[Cu2I]n and Cu2(CO2)

Nonmetallic Crystalline and Energy Conversion Materials, China Three Gorges ... chain and Cu2(CO2)4 paddlewheel SBUs which exhibited a new (3, 4, 4)- ...
0 downloads 0 Views 4MB Size
Article Cite This: Cryst. Growth Des. 2018, 18, 5449−5455

pubs.acs.org/crystal

A Microporous Heterovalent Copper−Organic Framework Based on [Cu2I]n and Cu2(CO2)4 Secondary Building Units: High Performance for CO2 Adsorption and Separation and Iodine Sorption and Release Jiaqi Yuan,† Jiantang Li,† Liang Kan,† Lifei Zou,† Jun Zhao,‡ Dong-Sheng Li,‡ Guanghua Li,† Lirong Zhang,*,† and Yunling Liu*,†

Downloaded via KAOHSIUNG MEDICAL UNIV on September 6, 2018 at 11:36:31 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



State Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, P. R. China ‡ College of Materials and Chemical Engineering, Key Laboratory of Inorganic Nonmetallic Crystalline and Energy Conversion Materials, China Three Gorges University, Yichang 443002, China S Supporting Information *

ABSTRACT: A microporous heterovalent Cu-based MOF, [(Cu2I)Cu2L2(H2O)2]22+·2NO3−·5DMF (1) (H2L = 5-(4H1,2,4-triazol-4-yl)benzene-1,3-dicarboxylic acid, DMF = N,Ndimethylformamide), has been successfully synthesized using the mixed secondary building unit (SBU) strategy. Compound 1 was constructed from [Cu2I]n chain and Cu2(CO2)4 paddlewheel SBUs and exhibited a new (3,4,4)-connected topology. Although compound 1 possesses moderate surface area, it exhibits better CO2 capture (91.9 cm3 g−1 at 273 K) and separation ability (selectivity for CO2 over CH4 = 8.2 under 1 bar at 298 K) than most of the reported CuxIy-based MOFs. Moreover, compound 1 is also a good candidate for the capture and release of iodine.



INTRODUCTION Currently, environmental pollution and energy shortage problems seem to be among the most intractable issues that scientists face.1−3 Reduction of emissions of CO2, one of the main gases that can cause the greenhouse effect, has attracted more and more attention from scientists all over the world.4−9 In addition, CO2 is also a main impurity of natural gas, which would reduce the energy conversion efficiency.10−13 Therefore, it is necessary to develop new materials that not only can effectively capture CO2 but also separate CO2 from natural gas to improve the energy conversion rate. In the last two decades, as a burgeoning category of porous materials, metal−organic frameworks (MOFs), which are constructed from organic ligands and metal ions or clusters through coordination bonds,14−17 have the features of high surface area and structural or functional tunability.18,19 MOFs constructed using metal halides (particularly hybrid silver(I) and copper(I) halides) as inorganic parts are a significant class of functional porous materials.20 Copper(I) halide MOFs have gained considerable interest because of their diverse structural and photophysical properties and potential applications as lowcost,21−28 plentiful materials in gas adsorption and separation,29,30 organic molecule adsorption and catalysis,31−36 etc. Secondary building unit (SBU) design is an effective synthetic strategy to construct diverse MOFs with robust and multifunctional properties.37−40 Copper(I) iodide-based © 2018 American Chemical Society

SBUs afford a series of geometrically diverse units, such as the tetrameric cubane-like [Cu4I4], the dimeric rhomboid [Cu2I2], the hexagonal-prism-shaped [Cu6I6], zigzag [CuxIy]n chains, etc.41−44 However, among the reported copper(I) iodide-based MOFs, only a few are robust enough to exhibit gas adsorption properties.45,46 CuI was used as the metal source since it can form various CuxIy clusters and Cu+ ions are also easy to partly convert into Cu2+ ions in air to enrich the diversity of SBUs.47−49 Cu2+ ions are more likely to form the classic Cu2(CO2)4 paddlewheel SBU by coordinating with O donors, which is helpful to improve the structural stability of MOFs based on CuxIy SBUs to develop gas adsorption ability.49 In addition, it is known that azolate ligands with N donors have the advantage of strong and directional coordination ability in connecting metal ions to construct stable frameworks, such as zeolitic imidazolate frameworks (ZIFs).50−52 Accordingly, Cu+ and I− ions can be coordinated with N donors to form much more stable [CuxIy]n chains than other CuxIy-based SBUs such as Cu4I4 clusters. On the basis of the above considerations, we chose the bifunctional organic ligand 5-(4H-1,2,4-triazol-4-yl)benzene1,3-dicarboxylate (L2−), which contains two types of Received: May 30, 2018 Revised: June 29, 2018 Published: July 11, 2018 5449

DOI: 10.1021/acs.cgd.8b00820 Cryst. Growth Des. 2018, 18, 5449−5455

Crystal Growth & Design

Article

Figure 1. Single-crystal structure of compound 1: (a) topology simplification of the Cu paddlewheel, [Cu2I]n chain, and ligand SBUs; (b) ball-andstick model of the 3D framework of compound 1 viewed along the [100] direction; (c) space-filling view of the structure of compound 1 along the [100] direction; (d) polyhedron view of the framework; (e) topological features of the compound displayed by tiles. For clarity, disordered atoms have been simplified, and H atoms on ligands have been omitted. purification. Powder X-ray diffraction (PXRD) data were collected on a Rigaku D/max-2550 diffractometer with Cu Kα radiation (λ = 1.5418 Å) over the 2θ range of 3−40° at room temperature. Elemental analysis (C, H, and N) was performed on a vario MICRO cube elemental analyzer (Elementar, Germany). Thermogravimetric analysis (TGA) was carried out on a TGA Q500 thermogravimetric analyzer used in air with a heating rate of 10 °C min−1. The liquidstate UV−vis spectra were recorded on a Shimadzu UV-2450 UV−vis spectrophotometer within 200−800 nm using the same solvent as the blank. N2, CO2, CH4, C2H6, and C3H8 gas adsorption measurements were carried out on a Micromeritics ASAP 2420 instrument at 77 K and a Micromeritics 3-Flex instrument at 273 and 298 K. Synthesis of Compound 1. A mixture of CuI (0.018 g, 0.09 mmol), H2L (0.009 g, 0.04 mmol), N, N-dimethylformamide (DMF) (1 mL), ethanol (0.5 mL), 1,4-diazabicyclo[2.2.2]octane (DABCO) (0.1 mL, 2 g of DABCO in 10 mL of DMF), HNO3 (0.3 mL, 2.2 mL of HNO3 in 10 mL of DMF) was sealed in a 20 mL vial and kept for 5 h at room temperature until it became a transparent dark-brown solution. The sealed vial was kept in an oven at 65 °C for 12 h and then cooled to room temperature. The blue block crystals were collected, washed with DMF, and air-dried (yield 76% based on CuI). Elemental analysis (%) calcd for C55H63Cu8I2N19O31: C, 29.35; H, 2.8; N, 11.83. Found: C, 30.25; H, 2.5; N, 12.36. The phase purity of compound 1 was proved by comparing the experimental PXRD pattern of the as-synthesized samples with the simulated one from the single-crystal X-ray data (Figure S1).

coordination groups (triazolate N donors and carboxylate O donors), as it can combine two different SBUs into one framework. Consequently, a novel microporous heterovalent copper−organic framework, [(Cu2I)Cu2L2(H2O)2]22+·2NO3−· 5DMF (1), has been successfully synthesized. It is constructed from two kinds of SBUs: [Cu2I]n chains and Cu2(CO2)4 paddlewheels53−55 and exhibits a novel (3,4,4)-connected topology. Compared with most of the CuxIy-based MOFs without adsorption ability, compound 1 possesses a moderate Brunauer−Emmett−Teller (BET) surface area (ca. 692 m2 g−1). Moreover, the results of ideal adsorbed solution theory (IAST) calculations show that compound 1 may be a candidate material for CO2 capture and separation since it exhibits good adsorption ability for CO2 and high selectivity for CO2 over CH4 (8.2 for 5:95 CO2/CH4 and 7.4 for 1:1 CO2/CH4 at 1 bar and 298 K). Meanwhile, the I2 sorption and release performance of activated samples of compound 1 in organic solvents has been investigated, and the results indicate that it is also a promising material for I2 capture.56



EXPERIMENTAL SECTION

Materials and Methods. All of the related chemicals were purchased from commercial sources and used without further 5450

DOI: 10.1021/acs.cgd.8b00820 Cryst. Growth Des. 2018, 18, 5449−5455

Crystal Growth & Design

Article

Figure 2. (a) CO2, (b) CH4, (c) C2H6, and (d) C3H8 gas sorption isotherms of compound 1 at 273 and 298 K under 1 bar.

ladder,59 the organic linker is considered as a 3-connected node with a triangular geometry, and the paddlewheel can be viewed as a 4-connected node with a square-planar geometry. Accordingly, the whole structure can be considered as a new (3,4,4)-connected topology with a Schläfli symbol of (42·63· 8)4(62·82·102)(62·82·92)(62·8)4. The topological characteristics presented by different tiles are displayed in Figures 1e and S6. The entire topology is composed of five types of tiles: [8·92], [62·102], [62·8·102], [42·62·92], and [42·62·82·92]. A total solvent-accessible volume of 4968.1 Å3 per unit cell was calculated using PLATON,61 corresponding to about 58.7% of the cell volume. The result indicates that compound 1 shows high porosity and potential application for gas storage. Thermogravimetric Analysis. TGA measurements were performed to explore the thermal stability of compound 1, and the results showed that compound 1 is stable up to about 230 °C (Figure S2). A weight loss of approximately 30% was witnessed prior to 200 °C, which can be ascribed to the removal of solvent molecules. The weight loss of about 47% between 230 and 500 °C can be attributed to removal of the organic ligands and collapse of the structure. The rest of the weight loss (22%) at 530 °C is ascribed to the remaining CuO. Property Characterization. Gas Adsorption and Separation Behaviors. N2 adsorption measurements at 77 K were carried out to verify the porosity of compound 1. As shown in Figure S4, the N2 uptake of activated compound 1 displays a typical type-I sorption isotherm, which is the feature of microporous materials, with a N2 uptake of 230 cm3 g−1 under 1 bar. The BET and Langmuir surface areas were calculated to be 692 and 932 m2 g−1, respectively. The BET surface area of compound 1 is higher than those of many reported Cu4I4based MOFs, such as COZ-1 (514 m2 g−1) and JLU-Liu23 (584 m2 g−1).62,63 The experimental pore volume was 0.36 cm3 g−1, which was close to the theoretical one (0.40 cm3 g−1). This indicates that the structure of compound 1 can be kept intact after the guest molecules are removed, which ensures its good adsorption performance. Considering the adsorption ability of compound 1, some other small-gas adsorption properties were also investigated.

Gas Adsorption Measurements. Before gas adsorption testing, the as-synthesized samples of compound 1 were exchanged in renewed ethanol for 3 days (six times per day) in order to totally remove the DMF molecules, as confirmed by TGA (Figure S2). The samples were activated by drying them in a vacuum oven at 90 °C for 20 min. Before the measurement, the activated samples were dehydrated again via the “outgas” function of the adsorption analyzer for 10 h at 80 °C. X-ray Crystallography. Single-crystal X-ray diffraction measurements were carried out on a Bruker Apex II CCD diffractometer with graphite-monochromatized Mo Kα radiation (λ = 0.71073 Å) at room temperature. The structure of compound 1 was solved by direct methods and refined on F2 by full-matrix least-squares with SHELXTL-NT version 5.1.57 All of the metal atoms were located first, and then the oxygen, carbon, and nitrogen atoms of the compound were subsequently found in difference Fourier maps. All non-hydrogen atoms were refined with anisotropic displacement parameters. The hydrogen atoms of the ligand were located geometrically. The final formula of the compound was designated from crystallographic data combined with elemental analysis and TGA data. The detailed crystallographic data and selected bond lengths and angles for the compound are listed in Tables S1 and S2, respectively. Related topology information on compound 1 was calculated using the TOPOS software.58



RESULTS AND DISCUSSION Structure Description. Single-crystal X-ray crystallographic analysis revealed that compound 1 crystallizes in the orthorhombic crystal system with space group Imma. There exists vibration of the ligands in the structure. As presented in Figure 1a,d, compound 1 is composed of three types of SBUs: organic SBUs composed of 3-connected ligands that have triazole and carboxyl groups, inorganic SBUs consisting of classical paddlewheels, and inorganic SBUs formed by [Cu2I]n chains.59,60 In the infinite zigzag SBUs, each Cu(I) atom is four-coordinated by two μ4-I atoms and two N atoms from two organic ligands, and each iodine atom adopts a fourfoldcoordination mode to link four Cu(I) atoms. As displayed in Figure 1b,c, along the [100] direction there exists a relatively large channel (13.3 Å × 4.7 Å) and a small one (6.7 Å × 4.0 Å) excluding the van der Waals radius. From the view of topology, the [Cu2I]n chain can be regarded as a 4-connected zigzag 5451

DOI: 10.1021/acs.cgd.8b00820 Cryst. Growth Des. 2018, 18, 5449−5455

Crystal Growth & Design

Article

Figure 3. (a, c) CO2, CH4, C2H6, and C3H8 adsorption isotherms at 298 K along with the dual-site Langmuir−Freundlich (DSLF) fits. (b, d) Gas mixture adsorption selectivities predicted by IAST at 298 K and 100 kPa for compound 1.

(Figure 3b), which were higher than the values for many reported microporous materials, especially most of the reported MOFs based on CuxIy SBUs, such as JLU-Liu31 (2.6, 2.7) and (Cu2I2)[Cu3PDC3(H2O)2]·2MeCN·2DMF (4) under identical measurement conditions.3,49 The practical application of compound 1 to separation of emblematic small hydrocarbons was explored by IAST as well. The identical fitting method was used to obtain the fitting parameters (Table S3). As presented in Figure 3d, the selectivities for C2H6 over CH4 and C3H8 over CH4 in equimolar mixtures were 15 and 40, respectively, at 298 K under 1 bar. The highly selective separation is ascribed to the difference in the interaction strengths between the framework and the small gases, as shown by Qst. The Qst values of compound 1 for CO2, CH4, C2H6, and C3H8 are 32.4, 22.6, 28.5, and 33.5 kJ mol−1, respectively, indicating that compound 1 prefers CO2 to CH4, C3H8 to CH4, and C2H6 to CH4 in separations. Iodine Sorption and Release Experiments. The I2 sorption and release performance of compound 1 in organic solvents has been investigated. Before the iodine adsorption testing, the microcrystalline samples were solvent-exchanged with renewed ethanol for 3 days in order to ensure complete removal of DMF molecules. Afterward, a 100 mg activated sample was immersed in 3 mL of a cyclohexane solution of I2 (0.01 M) sealed in a hermetic vial and placed under room temperature. As shown in Figure 4a, the adsorption of iodine could be detected by the fading of the color from deep purple to light pink after 48 h in the cyclohexane solution of I2, while the release of iodine could be witnessed by the color change to dark brown after a 20 mg iodine-adsorbed sample was soaked for 12 h in ethanol, as presented in Figure 4b. In the I2 adsorption process, the total uptake of I2 was about 1 per formula unit. The dynamic I2 release process was monitored by liquidstate UV−vis spectroscopy. As Figure 5 illustrates, the absorption bands in the UV−vis spectra presented λmax at 204, 288, and 360 nm, which can be attributed to I2 and polyiodide (I3−) formed by the reaction of I2 with decomposed iodide.70−72 The I2 release rate of compound 1 in ethanol is

First, CO2 adsorption measurements were carried out, and the adsorption isotherms are presented in Figure 2a. The CO2 uptakes by compound 1 were 92 and 59 cm3 g−1 at 273 and 298 K, respectively, under 1 bar. The isosteric adsorption enthalpy (Qst) of CO2 on compound 1 was also calculated by the virial model in order to explore the interaction between the framework and CO2 (Figure S9). The Qst value for compound 1 was 32.4 kJ mol−1 near zero coverage. The CO2 uptake (273 K) and Qst values for compound 1 are both higher than those for most of the CuxIy-based MOFs, such as JLU-Liu23 (39 cm3 g−1, 19 kJ mol−1)63 and (Cu2I2)[Cu3PDC3(H2O)2]·2MeCN· 2DMF (77 cm3 g−1, 27 kJ mol−1).49 These results illustrate the strong van der Waals interactions between the framework and CO2 molecules. Also, compound 1 should have the ability to adsorb other small hydrocarbons (e.g., CH4, C2H6, C3H8, etc.), which are the main constituents of natural gas. The adsorption performance of compound 1 for CH4, C2H6, and C3H8 was investigated at 273 and 298 K under 1 bar. The maximum uptakes of CH4 were 24 and 12 cm3 g−1, those of C2H6 were 70 and 55 cm3 g−1, and those of C3H8 were 61 and 53 cm3 g−1 at 273 and 298 K, respectively (Figure 2b−d). As calculated from the sorption isotherms at 273 and 298 K, the Qst values near zero coverage were 22.6, 28.5, and 33.5 kJ mol−1 for adsorption of CH4, C2H6, and C3H8, respectively (Figures S10−S12). To estimate the practical industrial separation ability of compound 1, the gas selectivities for CO2 (5% and 50%), C2H6 (50%), and C3H8 (50%) over CH4 were calculated using IAST as a common approach to analyze binary mixture adsorption according to the experimental single-component isotherms. By fitting the dual-site Langmuir−Freundlich equation data to match the experimental data (Figure 3a,c),64−66 we made the models match the isotherms at 298 K successfully (R2 > 0.999).67−69 Afterward, the fitting parameters (Table S3) were used to calculate multicomponent adsorption via IAST. According to the experimental data, the selectivities of compound 1 for CO2 over CH4 were 8.2 and 7.4 for 5:95 and 1:1 CO2/CH4 mixtures, respectively, at 298 K under 1 bar 5452

DOI: 10.1021/acs.cgd.8b00820 Cryst. Growth Des. 2018, 18, 5449−5455

Crystal Growth & Design

Article

Calculation of selectivity from IAST and PXRD patterns, TGA curves, N2 isotherms, XPS spectra, Qst plots, and structural data for compound 1 (PDF) Accession Codes

CCDC 1843778 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by e-mailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, U.K.; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Authors

*(L.Z.) E-mail: [email protected]. *(Y.L.) Fax: +86-431-85168624. E-mail: [email protected]. ORCID

Jiantang Li: 0000-0002-8963-5402 Dong-Sheng Li: 0000-0003-1283-6334 Guanghua Li: 0000-0003-3029-8920 Yunling Liu: 0000-0001-5040-6816

Figure 4. (a) Photographs of the time-dependent I2 adsorption process of 100 mg of compound 1 in 3 mL of cyclohexane. (b) Photographs of the time-dependent I2 release process of 20 mg of compound 1 in 3 mL of ethanol.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (21771078, 21671074, and 21621001), the 111 Project (B17020), and the National Key Research and Development Program of China (2016YFB0701100).



Figure 5. UV−vis spectra for I2 release from compound 1 in ethanol. The inset shows the plot of dynamic absorption intensity (monitored at 208 nm) vs time.

shown in the inset in Figure 5 as a plot of the absorption intensity versus time at 204 nm. The release of iodine rises abruptly at first and then mildly afterward, with a release rate of about 5.2 × 10−6 mol L−1 min−1 according to the standard curve (Figures S13 and S14), which is higher than those of JLU-Liu32 based on Cu4I4 clusters3 and [Cu4I3(DABCO)2]I3 based on [Cu8I6]2+ clusters.70



CONCLUSION By means of the helpful SBU synthesis strategy, a novel microporous heterovalent Cu-based MOF constructed from multiple SBUs with a new topology has been successfully solvothermally synthesized. Compound 1 displays moderate surface area and better adsorption ability for CO2 than most of the reported MOFs based on CuxIy SBUs. Meanwhile, it exhibits good selectivity for CO2 over CH4. In addition, compound 1 shows good performance for iodine sorption and release in organic solvents.



REFERENCES

(1) 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. (2) Yu, J.; Xie, L.-H.; Li, J.-R.; Ma, Y.; Seminario, J. M.; Balbuena, P. B. CO2 Capture and Separations Using MOFs: Computational and Experimental Studies. Chem. Rev. 2017, 117, 9674−9754. (3) Yao, S.; Sun, X.; Liu, B.; Krishna, R.; Li, G.; Huo, Q.; Liu, Y. Two Heterovalent Copper−Organic Frameworks with Multiple Secondary Building Units: High Performance for Gas Adsorption and Separation and I2 Sorption and Release. J. Mater. Chem. A 2016, 4, 15081−15087. (4) Zhang, K.; Qiao, Z.; Jiang, J. Molecular Design of Zirconium Tetrazolate Metal−Organic Frameworks for CO2 Capture. Cryst. Growth Des. 2017, 17, 543−549. (5) Liu, J.; Yang, G.-P.; Wu, Y.; Deng, Y.; Tan, Q.; Zhang, W.-Y.; Wang, Y.-Y. New Luminescent Three-Dimensional Zn(II)/Cd(II)Based Metal−Organic Frameworks Showing High H2 Uptake and CO2 Selectivity Capacity. Cryst. Growth Des. 2017, 17, 2059−2065. (6) Qi, Y.-J.; Zhao, D.; Li, X.-X.; Ma, X.; Zheng, W.-X.; Zheng, S.-T. Indium-Based Heterometal-Organic Frameworks with Different Nanoscale Cages: Syntheses, Structures, and Gas Adsorption Properties. Cryst. Growth Des. 2017, 17, 1159−1165. (7) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective Gas Adsorption and Separation in Metal−Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1477−1504. (8) Morris, R. E.; Wheatley, P. S. Gas Storage in Nanoporous Materials. Angew. Chem., Int. Ed. 2008, 47, 4966−4981. (9) Nandasiri, M. I.; Jambovane, S. R.; McGrail, B. P.; Schaef, H. T.; Nune, S. K. Adsorption, Separation, and Catalytic Properties of Densified Metal−Organic Frameworks. Coord. Chem. Rev. 2016, 311, 38−52. (10) Wang, D.; Zhao, T.; Li, G.; Huo, Q.; Liu, Y. A Porous SodaliteType MOF Based on Tetrazolcarboxylate Ligands and [Cu4Cl]7+

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.8b00820. 5453

DOI: 10.1021/acs.cgd.8b00820 Cryst. Growth Des. 2018, 18, 5449−5455

Crystal Growth & Design

Article

Squares with Open Metal Sites for Gas Sorption. Dalton Trans. 2014, 43, 2365−2368. (11) Adil, K.; Belmabkhout, Y.; Pillai, R. S.; Cadiau, A.; Bhatt, P. M.; Assen, A. H.; Maurin, G.; Eddaoudi, M. Gas/Vapour Separation Using Ultra-Microporous Metal−Organic Frameworks: Insights into the Structure/Separation Relationship. Chem. Soc. Rev. 2017, 46, 3402−3430. (12) Zhai, Q.-G.; Bu, X.; Zhao, X.; Li, D.-S.; Feng, P. Pore Space Partition in Metal−Organic Frameworks. Acc. Chem. Res. 2017, 50, 407−417. (13) Carrington, E. J.; McAnally, C. A.; Fletcher, A. J.; Thompson, S. P.; Warren, M.; Brammer, L. Solvent-Switchable ContinuousBreathing Behaviour in a Diamondoid Metal−Organic Framework and Its Influence on CO2 versus CH4 Selectivity. Nat. Chem. 2017, 9, 882−889. (14) Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Introduction to Metal− Organic Frameworks. Chem. Rev. 2012, 112, 673−674. (15) Zhou, H.-C.; Kitagawa, S. Metal−Organic Frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415−5418. (16) Zhai, Q.-G.; Bu, X.; Mao, C.; Zhao, X.; Daemen, L.; Cheng, Y.; Ramirez-Cuesta, A. J.; Feng, P. An Ultra-Tunable Platform for Molecular Engineering of High-Performance Crystalline Porous Materials. Nat. Commun. 2016, 7, 13645. (17) Li, B.; Wen, H.-M.; Cui, Y.; Zhou, W.; Qian, G.; Chen, B. Emerging Multifunctional Metal−Organic Framework Materials. Adv. Mater. 2016, 28, 8819−8860. (18) Li, J.; Luo, X.; Zhao, N.; Zhang, L.; Huo, Q.; Liu, Y. Two Finite Binuclear [M2(μ2-OH)(COO)2] (M = Co, Ni) Based Highly Porous Metal−Organic Frameworks with High Performance for Gas Sorption and Separation. Inorg. Chem. 2017, 56, 4141−4147. (19) Yuan, J.; Mu, L.; Li, J.; Zhang, L.; Li, G.; Huo, Q.; Liu, Y. A Water Stable Microporous Metal−Organic Framework Based on Rod SBUs: Synthesis, Structure and Adsorption Properties. CrystEngComm 2018, 20, 2169−2174. (20) Lei, X.-W.; Yue, C.-Y.; Zhao, J.-Q.; Han, Y.-F.; Yang, J.-T.; Meng, R.-R.; Gao, C.-S.; Ding, H.; Wang, C.-Y.; Chen, W.-D. LowDimensional Hybrid Cuprous Halides Directed by Transition Metal Complex: Syntheses, Crystal Structures, and Photocatalytic Properties. Cryst. Growth Des. 2015, 15, 5416−5426. (21) Jin, G.-X.; Wang, J.; Liu, J.-Y.; Ma, J.-P.; Dong, Y.-B. Visual Recognition and Removal of C2H2 from C2H4/C2H2 Mixtures by a CuI-MOF. Inorg. Chem. 2018, 57, 6218−6221. (22) Liu, J.; Wang, F.; Ding, Q.-R.; Zhang, J. Synthesis of an Enantipure Tetrazole-Based Homochiral CuI,II-MOF for Enantioselective Separation. Inorg. Chem. 2016, 55, 12520−12522. (23) Li, B.; Peng, Y.; Li, G.; Hua, J.; Yu, Y.; Jin, D.; Shi, Z.; Feng, S. Design and Construction of Coordination Polymers by 4-Amino-3,5bis(n-pyridyl)-1,2,4-triazole (n = 2, 3, 4) Isomers in a Copper(I) Halide System: Diverse Structures Tuned by Isomeric and Anion Effects. Cryst. Growth Des. 2010, 10, 2192−2201. (24) Liu, J.; Wang, F.; Zhang, J. Synthesis of Homochiral Zeolitic Tetrazolate Frameworks Based on Enantiopure Porphyrin-like Subunits. Cryst. Growth Des. 2017, 17, 5393−5397. (25) Yuan, S.; Wang, H.; Wang, D.-X.; Lu, H.-F.; Feng, S.-Y.; Sun, D. Reactant Ratio-Modulated Six New Copper(I)-iodide Coordination Complexes Based on Diverse [CumIm] Aggregates and Biimidazole Linkers: Syntheses, Structures and Temperature-Dependent Luminescence Properties. CrystEngComm 2013, 15, 7792−7802. (26) Peng, R.; Li, M.; Li, D. Copper(I) Halides: A Versatile Family in Coordination Chemistry and Crystal Engineering. Coord. Chem. Rev. 2010, 254, 1−18. (27) Deaton, J. C.; Switalski, S. C.; Kondakov, D. Y.; Young, R. H.; Pawlik, T. D.; Giesen, D. J.; Harkins, S. B.; Miller, A. J. M.; Mickenberg, S. F.; Peters, J. C. E-Type Delayed Fluorescence of a Phosphine-Supported Cu2(μ-NAr2)2 Diamond Core: Harvesting Singlet and Triplet Excitons in OLEDs. J. Am. Chem. Soc. 2010, 132, 9499−9508. (28) Liu, Z. W.; Qayyum, M. F.; Wu, C.; Whited, M. T.; Djurovich, P. I.; Hodgson, K. O.; Hedman, B.; Solomon, E. I.; Thompson, M. E.

A Codeposition Route to CuI-Pyridine Coordination Complexes for Organic Light-Emitting Diodes. J. Am. Chem. Soc. 2011, 133, 3700− 3703. (29) Banerjee, R.; Phan, A.; Wang, B.; Knobler, C.; Furukawa, H.; O’Keeffe, M.; Yaghi, O. M. High-Throughput Synthesis of Zeolitic Imidazolate Frameworks and Application to CO2 Capture. Science 2008, 319, 939−943. (30) Li, J.-R.; Sculley, J.; Zhou, H.-C. Metal−Organic Frameworks for Separations. Chem. Rev. 2012, 112, 869−932. (31) Yao, S.; Xu, T.; Zhao, N.; Zhang, L.; Huo, Q.; Liu, Y. An Anionic Metal−Organic Framework with Ternary Building Units for Rapid and Selective Adsorption of Dyes. Dalton Trans. 2017, 46, 3332−3337. (32) Wu, H.; Gong, Q.; Olson, D. H.; Li, J. Commensurate Adsorption of Hydrocarbons and Alcohols in Microporous Metal Organic Frameworks. Chem. Rev. 2012, 112, 836−868. (33) Lee, J. Y.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. B. T.; Hupp, J. T. Metal−Organic Framework Materials as Catalysts. Chem. Soc. Rev. 2009, 38, 1450−1459. (34) Ma, L.; Abney, C.; Lin, W. Enantioselective Catalysis with Homochiral Metal−Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1248−1256. (35) Yoon, M.; Srirambalaji, R.; Kim, K. Homochiral Metal− Organic Frameworks for Asymmetric Heterogeneous Catalysis. Chem. Rev. 2012, 112, 1196−1231. (36) Weng, Z.; Wu, Y.; Wang, M.; Jiang, J.; Yang, K.; Huo, S.; Wang, X.-F.; Ma, Q.; Brudvig, G. W.; Batista, V. S.; Liang, Y.; Feng, Z.; Wang, H. Active Sites of Copper-Complex Catalytic Materials for Electrochemical Carbon Dioxide Reduction. Nat. Commun. 2018, 9, 415. (37) Eddaoudi, M.; Moler, D. B.; Li, H.; Chen, B.; Reineke, T. M.; O’Keeffe, M.; Yaghi, O. M. Modular Chemistry: Secondary Building Units as a Basis for the Design of Highly Porous and Robust Metal− Organic Carboxylate Frameworks. Acc. Chem. Res. 2001, 34, 319−330. (38) O’Keeffe, M.; Yaghi, O. M. Deconstructing the Crystal Structures of Metal−Organic Frameworks and Related Materials into Their Underlying Nets. Chem. Rev. 2012, 112, 675−702. (39) Tranchemontagne, D. J.; Mendoza-Cortés, J. L.; O’Keeffe, M.; Yaghi, O. M. Secondary Building Units, Nets and Bonding in the Chemistry of Metal−Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1257−1283. (40) Wang, G.; Xue, Z.; Pan, J.; Wei, L.; Han, S.; Qian, J.; Wang, Z. Ligand-Oriented Assembly of a Porous Metal−Organic Framework by [CuI4I4] Clusters and Paddle-Wheel [CuII2(COO)4(H2O)2] Subunits. CrystEngComm 2016, 18, 8362−8365. (41) Abdulhalim, R. G.; Shkurenko, A.; Alkordi, M. H.; Eddaoudi, M. Supramolecular Isomers of Metal−Organic Frameworks Derived from a Partially Flexible Ligand with Distinct Binding Motifs. Cryst. Growth Des. 2016, 16, 722−727. (42) Kitagawa, H.; Ohtsu, H.; Kawano, M. Kinetic Assembly of a Thermally Stable Porous Coordination Network Based on Labile CuI Units and the Visualization of I2 Sorption. Angew. Chem., Int. Ed. 2013, 52, 12395−12399. (43) Wang, J.; Luo, J.; Luo, X.; Zhao, J.; Li, D.-S.; Li, G.; Huo, Q.; Liu, Y. Assembly of a Three-Dimensional Metal−Organic Framework with Copper(I) Iodide and 4-(Pyrimidin-5-yl) Benzoic Acid: Controlled Uptake and Release of Iodine. Cryst. Growth Des. 2015, 15, 915−920. (44) Fu, Z.; Lin, J.; Wang, L.; Li, C.; Yan, W.; Wu, T. Cuprous Iodide Pseudopolymorphs Based on Imidazole Ligand and Their Luminescence Thermochromism. Cryst. Growth Des. 2016, 16, 2322− 2327. (45) Hayashi, T.; Kobayashi, A.; Ohara, H.; Yoshida, M.; Matsumoto, T.; Chang, H.-C.; Kato, M. Vapochromic Luminescence and Flexibility Control of Porous Coordination Polymers by Substitution of Luminescent Multinuclear Cu(I) Cluster Nodes. Inorg. Chem. 2015, 54, 8905−8913. (46) Luo, X.; Sun, L.; Zhao, J.; Li, D.-S.; Wang, D.; Li, G.; Huo, Q.; Liu, Y. Three Metal−Organic Frameworks Based on Binodal 5454

DOI: 10.1021/acs.cgd.8b00820 Cryst. Growth Des. 2018, 18, 5449−5455

Crystal Growth & Design

Article

Inorganic Building Units and Hetero-O,N Donor Ligand: Solvothermal Syntheses, Structures, and Gas Sorption Properties. Cryst. Growth Des. 2015, 15, 4901−4907. (47) 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. (48) Yu, Q.; Zhu, L.-G.; Bian, H.-D.; Deng, J.-H.; Bao, X.-G.; Liang, H. A New Mixed-Valence Copper(I, II) Coordination Polymer with 1-D Chain Structure. Inorg. Chem. Commun. 2007, 10, 437−439. (49) He, H.; Sun, F.; Ma, S.; Zhu, G. Reticular Synthesis of a Series of HKUST-like MOFs with Carbon Dioxide Capture and Separation. Inorg. Chem. 2016, 55, 9071−9076. (50) 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. (51) Wang, F.; Fu, H.-R.; Kang, Y.; Zhang, J. A New Approach Towards Zeolitic Tetrazolate-Imidazolate Frameworks (ZTIFs) with Uncoordinated N-Heteroatom Sites for High CO2 Uptake. Chem. Commun. 2014, 50, 12065−12068. (52) Wang, F.; Liu, Z.-S.; Yang, H.; Tan, Y.-X.; Zhang, J. Hybrid Zeolitic Imidazolate Frameworks with Catalytically Active TO4 Building Blocks. Angew. Chem., Int. Ed. 2011, 50, 450−453. (53) Zeng, M.-H.; Yin, Z.; Tan, Y.-X.; Zhang, W.-X.; He, Y.-P.; Kurmoo, M. Nanoporous Cobalt (II) MOF Exhibiting Four Magnetic Ground States and Changes in Gas Sorption upon Post-Synthetic Modification. J. Am. Chem. Soc. 2014, 136, 4680−4688. (54) Śliwa, E. I.; Nesterov, D. S.; Kłak, J.; Jakimowicz, P.; Kirillov, A. M.; Smoleński, P. Unique Copper-Organic Networks Self-Assembled from 1,3,5-Triaza-7-Phosphaadamantane and Its Oxide: Synthesis, Structural Features and Magnetic and Catalytic Properties. Cryst. Growth Des. 2018, 18, 2814−2823. (55) Tomar, K.; Verma, A.; Bharadwaj, P. K. Exploiting Dimensional Variability in Cu Paddle-Wheel Secondary Building Unit Based Mixed Valence Cu(II)/Cu(I) Frameworks from a Bispyrazole Ligand by Solvent/pH Variation. Cryst. Growth Des. 2018, 18, 2397−2404. (56) Banerjee, D.; Chen, X.; Lobanov, S. S.; Plonka, A. M.; Chan, X.; Daly, J. A.; Kim, T.; Thallapally, P. K.; Parise, J. B. Iodine Adsorption in Metal Organic Frameworks in the Presence of Humidity. ACS Appl. Mater. Interfaces 2018, 10, 10622−10626. (57) Sheldrick, G. M. SHELXTL-NT, version 5.1; Bruker AXS: Madison, WI, 1997. (58) Blatov, V. A.; Shevchenko, A. P.; Proserpio, D. M. Applied Topological Analysis of Crystal Structures with the Program Package ToposPro. Cryst. Growth Des. 2014, 14, 3576−3586. (59) Schoedel, A.; Li, M.; Li, D.; O’Keeffe, M.; Yaghi, O. M. Structures of Metal−Organic Frameworks with Rod Secondary Building Units. Chem. Rev. 2016, 116, 12466−12535. (60) Zhang, J.-W.; Hu, M.-C.; Li, S.-N.; Jiang, Y.-C.; Zhai, Q.-G. Microporous Rod Metal−Organic Frameworks with Diverse Zn/CdTriazolate Ribbons as Secondary Building Units for CO2 Uptake and Selective Adsorption of Hydrocarbons. Dalton Trans. 2017, 46, 836− 844. (61) Spek, A. L. PLATON SQUEEZE: A Tool for the Calculation of the Disordered Solvent Contribution to the Calculated Structure Factors. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 9−18. (62) Kang, Y.; Wang, F.; Zhang, J.; Bu, X. Luminescent MTN-Type Cluster-Organic Framework with 2.6 nm Cages. J. Am. Chem. Soc. 2012, 134, 17881−17884. (63) Luo, X.; Cao, Y.; Wang, T.; Li, G.; Li, J.; Yang, Y.; Xu, Z.; Zhang, J.; Huo, Q.; Liu, Y.; Eddaoudi, M. Host-Guest Chirality Interplay: A Mutually Induced Formation of a Chiral ZMOF and Its Double-Helix Polymer Guests. J. Am. Chem. Soc. 2016, 138, 786−789. (64) Qadir, N. u.; Said, S. A. M.; Bahaidarah, H. M. Structural Stability of Metal Organic Frameworks in Aqueous Media-Controlling Factors and Methods To Improve Hydrostability and Hydrothermal Cyclic Stability. Microporous Mesoporous Mater. 2015, 201, 61−90. (65) 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. (66) Myers, A. L.; Prausnitz, J. M. Thermodynamics of Mixed-Gas Adsorption. AIChE J. 1965, 11, 121−127. (67) Lu, W.; Yuan, D.; Sculley, J.; Zhao, D.; Krishna, R.; Zhou, H.-C. Sulfonate-Grafted Porous Polymer Networks for Preferential CO2 Adsorption at Low Pressure. J. Am. Chem. Soc. 2011, 133, 18126− 18129. (68) Burd, S. D.; Ma, S.; Perman, J. A.; Sikora, B. J.; Snurr, R. Q.; Thallapally, P. K.; Tian, J.; Wojtas, L.; Zaworotko, M. J. Highly Selective Carbon Dioxide Uptake by [Cu(bpy-n)2(SiF6)] (bpy-1 = 4,4′-Bipyridine; bpy-2 = 1,2-Bis(4-pyridyl)ethene). J. Am. Chem. Soc. 2012, 134, 3663−3666. (69) Zhao, H.; Jin, Z.; Su, H.; Zhang, J.; Yao, X.; Zhao, H.; Zhu, G. Target Synthesis of a Novel Porous Aromatic Framework and Its Highly Selective Separation of CO2/CH4. Chem. Commun. 2013, 49, 2780−2782. (70) Xin, B.; Zeng, G.; Gao, L.; Li, Y.; Xing, S.; Hua, J.; Li, G.; Shi, Z.; Feng, S. An Unusual Copper(I) Halide-Based Metal−Organic Framework with a Cationic Framework Exhibiting the Release/ Adsorption of Iodine, Ion-Exchange and Luminescent Properties. Dalton Trans. 2013, 42, 7562−7568. (71) Wang, Z.; Huang, Y.; Yang, J.; Li, Y.; Zhuang, Q.; Gu, J. The Water-Based Synthesis of Chemically Stable Zr-Based MOFs Using Pyridine-Containing Ligands and Their Exceptionally High Adsorption Capacity for Iodine. Dalton Trans. 2017, 46, 7412−7420. (72) Yao, R.-X.; Cui, X.; Jia, X.-X.; Zhang, F.-Q.; Zhang, X.-M. A Luminescent Zinc(II) Metal−Organic Framework (MOF) with Conjugated π-Electron Ligand for High Iodine Capture and NitroExplosive Detection. Inorg. Chem. 2016, 55, 9270−9275.

5455

DOI: 10.1021/acs.cgd.8b00820 Cryst. Growth Des. 2018, 18, 5449−5455