Cd(II)-Based Metal

Mar 15, 2017 - (17, 18). MOFs, as a special kind of adsorption material and an exceptional class of CO2 capture and separation material, have been the...
2 downloads 11 Views 3MB Size
Article pubs.acs.org/crystal

New Luminescent Three-Dimensional Zn(II)/Cd(II)-Based Metal− Organic Frameworks Showing High H2 Uptake and CO2 Selectivity Capacity Jiao Liu, Guo-Ping Yang,* Yunlong Wu, Yuwei Deng, Qingshan Tan, Wen-Yan Zhang,* and Yao-Yu Wang Key Laboratory of Synthetic and Natural Functional Molecule Chemistry of the Ministry of Education, Shaanxi Key Laboratory of Physico-Inorganic Chemistry, College of Chemistry & Materials Science, Northwest University, Xi’an 710127, Shaanxi, P. R. China S Supporting Information *

ABSTRACT: Three new three-dimensional (3D) luminescent metal−organic frameworks (MOFs), namely, [Zn2(L)2]· 2DMA·3H2O (1), [Zn2(L)2]·H2O (2), and [Cd2(L)2]·H2O (3) [H2L = 2-(imidazol-1-yl)terephthalic acid], have been solvothermally synthesized by using d10 metal ions Zn(II)/ Cd(II) and H2L in different solvent systems, which have been well characterized by elemental analysis, Fourier transform infrared spectroscopy, powder X-ray diffraction, and thermogravimetric analysis. As influenced by the different solvents and metal ions, single-crystal X-ray diffraction shows that 1 is a three-dimensional (3D) microporous framework with one-dimensional (1D) pores (10.85 × 8.79 Å2) based on the [Zn2(μ2COO)4] secondary building units, and 2 and 3 are two 3D isostructural networks composed by 1D zigzag chains, which are further connected by L2− ligands into dense packing structures. Topology analyses reveal that 1 can be simplified as a binodal (6,3)-connected ant topological net with a point symbol of (44·63·83)(48·62), and 2 and 3 show binodal (4,4)-connected nets with a point symbol of (4·63·82). Gas sorption behaviors of 1 for N2, H2, CH4, and CO2 have been studied in detail at different temperatures, indicating that the high H2 uptake and high selectivity for CO2 will make it as potential gas storage and separation materials. Moreover, the solid state luminescent properties of 1−3 have also been measured and studied at room temperature.



solvent systems, and so on,22−27 may have a crucial influence on final structures of MOFs. Thus, how to synthesize targeted MOFs with unique structures and functions is still a great challenge for chemists. It is known that multidentate Nheterocyclic dicarboxylate ligands can be a prior choice for constructing functional MOFs with high dimensionalities, surface areas, and pore volumes.28−30 Up to now, our group has long been committed to building functional porous MOFs by employing various N-heterocyclic dicarboxylate ligands as potential storage and separation materials for H2 and CO2.27,49 Thus, an unsymmetrical N-heterocyclic dicarboxylate ligand 2(imidazol-1-yl)terephthalic acid (H2L) was chosen to construct microporous MOFs in this work. Fortunately, three MOFs [Zn 2 (L) 2 ]·2DMA·3H 2 O (1), [Zn 2 (L) 2 ]·H 2 O (2), and [Cd2(L)2]·H2O (3) have been successfully synthesized under different hydrothermal reaction systems. Structural analyses show that 1 is a three-dimensional (3D) microporous framework with one-dimensional (1D) pores (10.85 × 8.79 Å2) based on the [Zn2(μ2-COO)4] secondary building unit (SBUs), and 2 and 3 are two 3D isostructural networks with a dense packing structure.

INTRODUCTION Metal−organic frameworks (MOFs), emerging as a class of promising crystalline microporous materials generally formed by the self-assembly of metal ions/clusters and bridging organic ligands with functional groups via coordination bonds and weak interactions, have attracted great interest in past few several decades not only because of their fascinating structures,1−10 but also owing to their tremendous potential applications,11−16 such as magnetism, gas storage/separation, luminescence, ion exchange and catalysis, and so on. On the other hand, the effect of greenhouse gases becomes greatly serious day-by-day, and scientists have thus paid close attention to plans to reduce CO2 emissions in the atmosphere. The bulk of the anthropogenic carbon dioxide emissions are from hydrocarbons both as fuels and in industrial processing, and given their abundance and relatively low cost, hydrocarbons will continue to be used in the foreseeable future. Therefore, the minimization of carbon emissions will require much more effective carbon capture, sequestration technologies, and new green energy materials.17,18 MOFs, as a special kind of adsorption material and an exceptional class of CO2 capture and separation material, have been the subject of numerous studies.19−21 During the selfassembly process, different factors, such as the spacer length and substituent groups of organic ligands, the size of ion radius, © XXXX American Chemical Society

Received: January 9, 2017 Revised: February 27, 2017

A

DOI: 10.1021/acs.cgd.7b00042 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Table 1. Crystal Data and Structure Refinements for 1−3

a

complex

1

2

3

empirical formula formula mass crystal system space group a [Å] b [Å] c [Å] α [°] β [°] γ [°] V [Å3] Z Dcalcd.[g·cm−3] μ [mm−1] F [000] θ [°] reflections collected goodness-of-fit on F2 final Ra indices [I > 2σ(I)]

C30H36N6O13Zn2 819.39 monoclinic C2/c 15.701(3) 14.964(3) 17.339(4) 90 115.096(3) 90 3689.1(13) 4 1.848 1.369 1688 1.976−24.999 8836/3197 0.958 R1 = 0.0740 wR2 = 0.2430

C22H14N4O9Zn2 609.11 orthorhombic Pbcn 17.400(6) 7.780(2) 15.564(5) 90 90 90 2106.8(11) 4 1.920 2.345 1224 2.341−24.974 9627/1845 1.005 R1 = 0.0476 wR2 = 0.1496

C22H14Cd2N4O9 703.17 orthorhombic Pbcn 17.7780(18) 7.8994(8) 15.9127(17) 90 90 90 2234.7(4) 4 2.090 1.968 1368 2.29−24.99 10133/1965 0.0223 R1 = 0.0195 wR2 = 0.0760

R1 = ∑∥F0| − |Fc∥/∑|F0|, wR2 = [∑w(F02 − Fc2)2/∑w(F02)2]1/2.

Figure 1. (a) Coordination environment of Zn(II) ions in 1. Symmetry codes: #1: −0.5 + x, 1.5 − y, −0.5 + y; #2: −x, y, 0.5 − z; #3: 0.5 − x, 1.5 − y, 1 − z; #4: −x, 2 − y, 1 − z; #5: x, 2 − y, −0.5 + z. (b) The [Zn2(COO)4] SBU unit in 1. (c) 3D dense framework structure of 1. (d) Topological net for 1.

All of these complexes are well characterized by elemental analyses, Fourier transform infrared spectroscopy, powder X-ray diffraction (PXRD), and thermogravimetric analyses (TGA). Gas sorption behaviors of 1 have been studied in detail at different temperatures. Moderately high H2 uptake and high selectivity for CO2 favor it as a potential material for gas storage and separation. In addition, the solid state luminescent properties

for 1−3 at room temperature have also been measured and discussed.



EXPERIMENTAL SECTION

Materials and Measurements. All of solvents and reagents used in experiments were purchased without purification. And the H2L ligand was bought from Jinan Camolai Trading Company. Elemental analyses B

DOI: 10.1021/acs.cgd.7b00042 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 2. (a) Coordination environment of Zn(II) ions in 2. Symmetry codes: #1: 1 − x, 2 − y, 1 − z; #2: −0.5 + x, 0.5 + y, 0.5 − z; #3: −0.5 + x, 1.5 − y, 1 − z. (b) Structure of 1D chains generated by H2L ligand view along the b axis. (c) 3D dense framework structure of 2. (d) Topological net for 2. of C, H, and N were carried out with a PerkinElmer 2400C elemental analyzer. Infrared spectra were measured on a Bruker EQUINOX-55 spectrophotometer in the 4000−400 cm−1 region. PXRD data were tested on a Bruker D8 ADVANCE X-ray powder diffractometer (Cu− Kα, 1.5418 Å). TGA were carried out by using a NETZSCH STA 449C microanalyzer thermal analyzer in N2 protection atmosphere with a heating rate of 10 °C min−1. The solid state luminescent spectra were collected on a Hitachi F-4500 fluorescence spectrophotometer at room temperature. All the gas sorption isotherms were measured with ASAP 2020 M adsorption equipment. Synthesis of [Zn2(L)2]·2DMA·3H2O (1). A mixture of Zn(NO3)2·H2O (0.1 mmol, 29.7 mg), H2L (0.05 mmol, 11.6 mg), H2O (4 mL), and DMA (4 mL) were mixed in 15 mL Teflon-lined stainless steel vessel, and then heated at 120 °C for 72 h. After that the vessel was cooled to room temperature with a rate of 10 °C h−1, and at last colorless block crystals were obtained. Yield 92% (based on H2L). Elemental analysis of 1, calculated (%): C, 43.97; H, 4.42; N, 10.25. Found: C, 43.55; H, 3.29; N, 9.73. FT-IR (KBr, cm−1, Figure S1): 3447 (m), 3123 (w), 2932 (w), 1639 (s), 1516 (m), 1382 (m), 1241 (m), 1065 (m), 816 (w), 772 (m), 533 (w). Synthesis of [Zn2(L)2]·H2O (2). Similar to the synthesis process of 1 except that the solvents were instead NMP (4 mL) and H2O (4 mL), colorless needle-like crystals of 2 were acquired. Yield 54% (based on H2L). Elemental analysis of 2, calculated (%): C, 43.37; H, 2.31; N, 9.19. Found: C, 43.46; H, 1.99; N, 8.99. FT-IR (Figure S2): 3460 (m), 3142 (w), 1645 (m), 1422 (s), 1384 (s), 1237 (m), 1091 (m), 849 (w), 779 (m), 645 (w). Synthesis of [Cd2(L)2]·H2O (3). Similar to the synthesis procedure of 2, except that the Zn(NO3)2·H2O was replaced by Cd(NO3)2·4H2O. Pale-yellow crystals were obtained. Yield 49% (based on H2L). Elemental analysis of 3, calculated (%): C, 37.57; H, 2.00. N, 7.97. Found: C, 36.74; H, 1.94. N, 7.65. FT-IR (Figure S3): 3358 (m), 3142 (m), 1626 (s), 1594 (s), 1397(s), 1326(s), 1244 (m), 1091(m), 766 (m), 639 (m), 537(w). Crystal Structure Determination. Single crystal X-ray diffraction measurements were obtained using a Bruker SMART APEX II CCD diffractometer equipped with graphite monochromated MoKα radiation (λ = 0.71073 Å) by using the ϕ/ω scan technique. The diffraction data were corrected for Lorentz and polarization effects as well as for empirical absorption based on multiscan. The structures of three

complexes were solved by direct methods and refined anisotropically on F2 by a full-matrix least-squares refinement with the SHELXTL program.31 The reflection datas were corrected by using the SADABS program. Anisotropic thermal parameters were applied to non-hydrogen atoms, and all hydrogen atoms from organic ligands were calculated and added at ideal positions. The crystallographic data of complexes are given in Table 1, and selected bond lengths and angles are listed in Table S1 (Supporting Information). CCDC numbers are 1516099−1516101 for 1−3.



RESULTS AND DISCUSSION Structure Description of [Zn2(L)2]·2DMA·3H2O (1). Complex 1 belongs to the monoclinic space group C2/c, and its asymmetric unit consists of one Zn(II) ion, one L2− ligand, one DMA molecule, and one and a half lattice water molecules. The Zn(II) ion is five coordinated with four equatorial plane carboxylate O atoms from four L2− and one axial imidazole N atoms (Figure 1a), showing a tetragonal pyramid coordination geometry. The Zn−O bond distances are in the range of 2.024− 2.055 Å, and the Zn−N bond length is 1.977 Å. One kind of coordination fashion of carboxylate groups of L2− exists in 1. The deprotonated carboxylates exhibit bridging bidentate mode (η2μ2χ2) with two Zn(II) ions (Figure S4a), forming a [Zn2(COO)4] paddlewheel SBU (Figure 1b),32,33 and then these SBUs are further extended by L2− spacers to generate a 3D porous framework with 1D pores 10.85 × 8.79 Å2 by calculating the distance of the neighboring C−C atoms in the two lateral H2L ligands including the van der Waals radius (Figure 1c).34 Interestingly, the porous channels are taken up by the uncoordinated DMA and H2O molecules. The calculated void volume is 55.2% of the total crystal volume by using the PLATON program after excluding DMA and H2O molecules. Here, topology analysis revealed that the [Zn2(COO)4] SBU and organic linker can act as 6- and 3-connected nodes, respectively. Thus, the whole framework of 1 can be simplified as a binodal (6,3)-connected ant topological net with a point symbol of (44· 63·83)(48·62) (Figure 1d). C

DOI: 10.1021/acs.cgd.7b00042 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Langmuir surface areas of 632.2 cm2 g−1, and a total micropore volume evaluated by Dubinin-Astakhov (D-A) method of 0.223 cm3 g−1 (Figure 3). The H2 sorption experiments at 77 K exhibit the uptake of 132.9 cm3 g−1 (1.18 wt %) at 1 bar, which is higher than those of MOF-5 (127 cm3 g−1) and MOF-177 (123 cm3 g−1), slightly lower than those of IRMOF-8 (165.86 cm3 g−1) and IRMOF-11(178.72 cm3 g−1) (Figure 4a).37 In contrast to the nitrogen sorption isotherms, the hydrogen sorption isotherm displays remarkable hysteresis between adsorption and desorption curves at 77 K. The hysteretic desorption indicates strong interaction between H2 and host frameworks and allows H2 to be adsorbed at high pressures but stored at lower pressures, which has been reported in a few recent examples.38−41 Furthermore, MOFs with active sites can usually demonstrate the high selectivity for different gas adsorption.42 Therefore, the potential application in CO2/CH4 separation has been evaluated carefully herein. In this work, the gas sorption properties of 1a have been tested at different temperatures (CO2 and CH4 at 195, 273, and 298 K). As shown in Figure 4b, the CO2 uptake reaches 208.8 cm3 g−1 at 1 bar, and the CO2 adsorption shows an obvious two-step adsorption behavior in CO2 adsorption isotherm at 195 K. In the initial step, 1a adsorbs 13.6 wt % (69.7 cm3 g−1), and in the second step at P/P0 > 0.2, the isotherm exhibits another sharp increase and then finally attains saturation at 1 atm. The desorption isotherm also exhibits steps at the corresponding infection points of the adsorption isotherm, but does not trace the adsorption process. There is, therefore, a marked broad hysteresis, confirming that the adsorbed CO2 is not immediately released on reducing the external pressure and is thus trapped within the framework at very low pressures. The likely reason for the stepwise and hysteretic gas sorption isotherms of the present material is related to the framework of 1a with shrunken pores, and the host structure may expand above the gate opening pressure, which should be another typical example of the “breathing” MOFs.23,49 However, the CH4 uptake of 114.4 cm3 g−1 is much lower than the value of CO2 uptake at 195 K (Figure 4b). Also, the CO2 uptakes are 52.4 cm3 g−1 (3.7 wt %) at 273 K and 27.3 cm3 g−1 (1.9 wt %) at 298 K and 1 bar, respectively (Figure 5a). The CO2 uptake of 1a at 298 K is inferior to that of some Zn(II)-based MOFs with a large pore size (9.0−32.0 Å), such as UMCM-1 (3.8 wt %),43 MOF-177 (3.5 wt %),44 and SNU-70 (3.5 wt %),45 all have typical monotonic isotherms (Type I) in which the more intense gas adsorption at low pressure forms a “knee” in the isotherm. Meanwhile, the CH4 uptakes are 11.9 cm3 g−1 (0.85 wt %) at 273 K and 6.5 (0.46 wt %) at 298 K at 1 bar. In short, the examination of both CO2 and CH4 sorption at different temperatures reveals that there is indeed the existence of selectivity for respective sorbate molecules. Thus, such adsorption selectivity is more evident at high temperature. The selectivity for CO2/CH4 (298 K) in 1a is also estimated by the ideal adsorbed solution theory (ISAT),48 as shown in Figure S7. To simulate typical compositions of biogas, the gas phase mole fraction for CO2/CH4 is set as 50/50 at 298 K (Figure 5b). The IAST can be used to predict multicomponent adsorption behaviors and the results show that selectivity of CO2 over CH4 rapidly ascends with increasing loading for both mixture compositions in 1a. At 1 bar, the calculated CO2/CH4 selectivity is 5.29 at 298 K from an equimolar gas-phase mixture. Remarkably, the gas phase mole fraction for CO2/CH4 is set as 20/80 and 80/20 in the same temperature, and the selectivity is shown in Figure S8, indicating that this material has high loading and affinity toward CO2.

Structure Description of [Zn2(L)2]·H2O (2). Single crystal X-ray analysis reveals that 2 and 3 are isostructural frameworks, and herein, only 2 is discussed and depicted for clarity. Complex 2 crystallizes in the orthorhombic space group Pbcn, and its asymmetric unit also contains one Zn(II) ion, four partially deprotonated L2− ligands, and half lattice water. In 2, the Zn(II) ions are pentacoordinated in a distorted trigonal bipyramidal configuration formed by four oxygen atom from three carboxylate groups and one nitrogen atoms of ligands (Figure 2a). The distances of Zn−O bonds range from 1.954 to 2.447 Å, and the distances of Cd−O bonds range from 2.214 to 2.467 Å ([Zn−N = 2.002 Å] [Cd−N = 2.190 Å]) in 2 and 3, respectively. In 2, the two carboxylate groups display different coordination modes, i.e., monodentate (η1μ1χ1), bridging-chelating bidentate (η3μ2χ3) (Figure S4b), to link Zn(II) ions to form an infinite 1D chain motif (Figure 2b), and further extended by imidazole N atoms to afford a 3D network (Figure 2c). On the basis of the simplification principle, assuming that the Zn(II) ions and L2− ligands are considered as 4-connected nodes, the 3D structure of 2 can thus be simplified as the binodal (4,4)-connected net with a point symbol of (4·63·82) (Figure 2d). Thermal Analysis and X-ray Powder Diffraction. To estimate the stability of the complexes, TGA were performed in N2 atmosphere, heating 25−750 °C. As shown in Figure S5, the TGA curve of 1 shows a 29.8% weight loss at ∼32−200 °C, corresponding to the escape of all solvents molecules (calcd 30.0%). The structure remained relatively stable from ∼200−316 °C, and then it collapsed rapidly. PXRD was used to check the purity of the bulk samples in the solid state (Figure S6). The sample of 1 was steeped by CH2Cl2 for a week, and then heated to 200 °C for 6 h under a high vacuum to remove the guest molecules to obtain the desolvated 1a, as evidenced by the FT-IR of 1a, wherein the characteristic CO vibration of DMA solvents is absent (Figure S1). The PXRD pattern of 1a still matched well with the simulated pattern of 1 generated from single-crystal diffraction data, indicating that the host framework of 1 is stable and intact after the removal of the guest molecules. Gas Adsorption Behaviors. To investigate the permanent porous properties of 1, the adsorption/desorption isotherms for N2, H2, CO2, and CH4 were measured carefully.32−33,35−36 As shown in the Figure 3, the reversible N2 sorption isotherm of 1a at 77 K exhibits a type-I isotherm with a Brunauer−Emmett− Teller (BET) surface area up to 481.8 cm2 g−1 (much lower than the theoretical value of 4315.4 m2 g−1, possibly because of the incomplete activation and/or potential defects in the sample),

Figure 3. N2 adsorption and desorption isotherm at 77 K and pore size distribution (inset) of 1a. D

DOI: 10.1021/acs.cgd.7b00042 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 4. (a) H2 adsorption and desorption isotherm at 77 K. (b) CO2 and CH4 adsorption and desorption isotherm at 195 K.

Figure 5. (a) CO2 and CH4 adsorption and desorption isotherm at 273 and 298 K. (b) IAST adsorption selectivity of 1a for equimolar mixtures of CO2 and CH4 at 298 K.

Figure 6. Isosteric heat of CO2 (a) and CH4 (b) adsorption for 1a estimated by the virial equation from the adsorption isotherms 298 K.

shape of Qst curve reveals that 1a has a far higher affinity toward CO2 at high loading. Although the Qst displays a gradual decrease with the increasing of CO2 coverage, the Qst still reaches 33.9 kJ mol−1 (Figure 6), which makes the frameworks highly polar and then causes specific interactions with CO2 because of its large quadrupole moment.28,49 Overall, the studies provide the theoretical prediction in CO2 capture, and the fact that 1a has high selectivity of CO2 implies that MOF 1 may be a promising adsorbent in the process of industrial application, such as biogas treatment and natural gas clean up.

The significant selectivity for CO2 adsorption over CH4 and N2 in 1a is mainly attributed to rich-N pyrazole rings, making the framework very polar, as a result, to form specific affinity for CO2, which has a larger quadrupole moment and a higher polarizability value (CO2, quadrupole moment = 1.43 × 10−39 C m2; polarizability = 29.1× 10−25 cm−3; CH4, quadrupole moment = 0 C m2; polarizability = 25.9 × 10−25 cm−3; N2, polarizability = 17.4 × 10−25 cm−3) compared to CH4 and N2.42,49−50 To further explore the affinity of 1a for CO2 and CH4, their adsorption enthalpies (Qst) were calculated according to the virial equation from sorption isotherms at 298 K,46 as shown in Figure S9. The E

DOI: 10.1021/acs.cgd.7b00042 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

*(W.-Y.Z.) E-mail: [email protected].

Luminescent Properties. The luminescent properties of complexes 1−3 and free ligand have been examined in the solid state at room temperature (Figure 7). Ligand exhibits an intense

ORCID

Guo-Ping Yang: 0000-0002-0230-6834 Yao-Yu Wang: 0000-0002-0800-7093 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We acknowledge the financial support from NSFC (Grant Nos. 21531007, 21371142, and 21201139), China Postdoctoral Science Foundation (Grant No. 2016M600807), and Postdoctoral Science Foundation of Northwest University (Grant No. 334100049).



Figure 7. Luminescent emission spectra of the free ligand H2L and complexes 1−3 at room temperature.

emission at 459 nm upon excitation at 400 nm, which can be attributed to the π*−n or π*−π transitions. The maximum emission peaks of 1−3 locate in the area of purple or bluishpurple, giving the main emission peaks at 457 nm for 1 (λex = 380 nm), 453 nm for 2 (λex = 400 nm), and 453 nm for 3 (λex = 400 nm). Compared with the emission spectra of ligand, the emission peaks of 1−3 are similar to that of the free ligand, which probably is ascribed to the intraligand transitions modified by the metal coordination.47−49



CONCLUSION In summary, three new luminescent MOFs have been successfully assembled by a multidentate N-heterocyclic dicarboxylate ligand. Interestingly, complex 1 possess a binodal (6,3)-connected ant topological network. More importantly, the desolvated framework of 1 displays permanent porosity with a suitable pore size and the polar pore surface decorated Nheterocyclic groups, affording high CO2 and H2 loading and highly selective capacity for CO2 over CH4 at 298 K. The unique structures and excellent gas selectivity may draw much more attention to fabrication of functional MOF materials.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.7b00042. Table for selected bonds and distances for 1−3, PXRD patterns, IR spectra, thermogravimetric analysis for 1−3, and sorption measurements (PDF) Accession Codes

CCDC 1516099−1516101 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



REFERENCES

(1) Peplow, M. Swiss-cheese-like materials called metal−organic frameworks have long promised to improve gas storage, separation and catalysis. Nature 2015, 520, 148−150. (2) Matsuda, R. Selectivity from flexibility. Nature 2014, 509, 434− 435. (3) Ando, S.; Fukushima, T.; Yamaguchi, T. Discrete Self-Assembly and Functionality of Guest Molecules in an Organic Framework. Chem. Mater. 2016, 28, 5847−5854. (4) Levine, D. J.; Runcevski, T.; Kapelewski, M. T.; Keitz, B. K.; Oktawiec, J.; Reed, D. A.; Mason, J. A.; Jiang, Z. H.; Colwell, K. A.; Legendre, C. M.; FitzGerald, S. A.; Long, J. R. Olsalaiine-Based MetalOrganic Frameworks as Biocompatible Platforms for H2 Adsorption and Drug Delivery. J. Am. Chem. Soc. 2016, 138, 10143−10150. (5) Deng, Y.-K.; Su, H.-F.; Xu, J.-H.; Wang, W.-G.; Kurmoo, M.; Lin, S.-C.; Tan, Y.-Z.; Jia, J.; Sun, D.; Zheng, L.-S. Hierarchical Assembly of a {MnII15MnIII4} Brucite Disc: Step-by-Step Formation and Ferrimagnetism. J. Am. Chem. Soc. 2016, 138, 1328−1334. (6) Meek, S. T.; Greathouse, J. A.; Allendorf, M. D.; Meek, S. T.; Greathouse, J. A.; Allendorf, M. D. Metal-organic frameworks: A rapidly growing class of versatile nanoporous materials. Adv. Mater. 2011, 23, 249−267. (7) Liu, F.-L.; Kozlevcar, B.; Strauch, P.; Zhuang, G.-L.; Guo, L.-Y.; Wang, Z.; Sun, D. Robust Cluster Building Unit: Icosanuclear Heteropolyoxocopperate Templated by Carbonate. Chem. - Eur. J. 2015, 21, 18847−18854. (8) Sakata, Y.; Furukawa, S.; Kondo, M.; Hirai, K.; Horike, N.; Takashima, Y.; Uehara, H.; Louvain, N.; Meilikhov, M.; Tsuruoka, T.; Isoda, S.; Kosaka, W.; Sakata, O.; Kitagawa, S. Shape-Memory Nanopores Induced in Coordination Frameworks by Crystal Downsizing. Science 2013, 339, 193−196. (9) Jiang, H.-L.; Xu, Q. Porous Metal-organic framework as platforms for functional applications. Chem. Commun. 2011, 47, 3351−3370. (10) Li, J.-R.; Sculley, J.; Zhou, H.-C. Metal−Organic Frameworks for Separations. Chem. Rev. 2012, 112, 869−932. (11) Furukawa, S. H.; Reboul, J.; Diring, S.; Sumida, K.; Kitagawa, S. Structuring of metal-organic frameworks at the mesoscopic/macroscopic scale. Chem. Soc. Rev. 2014, 43, 5700−5734. (12) Huang, W.-H.; Yang, G.-P.; Chen, J.; Chen, X.; Zhang, C.-P.; Wang, Y.-Y.; Shi, Q.-Z. Solvent Influence on Sizes of Channels in Three New Co(II) Complexes, Exhibiting an Active Replaceable Coordinated Site. Cryst. Growth Des. 2013, 13, 66−73. (13) Yuan, S.; Deng, Y.-K.; Sun, D. Unprecedented Second-Timescale Blue/Green Emissions and Iodine-Uptake-Induced Single-Crystal-toSingle-Crystal Transformation in ZnII/CdII Metal−Organic Frameworks. Chem. - Eur. J. 2014, 20, 10093−10098. (14) Oh, Y.; Le, V.-D.; Maiti, U. N.; Park, W. J.; Lim, J.; Lee, K. E.; Bae, Y.-S.; Kim, Y.-H.; Kim, S. O.; Hwang, J. O. Selective and Regenerative Carbon Dioxide Capture by Highly Polarizing Porous Carbon Nitride. ACS Nano 2015, 9, 9148−9157. (15) Zhang, Z.; Yao, Z.-Z.; Xiang, S.; Chen, B. Perspective of microporous metal−organic frameworks for CO2 capture and separation. Energy Environ. Sci. 2014, 7, 2868−2899.

AUTHOR INFORMATION

Corresponding Authors

*(G.-P.Y.) E-mail: [email protected]. F

DOI: 10.1021/acs.cgd.7b00042 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

(16) He, Y.; Zhou, W.; Qian, G.; Chen, B. Methane storage in metalorganic frameworks. Chem. Soc. Rev. 2014, 43, 5657−5678. (17) Simmons, J. M.; Wu, H.; Zhou, W.; Yildirim, T. Carbon capture in metal−organic frameworksa comparative study. Energy Environ. Sci. 2011, 4, 2177−2185. (18) Guo, L.-Y.; Zeng, S.-Y.; Jaglicic, Z.; Hu, Q.-D.; Wang, S.-X.; Wang, Z.; Sun, D. A Pyridazine-Bridged Sandwiched Cluster Incorporating Planar Hexanuclear Cobalt Ring and Bivacant Phosphotungstate. Inorg. Chem. 2016, 55, 9006−9011. (19) Ramdin, M.; Amplianitis, A.; Bazhenov, S.; Volkov, A.; Volkov, V.; Vlugt, T.; de Loos, T. W. Solubility of CO2 and CH4 in Ionic Liquids: Ideal CO2/CH4 Selectivity. Ind. Eng. Chem. Res. 2014, 53, 15427− 15435. (20) Wang, H.-H.; Jia, L.-N.; Hou, L.; Shi, W.-J.; Zhu, Z.; Wang, Y.-Y. A New Porous MOF with Two Uncommon Metal−Carboxylate− Pyrazolate Clusters and High CO2/N2 Selectivity. Inorg. Chem. 2015, 54, 1841−1846. (21) D’Alessandro, D. M.; Smit, B.; Long, J. R. Carbon dioxide capture: Prospects for new materials. Angew. Chem., Int. Ed. 2010, 49, 6058− 6082. (22) Wang, B.; Huang, H.; Lv, X.-L.; Xie, Y.; Li, M.; Li, J.-R. Tuning CO2 Selective Adsorption over N2 and CH4 in UiO-67 Analogues through Ligand Functionalization. Inorg. Chem. 2014, 53, 9254−9259. (23) Liu, B.; Li, Y.; Hou, L.; Yang, G.; Wang, Y.-Y.; Shi, Q.-Z. Dynamic Zn-based metal−organic framework: stepwise adsorption, hysteretic desorption and selective carbon dioxide uptake. J. Mater. Chem. A 2013, 1, 6535−6538. (24) Dai, F.-R.; Sambasivam, U.; Hammerstrom, A. J.; Wang, Z. Synthetic Supercontainers Exhibit Distinct Solution versus Solid State Guest-Binding Behavior. J. Am. Chem. Soc. 2014, 136, 7480−7491. (25) Chen, D.-M.; Zhang, X.; Shi, W.; Cheng, P. Tuning twodimensional layer to three-dimensional pillar-layered metal-organic frameworks: polycatenation and interpenetration behaviors. Cryst. Growth Des. 2014, 14, 6261−6268. (26) Chen, D.-M.; Ma, X.-Z.; Zhang, X.-J.; Xu, N.; Cheng, P. Switching a 2D Co(II) Layer to a 3D Co7-cluster-based metal-organic framework: syntheses, crystal structures, and magnetic properties. Inorg. Chem. 2015, 54, 2976−2982. (27) Yang, G. P.; Hou, L.; Ma, L. F.; Wang, Y.-Y. Investigation on the prime factors influencing the formation of entangled metal-organic frameworks. CrystEngComm 2013, 15, 2561−2578. (28) Li, J.; Yang, G. P.; Hou, L.; Cui, L.; Li, Y.; Wang, Y.-Y.; Shi, Q.-Z. Three new solvent-directed 3D lead(II)−MOFs displaying the unique properties of luminescence and selective CO2 sorption. Dalton Trans. 2013, 42, 13590−13598. (29) Gong, Y.-N.; Meng, M.; Zhong, D.-C.; Huang, Y.-L.; Jiang, L.; Lu, T.-B. Counter-cation modulation of hydrogen and methane storage in a sodalite-type porous metal−organic framework. Chem. Commun. 2012, 48, 12002−12004. (30) Wang, H.; Qin, J.; Huang, C.; Han, Y.; Xu, W.; Hou, H. Mono/ bimetallic water-stable lanthanide coordination polymers as luminescent probes for detecting cations, anions and organic solvent molecules. Dalton Trans. 2016, 45, 12710−12716. (31) Sheldrick, G. M. SHELXL-97, Program for Refinement of Crystal Structures; University of Göttingen: Germany, 1997. (32) Li, L.-J.; Bell, J. G.; Tang, S.-F.; Thomas, K. M.; Lv, Z.; Wang, C.; Xing, Y.; Zhao, X. Gas Storage and Diffusion through Nanocages and Windows in Porous Metal−Organic Framework Cu2(2,3,5,6-tetramethylbenzene- 1,4-diisophthalate)(H2O)2. Chem. Mater. 2014, 26, 4679− 4695. (33) Qian, J.-J.; Jiang, F.-L.; Zhang, L.-J.; Su, K.-Z.; Li, Q.-P.; Pan, J.; Yuan, D.; Hong, M. Unusual pore structure and sorption behaviour in a hexanodal zinc−organic framework material. Chem. Commun. 2014, 50, 1678−1681. (34) Zhao, Y.; Wu, H.; Emge, T. J.; Gong, Q.; Nijem, N.; Chabal; Yves, J.; Kong, L.; Langreth, D. C.; Liu, H.; Zeng, H.; Li, J. Enhancing Gas Adsorption and Separation Capacity through Ligand Functionalization of Microporous Metal−Organic Framework Structures. Chem. - Eur. J. 2011, 17, 5101−5109.

(35) Wu, P.; Guo, X.; Cheng, L.; He, C.; Wang, J.; Duan, C. Photoactive Metal−Organic Framework and Its Film for Light-Driven Hydrogen Production and Carbon Dioxide Reduction. Inorg. Chem. 2016, 55, 8153−8159. (36) Dybtsev, D.; Chun, H.; Kim, K. Rigid and Flexible: A Highly Porour Metal-Organic Framework with Unusual Guest-Dependent Dynamic Behavior. Angew. Chem., Int. Ed. 2004, 43, 5033−5036. (37) Rowsell, J.; Millward, A. R.; Park, K. S.; Yaghi, O. M. Hydrogen Sorption in Functionalized Metal-Organic Frameworks. J. Am. Chem. Soc. 2004, 126, 5666−5667. (38) Basdogan, Y.; Keskin, S. Simulation and modelling of MOFs for hydrogen storage. CrystEngComm 2015, 17, 261−275. (39) Xiang, S.; Huang, J.; Li, L.; Zhang, J.; Jiang, L.; Kuang, X.; Su, C.-Y. Nanotubular Metal-Organic Frameworks with High Porosity Based on T-Shaped Pyridyl Dicarboxylate Ligands. Inorg. Chem. 2011, 50, 1743− 1748. (40) Liu, D.; Wu, H.; Wang, S.; Xie, Z.; Li, J.; Lin, W. A high connectivity metal−organic framework with exceptional hydrogen and methane uptake capacities. Chem. Sci. 2012, 3, 3032−3037. (41) Song, P.; Li, Y.; He, B.; Yang, J.; Zheng, J.; Li, X. Hydrogen storage properties of two pillared-layer Ni(II) metal-organic frameworks. Microporous Mesoporous Mater. 2011, 142, 208−213. (42) Bourrelly, S.; Llewellyn, P. L.; Serre, C.; Millange, F.; Loiseau, T.; Férey, G. Different Adsorption Behaviors of Methane and Carbon Dioxide in the Isotypic Nanoporous Metal Terephthalates MIL-53 and MIL-47. J. Am. Chem. Soc. 2005, 127, 13519−13521. (43) Yazaydin, A. Ö .; Snurr, R. Q.; Park, T.-H.; Koh, K.; Liu, J.; LeVan, M. D.; Benin, A. I.; Jakubczak, P.; Lanuza, M.; Galloway, D. B.; Low, J. J.; Willis, R. R. Screening of Metal-Organic Frameworks for Carbon Dioxide Capture from Flue Gas Using a Combined Experimental and Modeling Approach. J. Am. Chem. Soc. 2009, 131, 18198−18199. (44) Millward, A. R.; Yaghi, O. M. Metal-Organic Frameworks with Exceptionally High Capacity for Storage of Carbon Dioxide at Room Temperature. J. Am. Chem. Soc. 2005, 127, 17998−17999. (45) Prasad, T. K.; Suh, M. P. Control of Interpenetration and GasSorption Properties of Metal−Organic Frameworks by a Simple Change in Ligand Design. Chem. - Eur. J. 2012, 18, 8673−8680. (46) Pham, T.; Forrest, K. A.; Tudor, B.; Elsaidi, S. K.; Mohamed, M. H.; McLaughlin, K.; Cioce, C. R.; Zaworotko, M. J.; Space, B. Theoretical Investigations of CO2 and CH4 Sorption in an Interpenetrated Diamondoid Metal−Organic Material. Langmuir 2014, 30, 6454−6462. (47) Huo, J. Z.; Su, X. M.; Wu, X. X.; Liu, Y. Y.; Ding, B. Hydrothermal synthesis and characterization of a series of luminescent AgIJI) coordination polymers with two new multidentate bis-(1,2,3-triazole) ligands: structural diversity, polymorphism and photoluminescent sensing. CrystEngComm 2016, 18, 6640−6652. (48) Chen, D.-M.; Ma, X.-Z.; Shi, W.; Cheng, P. Solvent-Induced Topological Diversity of Two Zn(II) Metal−Organic Frameworks and High Sensitivity in Recyclable Detection of Nitrobenzene. Cryst. Growth Des. 2015, 15, 3999−4004. (49) Li, G.-P.; Liu, G.; Li, Y.-Z.; Hou, L.; Wang, Y.-Y.; Zhu, Z. Uncommon Pyrazoyl-Carboxyl Bifunctional Ligand-Based Microporous Lanthanide Systems: Sorption and Luminescent Sensing Properties. Inorg. Chem. 2016, 55, 3952−3959. (50) Lin, R.-B.; Chen, D.; Lin, Y.-Y.; Zhang, J.-P.; Chen, X.-M. A Zeolite-Like Zinc Triazolate Framework with High Gas Adsorption and Separation Performance. Inorg. Chem. 2012, 51, 9950−9955.

G

DOI: 10.1021/acs.cgd.7b00042 Cryst. Growth Des. XXXX, XXX, XXX−XXX