), MOF-177 (CO2/N2: 17.7, 298 K),30 SYSU (CO2/N2: 25.5, CO2/CH4: 4.7, 273 K)29a and {[Cu2(TCMBT)(bpp)(μ3-OH)]·6H2O}n (CO2/N2: 20.1, CO2/CH4: 4.0, 298 K).31 The significant CO2 sorption selectivity over CH4 and N2 at ambient condition for 1 can primarily be related to the vacant coordination sites, which function as charge-dense binding sites and interact more strongly with CO2 because of its greater quadrupole moment (CO2, 13.4 × 10−40 C·m2; N2, 4.7 × 10−40 C·m2) and polarizability (CO2, 29.1 × 10−25 cm−3; CH4, 25.9 × 10−25 cm−3; N2, 17.4 × 10−25 cm−3) compared with CH4 and N2. The isosteric heat (Qst) of CO2 and CH4 adsorption was calculated using experimental isotherm data at 273, 283, and 298 K (Figure S5, Supporting Information). Complex 1 exhibits strong binding affinity for CO2 (30.7 kJ mol−1) and CH4 (18.2 kJ mol−1) at zero coverage respectively (Figure 6a), which approach those of its prototype, HKUST-1 (35 and 18.7 kJ mol−1). Remarkably, the adsorption enthalpy for CO2 of 1 is much greater than those of “benchmark MOFs”, such as CuBTTri (21 kJ mol−1),32 MOF-5 (17 kJ mol−1),33 and NOTT-140 (25 kJ mol−1).34 The high initial Qst of CO2 can be mainly ascribed to the strong interactions between the exposed Cu2+ sites in the channel and CO2 molecules; therefore, the Qst decreases while the CO2 coverage rises because of the

Figure 6. Isosteric heats of CO2 and CH4 (a) and H2 adsorption (Qst) for complex 1 (b).

accessible vacant Cu2+ sites being reduced. More interestingly, the Qst of CO2 falls to 19.8 kJ mol−1 at loadings of 1.99 mmol/g but increases up to 23.8 kJ mol−1 at maximum loadings, which was attributed to increasing CO2−CO2 interactions.35 Then, its high-pressure adsorption was further investigated (Figure S8, Supporting Information), and the unsaturation excess CO2 uptake reaches 263.3 mg g−1 (5.98 mmol g−1) and 223.1 mg g−1 (5.07 mmol g−1) at 273 and 298 K (20 bar), respectively, which are close to the performance MSF series (MSF-1:24.1 wt %, MSF-3:21.6 wt %)36 and {[Cu2(TCMBT)(bpp)(μ3-OH)]·6H2O}n (25.5 wt %) at 298 K/20 bar.34 The unsaturation excess CH4 uptake reaches 38.42 mg g−1 (2.40 mmol g−1) and 30.26 mg g−1 (1.89 mmol g−1) at 273 and 298 K (20 bar), respectively.



CONCLUSIONS Two three-dimensional microporous compounds, Cu6(BTTC)4(H2O)6·xS (1) and [(CH3)2NH2]3[(Cu4Cl)3(BTTC)8]·yS (2), have been solvothermally synthesized and characterized, both of which are based on truncated octahedron subunits and contain uniform nanosized cavities but demonstrate different topological framework: tbo for 1 and the for 2. Notably, compound 1 exhibits high adsorption enthalpies for H2 and CO2 gas molecules, stemming principally from the presence of the exposed metal Cu(II) sites on the pore surface. In particular, activated 1 shows high efficiency for the selective adsorption of energy-correlated small molecules, including CO2 2870

dx.doi.org/10.1021/cg500175k | Cryst. Growth Des. 2014, 14, 2866−2872

Crystal Growth & Design

Article

Inubushi, Y.; Fukushima, T.; Kitagawa, S. Adv. Funct. Mater. 2013, 23, 3525−3530. (4) (a) Yuan, D. Q.; Zhao, D.; Timmons, D. J.; Zhou, H. C. Chem. Sci. 2011, 2, 103−106. (b) Yuan, D. Q.; Getman, R. B.; Wei, Z. W.; Snurr, R. Q.; Zhou, H. C. Chem. Commun. 2012, 48, 3297−3299. (5) Taerum, T.; Lukoyanova, O.; Wylie, R. G.; Perepichka, D. F. Org. Lett. 2009, 11, 3230−3233. (6) (a) Chen, B. L.; Ockwig, N. W.; Millward, A. R.; Contreras, D. S.; Yaghi, O. M. Angew. Chem., Int. Ed. 2005, 44, 4745−4749. (b) Yan, Y.; Lin, X.; Yang, S. H.; Blake, A. J.; Dailly, A.; Champness, N. R.; Hubberstey, P.; Schröder, M. Chem. Commun. 2009, 1025−1027. (c) Duan, J. G.; Zhen, Y.; Bai, J. F.; Zheng, B. S.; Li, Y. Z.; Li, S. H. Chem. Commun. 2012, 48, 3058−3060. (d) Hou, C.; Liu, Q.; Fan, J.; Zhao, Y.; Wang, P.; Sun, W. Y. Inorg. Chem. 2012, 51, 8402−8408. (7) SAINT, version 6.2; Bruker AXS, Inc.: Madison, WI, 2001. (8) Sheldrick, G. M. SADABS; University of Göttingen: Göttingen, Germany, 1997. (9) Sheldrick, G. M. SHELXTL, version 6.10; Bruker Analytical X-ray Systems: Madison, WI, 2001. (10) (a) Vandersluis, P.; Spek, A. L. Acta Crystallogr. A 1990, 46, 194−201. (b) Spek, A. L. J. Appl. Crystallogr. 2003, 36, 7−13. (11) (a) Blatov, V. A. TOPOS, A Multipurpose Crystallochemical Analysis with the Program Package; Samara State University, Russia, 2004. (b) Blatov, V. A.; Proserpio, D. M. Acta Crystallogr., Sect. A: Found. Crystallogr. 2009, 63, 329−343. (12) (a) Wong-Foy, A. G.; Matzger, A. J.; Yaghi, O. M. J. Am. Chem. Soc. 2006, 128, 3494−3495. (b) Wang, X. S.; Ma, S. Q.; Sun, D. F.; Parkin, S.; Zhou, H. C. J. Am. Chem. Soc. 2006, 128, 16474−16475. (c) Sun, D. F.; Ma, S. Q.; Ke, Y. X.; Collins, D. J.; Zhou, H. C. J. Am. Chem. Soc. 2006, 128, 3896−3897. (d) Ma, S. Q.; Sun, D. F.; Ambrogio, M.; Fillinger, J. A.; Parkin, S.; Zhou, H. C. J. Am. Chem. Soc. 2007, 129, 1858−1859. (13) (a) Dincǎ, M.; Dailly, A.; Liu, Y.; Brown, C. M.; Neumann, D. A.; Long, J. R. J. Am. Chem. Soc. 2006, 128, 16876−16883. (b) Tan, Y. X.; He, Y. P.; Zhang, J. Chem. Commun. 2011, 47, 10647−10649. (c) Ma, S. Q.; Zhou, H. C. J. Am. Chem. Soc. 2006, 128, 11734−11735. (14) Lee, J. Y.; Li, J.; Jagiello, J. J. Solid State Chem. 2005, 178, 2527− 2532. (15) Rowsell, J. L. C.; Yaghi, O. M. J. Am. Chem. Soc. 2006, 128, 1304−1315. (16) Lin, X.; Telepeni, I.; Blake, A. J.; Dailly, A.; Brown, C. M.; Simmons, J. M.; Zoppi, M.; Walker, G. S.; Thomas, K. M.; Mays, T. J.; Hubberstey, P.; Champness, N. R.; Schröder, M. J. Am. Chem. Soc. 2009, 131, 2159−2171. (17) Kim, T. K.; Suh, M. P. Chem. Commun. 2011, 47, 4258−4260. (18) Wong-Foy, A. G.; Lebel, O.; Matzger, A. J. J. Am. Chem. Soc. 2007, 129, 15740−15741. (19) Latroche, M.; Suble, S.; Serre, C.; Mellot-Draznieks, C.; Llewellyn, P. L.; Lee, J.-H.; Chang, J.-S.; Jhung, S. H.; Férey, G. Angew. Chem., Int. Ed. 2006, 45, 8227−8231. (20) Surblé, S.; Millange, F.; Serre, C.; Düren, T.; Latroche, M.; Bourrelly, S.; Llewellyn, P. L.; Férey, G. J. Am. Chem. Soc. 2006, 128, 14889−14896. (21) Zhao, D.; Yuan, D. Q.; Yakovenko, A.; Zhou, H. C. Chem. Commun. 2010, 46, 4196−4198. (22) Liu, Y.; Eubank, J. F.; Cairns, A. J.; Eckert, J.; Kravtsov, V. C.; Luebke, R.; Eddaoudi, M. Angew. Chem., Int. Ed. 2007, 46, 3278−3283. (23) Zhou, W.; Wu, H.; Hartman, M. R.; Yildirim, T. J. Phys. Chem. C 2007, 111, 16131−16137. (24) Ma, S. Q.; Eckert, J.; Forster, P. M.; Yoon, J. W.; Hwang, Y. K.; Chang, J.-S.; Collier, C. D.; Parise, J. B.; Zhou, H. C. J. Am. Chem. Soc. 2008, 130, 15896−15902. (25) Xiao, B.; Wheatley, P. S.; Zhao, X.; Fletcher, A. J.; Fox, S.; Rossi, A. G.; Megson, I. L.; Bordiga, S.; Regli, L.; Thomas, K. M.; Morris, R. E. J. Am. Chem. Soc. 2007, 129, 1203−1209. (26) Dietzel, P. D. C.; Panella, B.; Hirscher, M.; Blom, R.; Fjellvåg, H. Chem. Commun. 2006, 42, 959−961. (27) Wang, Y.; Cheng, P.; Chen, J.; Liao, D. Z.; Yan, S. P. Inorg. Chem. 2007, 46, 4530−4534.

over N2 and CH4 under ambient conditions, which enables complex 1 to possess potential applications in gas separation and purification. Attempts to obtain the large-scale samples of complex 2 are undergoing. We believe that our work will facilitate the exploration of more MOFs with versatile structures and excellent properties.



ASSOCIATED CONTENT

S Supporting Information *

X-ray crystallographic file in CIF format, additional crystal figures, TGA data, PXRD patterns, Qst calculation details for complex 1. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Authors

*(L.W.) E-mail: [email protected]. Fax: + 86 27 67867232. Phone: +86 27 67862900. *(D.L.) E-mail: dfl[email protected]. Author Contributions #

Y.H. and B.Z. contributed equally to this work.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the National Nature Science Foundation of China (Nos. 21171062, 21371065, 21371150 and 21301148), Program for Chenguang Young Scientists of Wuhan (2013070104010020).



REFERENCES

(1) (a) Rosi, N. L.; Eckert, J.; Eddaoudi, M.; Vodak, D. T.; Kim, J.; O’Keeffe, M.; Yaghi, O. M. Science 2003, 300, 1127−1129. (b) Yang, S. H.; Sun, J. L.; Ramirez-Cuesta, A. J.; Callear, S. K.; David, W. I. F.; Anderson, D. P.; Newby, R.; Blake, A. J.; Parker, J. E.; Tang, C. C.; Schröder, M. Nat. Chem. 2012, 11, 887−894. (c) Yang, S. H.; Lin, X.; Lewis, W.; Suyetin, M.; Bichoutskaia, E.; Parker, J. E.; Tang, C. C.; Allan, D. R.; Rizkallah, P. J.; Hubberstey, P.; Champness, N. R.; Thomas, K. M.; Blake, A. J.; Schröder, M. Nat. Mater. 2012, 11, 710− 716. (d) Jeazet, H. B. T.; Staudt, C.; Janiak, C. Chem. Commun. 2012, 15, 2140−2142. (e) Liang, X. Q.; Zhang, F.; Feng, W.; Zou, X. Q.; Zhao, C. J.; Na, H.; Liu, C.; Sun, F. X.; Zhu, G. S. Chem. Sci. 2013, 4, 983−992. (f) Chen, M. S.; Chen, M.; Takamizawa, S.; Okamura, T.-a.; Fan, J.; Sun, W. Y. Chem. Commun. 2011, 47, 3787−3789. (g) Zheng, B. S.; Bai, J. F.; Duan, J. G.; Wojtas, L.; Zaworotko, M. J. J. Am. Chem. Soc. 2011, 133, 748−751. (2) (a) Li, J. R.; Sculley, J.; Zhou, H. C. Chem. Rev. 2012, 112, 869− 932. (b) Wu, H. H.; Gong, Q. H.; Olson, D. H.; Li, J. Chem. Rev. 2012, 112, 836−868. (c) Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D. W. Chem. Rev. 2012, 112, 782−835. (d) Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van Duyne, R. P.; Hupp, J. T. Chem. Rev. 2012, 112, 1105−1125. (e) Jia, J. T.; Sun, F. X.; Ma, H. P.; Wang, L.; Cai, K.; Bian, Z.; Gao, L. X.; Zhu, G. S. J. Mater. Chem. A 2013, 1, 10112− 10115. (f) Cui, Y. J.; Yue, Y. F.; Qian, G. D.; Chen, B. L. Chem. Rev. 2012, 112, 1126−1162. (g) Lee, J. Y.; Farha, O. M.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Chem. Soc. Rev. 2009, 38, 1450−1459. (h) Qi, X. L.; Lin, R. B.; Chen, Q.; Lin, J. B.; Zhang, J. P.; Chen, X. M. Chem. Sci. 2011, 2, 2214−2218. (i) Han, Q. X.; He, C.; Zhao, M.; Qi, B.; Niu, J. Y.; Duan, C. Y. J. Am. Chem. Soc. 2013, 135, 10186−10189. (3) (a) Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D. W. Chem. Rev. 2012, 112, 782−835. (b) Zhang, Z. J.; Zhao, Y. G.; Gong, Q. H.; Li, Z.; Li, J. Chem. Commun. 2013, 49, 653−661. (c) Makal, T. A.; Li, J. R.; Lu, W. G.; Zhou, H. C. Chem. Soc. Rev. 2012, 41, 7761−7779. (d) Duan, J. G.; Higuchi, M.; Horike, S.; Foo, M. L.; Rao, K. P.; 2871

dx.doi.org/10.1021/cg500175k | Cryst. Growth Des. 2014, 14, 2866−2872

Crystal Growth & Design

Article

(28) (a) Banerjee, R.; Furukawa, H.; Britt, D.; Knobler, C.; Keeffe, M. O.; Yaghi, O. M. J. Am. Chem. Soc. 2009, 131, 3875−3877. (b) Cui, J. H.; Li, Y. Z.; Guo, Z. J.; Zheng, H. G. Chem. Commun. 2013, 49, 555− 557. (c) Millward, A. R.; Yaghi, O. M. J. Am. Chem. Soc. 2005, 127, 17998. (d) Deng, H.; Doonan, C. J.; Furukawa, H.; Ferriera, R. B.; Towne, J.; Knobler, C. B.; Wang, B.; Yaghi, O. M. Science 2010, 327, 846−850. (e) Zhang, Z.; Xiang, S.; Rao, X.; Zheng, Q.; Fronczek, F. R.; Qian, G.; Chen, B. Chem. Commun. 2010, 46, 7205−7207. (29) (a) Du, L. T.; Lu, Z. Y.; Zheng, K. Y.; Wang, J. Y.; Zheng, X.; Pan, Y.; You, X. Z.; Bai, J. F. J. Am. Chem. Soc. 2013, 135, 562−565. (b) Li, H. H.; Shi, W.; Zhao, K. N.; Niu, Z.; Chen, X. T.; Cheng, P. Chem.Eur. J. 2012, 18, 5715−5723. (c) Bae, Y.-S.; Farha, O. K.; Hupp, J. T.; Snurr, R. Q. J. Mater. Chem. 2009, 19, 2131−2134. (30) Saha, D.; Bao, Z. B.; Jia, F.; Deng, S. G. Environ. Sci. Technol. 2010, 44, 1820−1826. (31) Lu, Z. Y.; Xing, H.; Sun, R.; Bai, J. F.; Zheng, B. S.; Li, Y. Z. Cryst. Growth Des 2012, 12, 1081−1084. (32) Demessence, A. D.; Alessandro, M. D′; Foo, M. L.; Long, J. R. J. Am. Chem. Soc. 2009, 131, 8784−8786. (33) Choi, J.-S.; Son, W.-J.; Kim, J.; Ahn, W.-S. Microporous Mesoporous Mater. 2008, 116, 727−731. (34) Tan, C.; Yang, S.; Champness, N. R.; Lin, X.; Blake, A. J.; Lewis, W.; Schröder, M. Chem. Commun. 2011, 47, 4487−4489. (35) (a) Simmons, J. M.; Wu, H.; Zhou, W.; Yildirim, T. Energy Environ. Sci. 2011, 4, 2177−2185. (b) Dietzel, P. D. C.; Besikiotis, V.; Blom, R. J. Mater. Chem. 2009, 19, 7362−7370. (36) Zhang, J.; Xue, Y. S.; Liang, L. L.; Ren, S. B.; Li, Y. Z.; Du, H. B.; You, X. Z. Inorg. Chem. 2010, 49, 7685−7691.

2872

dx.doi.org/10.1021/cg500175k | Cryst. Growth Des. 2014, 14, 2866−2872