A Highly Symmetric Metal–Organic Framework Based on a Propeller

Oct 29, 2013 - Ru(H2dcbpy)32+, one of the Ru(bpy)32+ (dcbpy = 2,2′-bipyridine-4,4′-dicarboxylic acid, bpy = 2,2′-bipyridine) derivatves, has bee...
8 downloads 8 Views 4MB Size
Article pubs.acs.org/crystal

A Highly Symmetric Metal−Organic Framework Based on a Propeller-Like Ru-Organic Metalloligand for Photocatalysis and Explosives Detection Shuquan Zhang,†,‡ Liang Han,†,§ Lina Li,†,§ Jing Cheng,† Daqiang Yuan,‡ and Junhua Luo*,†,‡ †

Key Laboratory of Optoelectronic Materials Chemistry and Physics and ‡State Key Laboratory of Structure Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China § Graduate School of the Chinese Academy of Sciences, Beijing 100039, China S Supporting Information *

ABSTRACT: Ru(H2dcbpy)32+, one of the Ru(bpy)32+ (dcbpy = 2,2′-bipyridine-4,4′-dicarboxylic acid, bpy = 2,2′-bipyridine) derivatves, has been used as a propeller-like photoactive metalloligand to coordinate with indium(III) ions to form a highly symmetric metal−organic framework [InRu(dcbpy)3][(CH3)2NH2]·6H2O (1), and the cubic microcrystals of 1 have been acquired through modified procedures. Compound 1 manifests broad visible light absorption band and strong red light emission with long decay lifetime, both of which are originated from the metal-to-ligand charge transfer of the Ru(dcbpy)34− metalloligands. Because of the highly lightharvesting and strong redox nature of the Ru(dcbpy)32+ units in 1, its photocatalysis activities were determined by visible lightinduced photodegradation of methyl orange experiments. The results indicate that 1 can be a stable and good visible-light driving heterogeneous photocatalyst. Meanwhile, the sensing properties of 1 were also evaluated, and the result shows that 1 can selectively detect the nitro explosives molecules.



INTRODUCTION In the past two decades, the rapidly evolved multifunctional metal−organic frameworks (MOFs) have attracted great interest due to their application in gas storage, separation, light emitting, biomedical imaging, chemical sensing, and heterogeneous catalysis.1 Incorporation of active functional units into MOFs has become an effective way to build multifunctional materials. One general methodology to introduce functional sites is to make use of functional organic groups,2 such as −NH2, −OH, and −CF3, decorated organic linkers. For instance, an −NH2 functionalized MIL-125(Ti) can not only enhance hydrogen adsorption but also can display high visible light absorption that makes it to be an effective photocatalyst for visible light-induced CO2 reduction.3 Introduction of open metal sites into traditional porous MOFs is another viable route to obtain multifunctional materials, as open metal sites usually play important roles in the gas storage, separation, and heterogeneous catalysis.4 Recently, a series of porous MOFs immobilized metalloporphyrin units in the structure exhibited both adsorption and heterogeneous catalysis properties.5 So far, chemists are still devoting their efforts to make the functional units into MOFs in order to explore novel multifunctional materials.6 Ru(II)-polypyridine complexes, one type of multifunctional metallo-organic compounds, have been extensively utilized in many research areas.7 For example, due to their broad © XXXX American Chemical Society

absorption band in visible region and their abilities to undergo photoredox quenching cycles, Stephenson and co-workers used Ru(bpy)3·Cl2 as homogeneous photoredox catalyst for many visible light induced radical addition reactions.7e,f Meanwhile, Yoon and co-workers explored new methods for [2 + 2] and [3 + 2] cycloaddition reactions by visible light photocatalysis using Ru(bpy)3·Cl2 as homogeneous photocatalyst.7c,d At the same time, Ru(bpy)32+ derivates are frequently used as the photosensitizers, because they not only have long-lived excited states but also are stable in either oxidized or reduced form.7i Although many reports focus on the Ru(II)-polypyridine monomers and their applications in homogeneous photochemical systems, the researches of MOFs based on Ru-organic functional units are limited. In the past decade, Ward and coworkers developed some Ru-(bpy)-cyanogens-based metalloligands and used them to construct ruthenium−lanthanide heteronuclear MOFs, which showed sensitized near-IR Ln(III) luminescence through effective Ru→Ln (Ln = Yb, Er, Nd) energy-transfer.8 Recently, Lin and co-workers reported some MOFs with Ru(bpy)32+ based metalloligands9 and discovered an interesting intraframework excited state energy migration phenomenon.9a,b Also, It is more attractive that doping Received: September 26, 2013 Revised: October 25, 2013

A

dx.doi.org/10.1021/cg401438j | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 1. (a) The synthesis of 1. (b) Stick/polyhedra model showing the connectivity of metalloligand and In centers in 1. (c) Space-filling model viewed along the [010] direction. (d) Schematic showing pcu topology of 1. crystals were filtered off and washed a few times with water. Yield: 152 mg (83%). Synthesis of Single Crystals of 1. A mixture of [Ru(H2dcbpy)3]·Cl2 (0.005 mmol, 5 mg) and InCl3 (0.015 mmol, 3.3 mg) was dissolved in 10 mL of DMF/H2O solution in a 20 mL vial and then a small amount of acid was added. The resulting solution was heated for 5 days at 90 °C. The product was obtained as a crystal (yield 57% based on [Ru(H2dcbpy)3]·Cl2). The phase purity of 1 was confirmed by comparing the powder X-ray diffraction (PXRD) patterns of the pristine sample and the simulated pattern from the crystal structure (Supporting Information Figure S3). The formula [InRu(dcbpy)3][(CH3)2NH2]·6H2O for 1 was obtained by the combination of TGAMS (see Supporting Information for detailed analysis, Figure S4) and elemental analysis. Anal. Calcd for C38H38N7O18RuIn: C, 41.62; H, 3.49; N, 8.94%. Found: C, 42.03; H, 3.23; N, 9.22%. IR (cm−1): 3384 (m, br), 3070 (w), 1595 (s), 1540 (s), 1377 (s), 1234 (w), 1112 (w), 1026 (w), 916 (w), 860 (w), 782 (m), 698 (m). Synthesis of Microcrystals of 1. A mixture of [Ru(H2dcbpy)3]·Cl2 (0.0025 mmol, 2.5 mg) and InCl3 (0.015 mmol, 3.3 mg) was dissolved in 20 mL of DMF/H2O solution in a 40 mL via, and then a small amount of acid was added. The resulting solution was heated for 10 h at 100 °C. After cooling to room temperature, the red precipitate was collected by centrifugation and washed with methanol twice. After drying in air, the yield based on [Ru(H2dcbpy)3]·Cl2 was 38%. Photodegradation of Methyl Orange Experiments. To characterize the photocatalytic activities of the samples under visible light irradiation, measurements on photodegradation of methyl orange (MO) were carried out at room temperature. For each experiment, 30 mg of the sample was put in a beaker containing 80 mL of 10 mg/L MO aqueous solution. A 500 W Xe lamp with a cutoff filter to remove the radiation below 420 nm was used as the visible light source. The solution was magnetically stirred for 1 h in the dark to establish an adsorption/desorption equilibrium of MO on the surfaces of the samples, and then the light was turned on to initiate the photocatalytic reaction. Three milliliters of the suspension was extracted and centrifuged to remove the MOF particles after every 20 min during the course of 120 min irradiation. The concentration of MO in the photocatalytic degradation process was calibrated by the intensity of its characteristic absorption peak located at 464 nm measured with a UV−vis Lambda 35 spectrometer. The control experiments on the

Ru(bpy)32+-derived bridging ligands into the traditional porous MOF (UIO-67) can make these MOFs as good heterogeneous photocatalysts for visible light-induced mild organic transformations.9c To the best of our knowledge highly symmetric MOFs with such functional Ru(II)-polypyridine building units have not been reported. Here, we want to report the design and synthesis of a highly symmetric MOF with broad visible-light absorption and strong red luminescence based on a photoactive Ru(dcbpy)32+ metalloligand and its applications in heterogeneous photodegradation of organic dye and detection of nitro explosive molecules.



EXPERIMENTAL SECTION

Materials and Measurements. All reactants were reagent grade and used as purchased without further purification. Elemental analyses for C, H, and N were carried out on a German Elementary Vario EL III instrument. The FT-IR spectra were performed on a Nicolet Magna 750 FT-IR spectrometer using KBr pellets in the range of 4000−400 cm−1. The thermal decomposition behavior was analyzed by thermogravimetric analysis-mass spectrometry (TGA-MS) using a NETSCH STA-449C thermoanalyzer coupled with a NETSCH QMS403C massspectrometer. The power X-ray diffraction (PXRD) patterns were collected by a Rigaku DMAX2500 X-ray diffractometer using Cu Kα radiation (λ = 0.154 nm). Fluorescent analysis was performed on an Edinburgh Instruments FLS920 spectrofluorimeter equipped with both continuous (450 W) and pulse xenon lamps. UV/ visible absorbance was collected in the solid state at room temperature on a Perkin-Elmer Lambda 650S UV/vis spectrometer equipped with Labsphere integrating over the spectral range 300−800 nm using BaSO4 as reflectance standards. The diffuse reflectance measured was converted to Kubelka−Munk Function.10 The optical band gap was estimated to be 2.19 eV, as indicated in Supporting Information Figure S18. Synthesis of the Metalloligand [Ru(H 2 dcbpy) 3 ]·Cl 2 . [Ru(H2dcbpy)3]·Cl2 was synthesized by following the published procedure.11 RuCl3·3H2O (50 mg) and H2dcbpy (140 mg) were placed in a 23 mL Teflon-lined stainless steel container where 1 mL of HCl (37%) and 2 mL of water were added. The mixture was heated at 200 °C for 4 h and cooled to room temperature. The dark red block B

dx.doi.org/10.1021/cg401438j | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 2. SEM image of cubic shaped microcrystals of 1. Scale bars = (a) 10 μm, (b) 100 μm, (c) 200 μm. photodegradation of MO were carried out in the following conditions: (1) in the dark; (2) without catalyst. Nitro Compound Sensing Experiment. The nitro-explosives sensing experiments were carried out using a thin layer of 1 cubic microcrystals. The thin layers were prepared through a handy method. Glass slides, 20 mm × 10 mm in size, were rinsed by deionized water and acetone and dried by nitrogen flow. Double-sided tapes (8 mm × 2 mm) were then applied to the slightly lower half of the slides. The microcrystals of 1 were then evenly sprinkled onto the surfaces of the slides. The slides were then turned face-down and gently tapped to remove any powder that was not glued well to the surface of the slides. Before the sensing experiments were performed, the fluorescence of glass slides, the glass slides with double-sided tapes and thin layers of 1 were measured respectively. The glass slides and the glass slides with double-sided tapes show almost no luminescence when compared to the thin layer of 1 (Supporting Information Figure S6). To demonstrate the applicability of 1 for nitro compounds detection, the fluorescence spectra of the thin layers were monitored, before and after immersing them in the different methanol solution of nitro and non-nitro compounds (0.0044 M for TNT and 0.1 M for other compounds) for varied periods of time (10, 20, 40, 80, 180, and 600 s, Figure S7−S15, Supporting Information). The quenching efficiency (%) is estimated using the formula (I0 − I)/I0 × 100%, where I0 = original peak maximum intensity and I = maximum intensity after immersion.

within the cavities are highly disordered, and the potential solvent and cation-accessible volume in 1 was estimated by PLATON12 to be 70.2% (Supporting Information Figure S2). In recent years, nano- and microscale particles of MOFs become more and more attractive due to their highlighted performances over bulk crystals, which have reported to be usable in biomedical, chemical sensing, and some other applications.13 Thus, we prepared the cubic microcrystals of 1 by modified procedures. SEM images show that the particles are in cubic shape, which is in accordance with the Ohsymmetric crystal structure of 1, and the dimensional sizes of the cubic particles are ranging from 5 to 10 μm (Figure 2). The PXRD studies indicate that the micrometer-sized particles keep the cubic framework structure of 1 (Supporting Information Figure S20). In this work, three reasons spur us to prepare the cubic microcrystals. First, scaling down the crystal size can make the particles easy to be dispersed in the solution. Second, small-sized cubic microcrystals offer large surface area for the close contact between them and solution. Finally, microsized crystals are helpful to excited state energy migration from interior to surface of crystals, which makes the excited state energy to be used efficiently. These advantages are beneficial to the photocatalysis and sensing experiments described below. The Photophysical Properties. Diffuse reflectance UV− vis measurements showed a broad metal-to-ligand charge transfer (MLCT) absorption band between 250 and 650 nm for 1, consistent with the spectrum of the metalloligand [Ru(H2dcbpy)3]·Cl2 in solid state (Figure 3). The broad absorption band in visible light region makes this material



RESULTS AND DISCUSSION Description of the Crystal Structure and Morphology. X-ray crystallographic analysis reveals that 1 crystallizes in the high symmetric cubic space group Fm3̅m, with 1/48 [Ru(dcbpy)3]4− metalloligand, 1/48 In(III) ion, 1/8 lattice water molecule, and 1/48 dimethylamine cation in the asymmetric unit. It is noticeable that the [Ru(dcbpy)3]4− metalloligand in the structure is disordered at four equivalent positions with the ratio of 1:1:1:1 (Supporting Information Figure S1a). Therefore, the pseudo-C3-symmetric propeller-like metalloligand [Ru(dcbpy)3]4‑ in 1 can be seen as an Oh-symmetric octahedral building unit (Supporting Information Figure S1b) which bridges six In(III) ions to act as a 6-connected node. At the same time, the In(III) center adopts octahedral geometry coordinated by six carboxyl oxygen atoms from six different [Ru(dcbpy)3]4− metalloligands, serving as another 6-connected node (Figure 1b). Interestingly, the angles between Ru(II) ions that locate at the center of [Ru(dcbpy)3]4− building units and In(III) ions are either 90 or 180° and the distances between them are equal. As a result, the two types of 6-connected nodes are equivalent in topology and the highly symmetric threedimensional (3D) microporous framework 1 can be rationalized as six-connected cubic pcu network structure (Figure 1d), which displays 3D open channel along the crystallographic a, b, and c axes with square windows of diagonal about 4 Å (Figure 1c). The solvent molecules and the Me2NH2+ cations located

Figure 3. Absorptance and emission spectra of metalloligand [Ru(H2dcbpy)3]·Cl2 and 1. C

dx.doi.org/10.1021/cg401438j | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 4. (a) UV−vis absorption spectra of methyl orange solution degraded by 1 after the UV−visible light irradiation for different time intervals. (b) Control experiments on the photodegradation of MO: in the dark, without catalyst, and 1 with visible light. (c) Photograph showing the photocatalytic degradation under UV−vis light for 0, 20, 40, 60, 80, 100, and 120 min (from left to right, respectively).

Figure 5. (a) Schematic of the experimental procedures for luminescence quenching measurements. (b) Quenching efficiency of different nitro and non-nitro compounds. (c) Comparison of quenching efficiency between p-DNB and TNT at different concentrations. (d) Photograph of thin layer without UV light irradiation (left), under UV light irradiation before immersing in p-DNB (middle), and under UV light irradiation after immersing in p-DNB (right).

we also investigated the quantum yield (QY) of these complexes. Similar to the fluorescence lifetime, the QY of 1 is a slightly smaller than the metalloligand [Ru(H2dcbpy)3]·Cl2, which is 2.8% for 1 and 4.2% for Ru-based metalloligand. These values are comparable to other Ru-metalloligand based MOFs.9d,e Also, the differences of fluorescence lifetime and QY between the metalloligand and the MOF can be tentatively ascribed to the disordered thermal motion of the solvent molecules and the Me2NH2+ cations in the cavities of 1, which may quench the fluorescence of MOF through the collision between framework and guest molecules.14 Photocatalysis. The broad absorption band in visible region of 1 and the redox nature of Ru(II)-polypyridine based metalloligand motivated us to explore its potential applications

potential to utilize the solar radiation. The free metalloligand displays a strong red photoluminescence with the emission maximum at 633 nm upon excitation at 400 nm in the solid state, and the as-synthesized 1 fluoresces with an emission centered at 657 nm (Figure 3). The similar shape and energy of the emission bands of two compounds indicate that the fluorescence of 1 originated from the metalloligand. The tiny red shift may be attributed to the coordination of metalloligands with In(III) centers and the stability of the coordination framework. The emission of both metalloligand and 1 are in accord with the fluorescence of Ru-based complexes reported previously.9d The fluorescence lifetime of the metalloligand [Ru(H2dcbpy)3]·Cl2 was determined to be 512 ns, while it is found to be 301 ns for 1. At the same time, D

dx.doi.org/10.1021/cg401438j | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

more sensitive to TNT than other nitro compounds. And it is noticeable that the detectable luminescence responses of TNT were observed at ppm scale and the quenching efficiency reaches to 25% at a concentration of 125 ppm. Meanwhile, the luminescent intensity of the thin layers could be fully recovered simply by rinsing them in methanol for several minutes after each measurement (Supporting Information Figure S19). Mechanism for the Explosives Detecting. From the crystal structure of 1, we can find that the small pore size of this complex might exclude their encapsulation of the analytes, which indicates the sensing mechanism of our experiments might be different from most cases of guest-induced luminescent response reported before.18 Therefore, this fluorescence quenching might be attributed to the photoinduced electron transfer from electron-donating framework to high electron-withdrawing analytes adsorbed on the surface of the MOF particles on the thin layer. The selectivity of this MOF for different analytes could be caused by the electrondonating and electron-withdrawing effect of different substituent groups on the analytes. The strongly electronwithdrawing −NO2 groups can increase the electron deficient degrees of the analytes that will enhance their quenching abilities. In contrast, electron-donating −CH3 groups can decrease the electron deficiency and weaken the quenching abilities. It is noteworthy that the quenching efficiency of DMNB is lower than NB may due to the weaker interaction of aliphatic nitro compound than aromatic ones with MOF particles.17d This quenching phenomenon resembles the organic polymer films that are usually used in organic compound vapor detection.19 In addition, as depicted in Supporting Information Figure S18, the large optical band gap (ca. 2.19 eV) indicates that the framework in the excited state is highly reductive, providing the driving force for the electron transfer to the electron deficient analytes.17d

in heterogeneous photocatalysis. The photocatalytic activity of 1 was evaluated by photodecomposition of methyl orange (MO), one of the most stable azo dyes that is extensively used in the textile industry and resistant to biodegradation, photodegradation under visible-light illumination.15 As shown in Figure 4a, while with 1 as a catalyst and under the irradiation of visible light, the intensity for the absorption band of MO at 464 nm gradually decreased as a function of increasing reaction time, whereas in the dark or without catalyst, no decomposition of MO was observed (Figure 4b). This suggests that the MO can be obviously photodegraded by 1 under visible light. Such degradation was easily observed by the naked eye (Figure 4c), and about 80% of MO had been decomposed upon visible light irradiation for 120 min. To the best of our knowledge, the decomposition of MO may attribute to the highly active hydroxyl radical (·OH)15b that might be generated during the redox cycles of the [Ru(dcbpy)3]4− metalloligands in 1. Finally, the catalyst can be readily recovered from the reaction mixture via simple filtration. In our experiment, about 89% catalyst can be recovered (26.6 mg), and the loss might be caused during the process of recovery, which is hard to avoid. The photostability of the catalysts was monitored by using PXRD patterns during the course of photocatalytic reactions (Supporting Information Figure S20), and the results showed that the recovered catalysts are identical to the as-synthesized compounds, implying that this MOF photocatalyst maintains its structural integrity after photocatalysis reaction. All these experiments confirmed that 1 could be one stable and efficient heterogeneous photocatalyst. Detection of Explosives. Detection of explosives and explosive-like substances, which usually are nitro compounds, is a very important for the society security and environmental safety. Compared with the traditional detection methods,16 such as canines or sophisticated instruments, luminescence sensor emerged as a fast, convenient, and low-cost approach recently.17 Owing to its long MLCT lifetime and outstanding light emitting properties, 1 may be a potential sensory material for nitro explosives detecting. As shown in Figure 5b, the fluorescence intensity of the thin layer of 1 was almost unchanged after soaking in methanol solution of some nonnitro analytes such as benzene, benzoic acid (BA), toluene, and phenol. But obvious quenching of fluorescence intensity was observed upon immersing the thin layer in methanol solution of nitro compounds such as nitrobenzene (NB), nitrotoluene (NT), 2,3-dimethyl-2,3-dinitrobutane (DMNB), 2,4-dinitrotoluene (DNT), dinitrobenzene (m-DNB and p-DNB), and trinitrotoluene (TNT). The results indicate that 1 has potential for selective sensing of nitro compounds. Further studies indicate that during the same immersing time (80 s), DNB shows most significant quenching ability (p-DNB ≈ m-DNB, 63.5 and 61.7%, respectively) and the order of quenching efficiencies for the selected nitro compounds is p-DNB ≈ mDNB > DNT > NB > DMNB > NT. Because the concentration of TNT is about 22.7 times lower than other nitro compounds, the quenching efficiency of TNT is only 38.3%. To compare the quenching abilities of TNT with other nitro compounds, we measured the fluorescent quenching abilities after immersing the thin layers of 1 in methanol solution of TNT and p-DNB with the same concentration (1.25, 12.5, 125, and 1250 ppm) during the same periods of time (80 s), respectively (Figure 5c and Supporting Information Figure S16, 17). The result shows that the quenching efficiency of TNT is apparently larger than that of the p-DNB with the same concentration. Therefore, 1 is



CONCLUSIONS In summary, an Oh-symmetric metal−organic framework 1 has been successfully constructed with the photoactive metalloligand [Ru(H 2 dcbpy)3]·Cl 2. The 1 displays a broad absorption band in visible spectrum region and visible red light emission with long MLCT lifetime. Furthermore, due to its encouraging light-harvesting and photoluminescent merits, its useful applications in heterogeneous photocatalysis and explosive molecules detection have been studied in detail. The compound 1 exhibits effective visible-light driving photodegradation of MO, and meanwhile, it also could selectively detect the nitro explosive molecules in parts per million scale. The easily recoverable and reusable abilities make 1 a good candidate for both heterogeneous photocatalyst and explosivedetectable material. Our group will devote our efforts to explore the applications of this type of photoactive MOF materials and some artificial photosyntheses are currently being pursued.



ASSOCIATED CONTENT

S Supporting Information *

Crystallographic data in CIF format, selected bond lengths (Å) and bond angles (°) for 1, structural resolving, thermogravimetric analysis, XRD patterns of 1, and some additional figures. This material is available free of charge via the Internet at http://pubs.acs.org. E

dx.doi.org/10.1021/cg401438j | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design



Article

(5) (a) Feng, D. W.; Gu, Z. Y.; Li, J. R.; Jiang, H. L.; Wei, Z. W.; Zhou, H. C. Angew. Chem., Int. Ed. 2012, 51, 10307−10310. (b) Hupp, J. T. Nature Chem. 2010, 2, 432−433. (c) Yang, X. L.; Xie, M. H.; Zou, C.; He, Y. B.; Chen, B. L.; O’Keeffe, M.; Wu, C. D. J. Am. Chem. Soc. 2012, 134, 10638−10645. (d) Zou, C.; Wu, C. D. Dalton Trans. 2012, 41, 3879−3888. (e) Zhang, Z.; Zhang, L.; Wojtas, L.; Nugent, P.; Eddaoudi, M.; Zaworotko, M. J. J. Am. Chem. Soc. 2012, 134, 924−927. (f) Zhang, Z.; Zhang, L.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M. J. J. Am. Chem. Soc. 2012, 134, 928−933. (6) (a) Das, M. C.; Xiang, S. C.; Zhang, Z. J.; Chen, B. L. Angew. Chem., Int. Ed. 2011, 50, 10510−10520. (b) Grunker, R.; Bon, V.; Heerwig, A.; Klein, N.; Muller, P.; Stoeck, U.; Baburin, I. A.; Mueller, U.; Senkovska, I.; Kaskel, S. Chem.Eur. J. 2012, 18, 13299−13303. (c) Li, H. H.; Shi, W.; Zhao, K. N.; Niu, Z.; Li, H. M.; Cheng, P. Chem.Eur. J. 2013, 19, 3358−3365. (d) Maspoch, D.; Ruiz-Molina, D.; Veciana, J. Chem. Soc. Rev. 2007, 36, 770−818. (e) Qiu, S. L.; Zhu, G. S. Coord. Chem. Rev. 2009, 253, 2891−2911. (f) Rocha, J.; Carlos, L. D.; Paz, F. A. A.; Ananias, D. Chem. Soc. Rev. 2011, 40, 926−940. (f) Xiong, S. S.; He, Y. B.; Krishna, R.; Chen, B. L.; Wang, Z. Y. Cryst. Growth Des. 2013, 13, 2670−2674. (g) Kobayashi, A.; Suzuki, Y.; Ohba, T.; Noro, S. I.; Chang, H. C.; Kato, M. Inorg. Chem. 2011, 50, 2061−2063. (7) (a) Numata, Y.; Singh, S. P.; Islam, A.; Iwamura, M.; Imai, A.; Nozaki, K.; Han, L. Y. Adv. Funct. Mater. 2013, 23, 1817−1823. (b) Hussain, M.; El-Shafei, A.; Islam, A.; Han, L. Y. Phys. Chem. Chem. Phys. 2013, 15, 8401−8408. (c) Lu, Z.; Shen, M. H.; Yoon, T. P. J. Am. Chem. Soc. 2011, 133, 1162−1164. (d) Du, J.; Yoon, T. P. J. Am. Chem. Soc. 2009, 131, 14604−14605. (e) Narayanam, J. M. R.; Stephenson, C. R. J. Chem. Soc. Rev. 2011, 40, 102−113. (f) Wallentin, C. J.; Nguyen, J. D.; Finkbeiner, P.; Stephenson, C. R. J. J. Am. Chem. Soc. 2012, 134, 8875−8884. (g) Zhao, Q.; Huang, C. H.; Li, F. Y. Chem. Soc. Rev. 2011, 40, 2508−2524. (h) Zhao, Q. A.; Li, F. Y.; Huang, C. H. Chem. Soc. Rev. 2010, 39, 3007−3030. (i) Yui, T.; Tamaki, Y.; Sekizawa, K.; Ishitani, O. Top. Curr. Chem. 2011, 303, 151−184. (8) (a) Baca, S. G.; Adams, H.; Sykes, D.; Faulkner, S.; Ward, M. D. Dalton Trans. 2007, 2419−2430. (b) Miller, T. A.; Jeffery, J. C.; Ward, M. D.; Adams, H.; Pope, S. J. A.; Faulkner, S. Dalton Trans. 2004, 1524−1526. (c) Adams, H.; Alsindi, W. Z.; Davies, G. M.; Duriska, M. B.; Easun, T. L.; Fenton, H. E.; Herrera, J. M.; George, M. W.; Ronayne, K. L.; Sun, X. Z.; Towrie, M.; Ward, M. D. Dalton Trans. 2006, 39−50. (d) Herrera, J. M.; Ward, M. D.; Adams, H.; Pope, S. J. A.; Faulkner, S. Chem. Commun. 2006, 1851−1853. (e) Herrera, J. M.; Pope, S. J. A.; Meijer, A. J. H. M.; Easun, T. L.; Adams, H.; Alsindi, W. Z.; Sun, X. Z.; George, M. W.; Faulkner, S.; Ward, M. D. J. Am. Chem. Soc. 2007, 129, 11941−11504. (9) (a) Xie, Z. G.; Ma, L. Q.; deKrafft, K. E.; Jin, A.; Lin, W. B. J. Am. Chem. Soc. 2010, 132, 922−923. (b) Barrett, S. M.; Wang, C.; Lin, W. B. J. Mater. Chem. 2012, 22, 10329−10334. (c) Wang, C.; Xie, Z. G.; deKrafft, K. E.; Lin, W. L. J. Am. Chem. Soc. 2011, 133, 13445−13454. (d) Kent, C. A.; Liu, D. M.; Ma, L. Q.; Papanikolas, J. M.; Meyer, T. J.; Lin, W. B. J. Am. Chem. Soc. 2011, 133, 12940−12943. (e) Wang, C.; Lin, W. B. J. Am. Chem. Soc. 2011, 133, 4232−4235. (f) Kent, C. A.; Mehl, B. P.; Ma, L. Q.; Papanikolas, J. M.; Meyer, T. J.; Lin, W. B. J. Am. Chem. Soc. 2010, 132, 12767−12769. (g) Wang, C.; deKrafft, K. E.; Lin, W. B. J. Am. Chem. Soc. 2012, 134, 7211−7214. (h) Kent, C. A.; Liu, D.; Meyer, T. J.; Lin, W. B. J. Am. Chem. Soc. 2012, 134, 3991− 3994. (i) Liu, D. M.; Huxford, R. C.; Lin, W. B. Angew. Chem., Int. Ed. 2011, 50, 3696−3700. (10) Kubelka, P.; Munk, F. Tech. Z. Phys. 1931, 12, 593. (11) Eskelinen, E.; Luukkanen, S.; Haukka, M.; Ahlgren, M.; Pakkanen, T. A. J. Chem. Soc., Dalton Trans. 2000, 2745−2752. (12) Spek, A. L. PLATON, A Multipurpose Crystallographic Tool; Utrecht University: Utrecht The Netherlands, 2008. (13) (a) Zhang, J.; Sun, L. X.; Xu, F.; Li, F.; Zhou, H. Y.; Liu, Y. L.; Gabelica, Z.; Schick, C. Chem. Commun. 2012, 48, 759−761. (b) Ma, M. Y.; Zacher, D.; Zhang, X. N.; Fischer, R. A.; Nolte, N. M. Cryst. Growth Des. 2011, 11, 185−189. (c) Xu, H.; Rao, X. T.; Gao, J. K.; Yu, J. C.; Wang, Z. Q.; Dou, Z. S.; Cui, Y. J.; Yang, Y.; Chen, B. L.; Qian, G. D. Chem. Commun. 2012, 48, 7377−7379. (d) Xu, H.; Liu, F.; Cui,

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Fax: (+86) 591-83730955. Tel: (+86) 591-83730955. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the National Nature Science Foundation of China (21222102, 21373220, 21301172, 21171166, and 51102231), the 973 key programs of the MOST (2010CB933501 and 2011CB935904), the Key Project of Fujian Province (2012H0045) and the One Hundred Talent Program of the Chinese Academy of Sciences.



REFERENCES

(1) (a) Murray, L. J.; Dinca, M.; Long, J. R. Chem. Soc. Rev. 2009, 38, 1294−1314. (b) Ma, L. Q.; Abney, C.; Lin, W. B. Chem. Soc. Rev. 2009, 38, 1248−1256. (c) Li, J. R.; Kuppler, R. J.; Zhou, H. C. Chem. Soc. Rev. 2009, 38, 1477−1504. (d) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Chem. Soc. Rev. 2009, 38, 1450−1459. (e) Allendorf, M. D.; Bauer, C. A.; Bhakta, R. K.; Houk, R. J. T. Chem. Soc. Rev. 2009, 38, 1330−1352. (f) Yoon, M.; Srirambalaji, R.; Kim, K. Chem. Rev. 2012, 112, 1196−1231. (g) Wu, H. H.; Gong, Q. H.; Olson, D. H.; Li, J. Chem. Rev. 2012, 112, 836− 868. (h) Wang, C.; Zhang, T.; Lin, W. B. Chem. Rev. 2012, 112, 1084− 1104. (i) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.; Bae, T. H.; Long, J. R. Chem. Rev. 2012, 112, 724−781. (j) Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D. W. Chem. Rev. 2012, 112, 782−835. (k) Li, J. R.; Sculley, J.; Zhou, H. C. Chem. Rev. 2012, 112, 869−932. (l) Kreno, L. E.; Leong, K.; Farha, O. K.; Allendorf, M.; Van Duyne, R. P.; Hupp, J. T. Chem. Rev. 2012, 112, 1105−1125. (m) Horcajada, P.; Gref, R.; Baati, T.; Allan, P. K.; Maurin, G.; Couvreur, P.; Ferey, G.; Morris, R. E.; Serre, C. Chem. Rev. 2012, 112, 1232−1268. (n) Cui, Y. J.; Yue, Y. F.; Qian, G. D.; Chen, B. L. Chem. Rev. 2012, 112, 1126−1162. (o) Li, L. N.; Ma, J. X.; Song , C.; Chen , T. L.; Sun, Z. H.; Wang, S. Y.; Luo, J. H.; Hong , M. C. Inorg. Chem. 2012, 51, 2438−2442. (p) Li, L. N.; Wang, S. Y.; Chen, T. L.; Sun, Z. H.; Luo, J. H.; Hong, M. C. Cryst. Growth Des. 2012, 12, 4109−4115. (q) Li, L. N.; Luo, J. H.; Wang, S. Y.; Sun, Z. H.; Chen, T. L.; Hong, M. C. Cryst. Growth Des. 2011, 11, 3744−3747. (r) Li, L. N.; Zhang, S. Q.; Han, L.; Sun , Z. H.; Luo, J. H.; Hong, M. C. . Cryst. Growth Des. 2013, 13, 106−110. (2) (a) Cohen, S. M. Chem. Sci. 2010, 1, 32−36. (b) Cohen, S. M. Chem. Rev. 2012, 112, 970−1000. (c) Garibay, S. J.; Wang, Z. Q.; Cohen, S. M. Inorg. Chem. 2010, 49, 8086−8091. (d) Garibay, S. J.; Wang, Z. Q.; Tanabe, K. K.; Cohen, S. M. Inorg. Chem. 2009, 48, 7341−7349. (e) Gurunatha, K. L.; Mohapatra, S.; Suchetan, P. A.; Maji, T. K. Cryst. Growth Des. 2009, 9, 3844−3847. (f) Cai, Y.; Zhang, Y. D.; Huang, Y. G.; Marder, S. R.; Walton, K. S. Cryst. Growth Des. 2012, 12, 3709−3713. (3) (a) Fu, Y. H.; Sun, D. R.; Chen, Y. J.; Huang, R. K.; Ding, Z. X.; Fu, X. Z.; Li, Z. H. Angew. Chem., Int. Ed. 2012, 51, 3364−3367. (b) Zlotea, C.; Phanon, D.; Mazaj, M.; Heurtaux, D.; Guillerm, V.; Serre, C.; Horcajada, P.; Devic, T.; Magnier, E.; Cuevas, F.; Ferey, G.; Llewellyn, P. L.; Latroche, M. Dalton Trans. 2011, 40, 4879−4881. (4) (a) Chen, B. L.; Zhao, X.; Putkham, A.; Hong, K.; Lobkovsky, E. B.; Hurtado, E. J.; Fletcher, A. J.; Thomas, K. M. J. Am. Chem. Soc. 2008, 130, 6411−6423. (b) Nguyen, H. G. T.; Weston, M. H.; Sarjeant, A. A.; Gardner, D. M.; An, Z.; Carmieli, R.; Wasielewski, M. R.; Farha, O. K.; Hupp, J. T.; Nguyen, S. T. Cryst. Growth Des. 2013, 13, 3528−3534. (c) Song, F. J.; Wang, C.; Falkowski, J. M.; Ma, L. Q.; Lin, W. B. J. Am. Chem. Soc. 2010, 132, 15390−15398. (d) Xiang, S. C.; Zhang, Z. J.; Zhao, C. G.; Hong, K. L.; Zhao, X. B.; Ding, D. R.; Xie, M. H.; Wu, C. D.; Das, M. C.; Gill, R.; Thomas, K. M.; Chen, B. L. Nat. Commun. 2011, 2, 204. (e) Hong, K.; Bak, W.; Moon, D.; Chun, H. Cryst. Growth Des. 2013, 13, 4066−4070. F

dx.doi.org/10.1021/cg401438j | Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Y. J.; Chen, B. L.; Qian, G. D. Chem. Commun. 2011, 47, 3153−3155. (e) Spokoyny, A. M.; Kim, D.; Sumrein, A.; Mirkin, C. A. Chem. Soc. Rev. 2009, 38, 1218−1227. (f) Della Rocca, J.; Liu, D. M.; Lin, W. B. Acc. Chem. Res. 2011, 44, 957−968. (g) Carne, A.; Carbonell, C.; Imaz, I.; Maspoch, D. Chem. Soc. Rev. 2011, 40, 291−305. (14) Lakowicz, J. R. Principles of fluorescence spectroscopy, 3rd ed.; Springer-Verlag: Heidelberg, 2006. (15) (a) Yang, H.; He, X. W.; Wang, F.; Kang, Y.; Zhang, J. J. Mater. Chem. 2012, 22, 21849−21851. (b) Das, M. C.; Xu, H.; Wang, Z. Y.; Srinivas, G.; Zhou, W.; Yue, Y. F.; Nesterov, V. N.; Qian, G. D.; Chen, B. L. Chem. Commun. 2011, 47, 11715−11717. (c) Chen, Y. Q.; Li, G. R.; Qu, Y. K.; Zhang, Y. H.; He, K. H.; Gao, Q.; Bu, X. H. Cryst. Growth Des. 2013, 13, 901−907. (16) (a) Furton, K. G.; Myers, L. J. Talanta. 2001, 54, 487−500. (b) Davies, A. G.; Burnett, A. D.; Fan, W. H.; Linfield, E. H.; Cunningham, J. E. Mater. Today 2008, 11, 18−26. (17) (a) Zhang, Z. J.; Xiang, S. C.; Rao, X. T.; Zheng, Q. A.; Fronczek, F. R.; Qian, G. D.; Chen, B. L. Chem. Commun. 2010, 46, 7205−7207. (b) Zhang, C. Y.; Che, Y. K.; Zhang, Z. X.; Yang, X. M.; Zang, L. Chem. Commun. 2011, 47, 2336−2338. (c) Guo, M.; Sun, Z. M. J. Mater. Chem. 2012, 22, 15939−15946. (d) Gole, B.; Bar, A. K.; Mukherjee, P. S. Chem. Commun. 2011, 47, 12137−12139. (e) Li, L. N.; Zhang, S. Q.; Xu, L. J.; Han, L.; Chen, Z. N.; Luo, J. H. Inorg. Chem. 2013, 52, 12323−12325. (18) (a) Wang, P.; Ma, J. P.; Dong, Y. B.; Huang, R. Q. J. Am. Chem. Soc. 2007, 129, 10620−10621. (b) Takashima, Y.; Martinez, V. M.; Furukawa, S.; Kondo, M.; Shimomura, S.; Uehara, H.; Nakahama, M.; Sugimoto, K.; Kitagawa, S. Nat. Commun. 2011, 2, 168. (c) Stylianou, K. C.; Heck, R.; Chong, S. Y.; Bacsa, J.; Jones, J. T. A.; Khimyak, Y. Z.; Bradshaw, D.; Rosseinsky, M. J. J. Am. Chem. Soc. 2010, 132, 4119− 4130. (d) Pramanik, S.; Zheng, C.; Zhang, X.; Emge, T. J.; Li, J. J. Am. Chem. Soc. 2011, 133, 4153−4155. (e) Lan, A. J.; Li, K. H.; Wu, H. H.; Olson, D. H.; Emge, T. J.; Ki, W.; Hong, M. C.; Li, J. Angew. Chem., Int. Ed. 2009, 48, 2334−2338. (19) (a) Zhang, C. Y.; Che, Y. K.; Yang, X. M.; Bunes, B. R.; Zang, L. Chem. Commun. 2010, 46, 5560−5562. (b) Wang, Y.; La, A.; Ding, Y.; Liu, Y. X.; Lei, Y. Adv. Funct. Mater. 2012, 22, 3547−3555. (c) Naddo, T.; Che, Y. K.; Zhang, W.; Balakrishnan, K.; Yang, X. M.; Yen, M.; Zhao, J. C.; Moore, J. S.; Zang, L. J. Am. Chem. Soc. 2007, 129, 6978− 6979. (d) Clulow, A. J.; Burn, P. L.; Meredith, P.; Shaw, P. E. J. Mater. Chem. 2012, 22, 12507−12516. (e) Che, Y. K.; Yang, X. M.; Liu, G. L.; Yu, C.; Ji, H. W.; Zuo, J. M.; Zhao, J. C.; Zang, L. J. Am. Chem. Soc. 2010, 132, 5743−5750.

G

dx.doi.org/10.1021/cg401438j | Cryst. Growth Des. XXXX, XXX, XXX−XXX