Assembly of Two Flexible Metal–Organic Frameworks with Stepwise

Ran Zhao , Lei Mei , Lin Wang , Zhi-fang Chai , and Wei-qun Shi. Inorganic .... Jing Wang , Xuemin Jing , Yu Cao , Guanghua Li , Qisheng Huo , Yunling...
0 downloads 0 Views 423KB Size
Article pubs.acs.org/crystal

Assembly of Two Flexible Metal−Organic Frameworks with Stepwise Gas Adsorption and Highly Selective CO2 Adsorption Jing Wang,† Jiahuan Luo,† Jun Zhao,‡ Dong-Sheng Li,*,‡ Guanghua Li,† Qisheng Huo,† and Yunling Liu*,† †

State Key Laboratory of Inorganic Synthesis and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, P. R. China ‡ College of Mechanical & Material Engineering, Research Institute of Materials, China Three Gorges University, Yichang 443002, China S Supporting Information *

ABSTRACT: [Co3(bpydc)2(HCOO)2H2O]·2DMF (JLULiu3) and [Zn3(bpydc)2(HCOO)2]·H2O·DMF (JLU-Liu4), two kinds of chiral flexible metal−organic frameworks, have been synthesized using 2,2′-bipyridyl-5,5′-dicarboxylate ligand (H2bpydc) and trinuclear clusters. The two materials reveal ant topology and exhibit uncommon stepwise N2 and CO2 adsorption behaviors with large hysteresis loops. The threedimensional frameworks of JLU-Liu3 and JLU-Liu4 undergo structural transformation upon the removal of the guest molecules in the pores, which is indicated by the shift of powder X-ray diffraction peaks and the alteration of the peak intensity. These two flexible metal organic frameworks can be switched between the unique “closed” and “opened” forms via the removal of guest molecules. Moreover, after activation, both samples show high selective CO2 uptake over N2 and CH4.



CO2 capture,11 among which MOFs materials are promising adsorbents for CO2 fixation. So far, some flexible MOFs have been reported, but only a few exhibit selective adsorption capability. Therefore, it is imperative to develop effective ways to selectively capture and sequester CO2 to reduce the greenhouse effect in the atmosphere.12 Herein, we present two chiral flexible MOFs, JLU-Liu3 and JLU-Liu4, which are obtained by heating a mixture of Co(NO3)2·6H2O/Zn(NO3)2· 6H2O and 2,2′-bipyridyl-5,5′-dicarboxylic acid (H2bpydc)13 in DMF solution. The two materials exhibit uncommon stepwise N2 and CO2 adsorption behaviors and reveal large hysteresis loops. Also, they display high selective CO2 uptake over N2 and CH4.

INTRODUCTION Recently, metal−organic frameworks (MOFs) have attracted much attention due to their fascinating structural diversities and potential specific applications,1 such as gas adsorption and separation, catalysis, magnetism, fluorescence, drug delivery, and so on. Compared to the ordinary MOFs, MOFs with structural flexibility are able to “breathe”,2 possessing three main advantages: (1) they are amenable to structural and topological tailoring by adaption to their surroundings,3 while common MOFs materials are unstable when the environment is changed because their entire framework is linked by coordination bonds and/or other weak cooperative interactions such as H-bonding,4 π−π stacking,5 and van der Waals interaction;6 (2) they exhibit unusual stepwise and uncommon hysteretic gas adsorption behavior;7 some hysteretic gas adsorption behavior depends on the strength of the intermolecular interaction, which permits gas molecules to pass through the gate at a specific gate-opening pressure;8 (3) they display strong selective adsorption ability.9 These MOFs display particular structural flexibility on adsorption/desorption of the specific gases. There are rare works that succeed in designing and synthesizing flexible MOFs, and one of the most studied flexible MOFs is MIL-53.10 However, the cause for the structural flexibility is still far from clear, and the specific applications of such flexible MOFs is of high interest. As a greenhouse gas, CO2 has become a major worldwide environmental issue due to its rapid accumulation. To date, COFs, zeolite and MOFs have been utilized as adsorbents for © 2014 American Chemical Society



EXPERIMENTAL SECTION

Physical Measurements and Materials. All chemicals were obtained from commercial sources and used without further purification. Powder X-ray diffraction (PXRD) data were collected on a Rigaku D/max-2550 diffractometer with Cu Kα radiation (λ = 1.5418 Å). Elemental analyses were performed on a Perkin-Elmer 2400 element analyzer. The infrared (IR) spectra were recorded within the 4000−400 cm−1 region on a Nicolet Impact 410 FTIR spectrometer with KBr pellets. Thermogravimetric (TG) analyses were performed on TGA Q500 V20.10 Build 36 thermogravimetric analyzer in the temperature range 35−800 °C under air flow with a Received: January 16, 2014 Revised: March 3, 2014 Published: March 18, 2014 2375

dx.doi.org/10.1021/cg500091k | Cryst. Growth Des. 2014, 14, 2375−2380

Crystal Growth & Design

Article

heating rate of 10 °C min−1. The samples of JLU-Liu3 and JLU-Liu4 were solvent exchanged with CHCl3 and activated at 120 °C for 10 h for gas adsorption test. N2, CH4, and CO2 sorption isotherm measurements of JLU-Liu3 and JLU-Liu4 were carried out on a Micromeritics ASAP 2020 instrument. Synthesis of JLU-Liu3. A mixture of Co(NO3)2·6H2O (0.017 g, 0.06 mmol), H2bpydc (0.0025 g, 0.01 mmol), HCOOH (0.04 mL, 0.001 mmol), 4,4′-dipyridine (0.05 mL, 1.28 M in DMF), HNO3 (0.3 mL, 2.8 M in DMF), and DMF (1 mL) was added, respectively, into a 20 mL vial. And then the vial was sealed and kept at 85 °C for 12 h and 105 °C for another 24 h. Eventually, the mixture was cooled to room temperature at a rate of 15 °C/min, after which red octahedral crystals [Co3(bpydc)2(HCOO)2H2O]·2DMF were collected and dried in the air (60% yield based on Co(NO3)2·6H2O). Elemental analysis (wt %) for JLU-Liu3: calcd C 41.98, H 3.30, N 9.18; found C 42.90, H 3.82, N 9.64. FT-IR (KBr, cm−1): 3058 (w), 2942 (w), 2866 (w), 1627 (s), 1382 (s), 1295 (m), 1037 (m), 855 (s), 777 (s), 702 (m). Synthesis of JLU-Liu4. A mixture of Zn(NO3)2·6H2O (0.018 g, 0.06 mmol), H2bpydc (0.0025 g, 0.01 mmol), HCOOH (0.04 mL, 0.001 mmol), HNO3 (0.2 mL, 2.8 M in DMF), and DMF (1 mL) was added, respectively, into a 20 mL vial. And then, the vial was sealed and kept at 85 °C for 12 h. Eventually, the mixture was cooled to room temperature at a rate of 15 °C/min, after which colorless octahedral crystals [Zn3(bpydc)2(HCOO)2]·H2O·DMF were collected and dried in the air (63% yield based on Zn(NO3)2·6H2O). Elemental analysis (wt %) for JLU-Liu4: calcd C 40.42, H 2.69, N 8.13; found C 40.68, H 3.02, N 8.64. FT-IR (KBr, cm−1): 3077 (w), 2942 (w), 2875 (w), 1629 (s), 1382 (s), 1247 (m), 1038 (m), 855 (s), 777 (s), 700 (m). The phase purity of as-synthesized samples was confirmed by the evident similarities between the calculated and the experimental PXRD patterns (see Supporting Information Figure S3). The IR spectra for compounds are shown in Supporting Information Figure S4. X-ray Crystallography. Data were collected on a Bruker Apex II CCD diffractometer at 293(2) K for JLU-Liu3 and JLU-Liu4, with graphite-monochromated Mo Kα radiation (λ = 0.71073 Å). The structure was solved by direct methods and refined by full-matrix leastsquares methods with SHELXTL. All non-hydrogen atoms were easily recognized from the difference Fourier map and were refined anisotropically. Since the highly disordered guest DMF and water molecules were trapped in the channels of JLU-Liu3 and JLU-Liu4 and could not be modeled properly, there are “Alert level A” about “Check Reported Molecular Weight” and “VERY LARGE Solvent Accessible VOID(S) in Structure” in the “checkCIF/PLATON report” files for JLU-Liu3 and JLU-Liu4. The final formulas of JLU-Liu3 and JLU-Liu4 were derived from crystallographic data combined with elemental and thermogravimetric analysis data. Crystallographic data have been deposited with the Cambridge Crystallographic Data Centre, CCDC 927169 for JLU-Liu3 and 927170 for JLU-Liu4. Data can be obtained free of charge upon request at www.ccdc.cam.ac.uk/ data_request/cif. Crystal parameters and structure refinement is summarized in Table 1. Topology information for JLU-Liu3 and JLU-Liu4 were calculated by TOPOS 4.0.

Table 1. Crystal Data and Structure Refinement for Compounds JLU-Liu3 and JLU-Liu4a

a

compound

JLU-Liu3

JLU-Liu4

formula fw temp (K) crystal syst space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Z flack factor Dc (Mg/m3) μ (mm−1) F(000) reflections collected reflections unique Rint GOF on F2 R1, wR2 [I > 2σ(I)] R1, wR2 (all data)

C32H30Co3N6O15 915.41 293(2) orthorhombic C2221 20.383(4) 22.674(5) 23.215(5) 90 90 90 10729(4) 8 0.00(1) 1.133 0.971 3720 64278/12957 [R(int) = 0.0869] 1.069 0.0767, 0.1943 0.0937, 0.2062

C29H23Zn3N5O14 861.63 296(2) tetragonal P43212 15.3019(4) 15.3019(4) 23.1993(9) 90 90 90 5432.1(3) 4 0.00(1) 1.054 1.361 1736 34089/6740 [R(int) = 0.0582] 1.051 0.0583, 0.1579 0.0793, 0.1714

R1 = ∑∥Fo| − |Fc∥/∑|Fo|. wR2 = [∑[w(Fo2 − Fc2)2]/∑[w(Fo2)2]]1/2.

molecule and a terminal water molecule to form octahedral [CoN2O4] units, respectively. Like Co2 atoms, Co4 atoms are associated with four oxygen atoms from four individual ligands and two oxygen atoms from two different formic acid molecules (Figure S1b, Supporting Information). The Co−O bond distance is in the range of 2.004−2.382 Å, and the Co−N bond distance is in the range of 2.080−2.165 Å. It is noteworthy that H2bpydc ligands coordinate to Co centers in chelate (two N atoms) and bidentate bridging (two O atoms). In the structure of JLU-Liu3, two types of linear-like trinuclear clusters exist: one is formed by one Co2 and two Co1 atoms via carboxyl bridging oxygen atoms, and the other consists of one Co4 and two Co3 atoms via oxygen atoms from carboxyl bridges. These trinuclear clusters are further connected with the ligand to construct a 3D framework (Figure 1b). JLU-Liu3 displays two types of helical chains with opposite helical directions along the [001] direction (lefthanded and right-handed) (Figure S1c, Supporting Information), which are composed of vertex-sharing 4-rings propagating along the c axial direction. Both helical channels can be viewed as two independent 4-ring chains arranged in an anticlockwise/clockwise direction with dihedral angle about 93.54° and 91.34°, 96.09° and 105.95° to generate two helical chains with a pitch of 23.2 Å (Figure 1c). There are two different types of channels along the c axial direction with dimensions of 5.3 × 5.4 Å and 4.6 × 5.3 Å, respectively. To simplify the structure of JLU-Liu3, the trinuclear cluster and the ligand can be viewed as a 6-connected node and 3connected node, respectively, and then they are connected to form a (3, 6)-connected ant network (Figure S1d, Supporting Information). PLATON14 analysis reveals that the 3D porous structure is composed of a large solvent area volume of 6683.1 Å3, which represents 62.3% per unit cell volume. Crystal Structure of JLU-Liu4. Single-crystal X-ray diffraction analysis reveals that JLU-Liu4 crystallizes in the



RESULTS AND DISCUSSION Crystal Structure of JLU-Liu3. Single-crystal X-ray diffraction analysis reveals that JLU-Liu3 crystallizes in the orthorhombic crystal system with chiral space group of C2221. The asymmetric unit of JLU-Liu3 consists of four crystallographically independent Co (II) atoms, two ligands, two formic acid molecules, and one water molecule. Co1 atoms are linked with two nitrogen atoms, two oxygen atoms from three individual ligands, and two oxygen atoms from two distinct formic acid molecules, forming octahedral [CoN2O4] units. Co2 atoms are coordinated with four oxygen atoms from four individual ligands and two oxygen atoms from two different formic acid molecules (Figures 1a and S1a). Co3 atoms are connected with two nitrogen and two oxygen atoms from three individual ligands and two oxygen atoms from one formic acid 2376

dx.doi.org/10.1021/cg500091k | Cryst. Growth Des. 2014, 14, 2375−2380

Crystal Growth & Design

Article

Figure 1. A description of the structure of JLU-Liu3: (a) illustration of the ligand and trinuclear Co(II) units, viewed as 3- and 6-connected nodes; (b) view of channels with helical chains along the [001] direction (green stick: left-handed helical channels, purple stick: right-handed helical channels); (c) characteristics of helical chains.

Figure 2. A description of the structure of JLU-Liu4: (a) illustration of the trinuclear Zn(II) units and ligand; (b) view of channels with helical chains along the [001] direction (green stick: left-handed helical channels, purple stick: right-handed helical channels); (c) characteristics of helical chains; (d) ant topology displayed by tiling.

tetragonal crystal system with chiral space group of P43212. The asymmetric unit of JLU-Liu4 consists of two Zn (II) atoms, one ligand, and one formic acid molecule. Zn1 atom is coordinated with two nitrogen, two oxygen atoms from three individual ligands, and one oxygen atom from a formic acid molecule to form distorted square pyramidal [ZnN2O3] unit. Zn2 atom is connected with four oxygen atoms from four individual ligands and two oxygen atoms from two different formic acid molecules (Figures 2a and S2a). The Zn−O bond distance is in the range of 1.979−2.092 Å, and the Zn−N bond distance is in the range of 2.073−2.186 Å. The coordination of the ligands is the same as JLU-Liu3. In the structure of JLU-Liu4, each Zn2 atom is linked with two Zn1 atoms via oxygen atoms from carboxyl bridges to form linear-like trinuclear cluster building units (Figure 2a). Similar to JLU-Liu3, JLU-Liu4 also exhibits two types of helical chains

along the [001] direction (Figures 2b and S2b). These helical chains are made up of vertex-sharing 4-rings propagating along the c axial directions. Both helical channels can be viewed as two independent 4-ring chains arranged in an anticlockwise/ clockwise direction with a dihedral angle about 90.58°, 95.55°, and 83.36° to generate two helical chains with a pitch of 23.19 Å (Figure 2c). The space-filling representation of the 3D framework displays one type of channel with dimensions of 4.9 × 5.1 Å. JLU-Liu4 can also be simplified as a (3,6)-connected framework with ant topology (Figures 2d and S2c). PLATON analysis reveals that the 3D porous structure is composed of a solvent area volume of 3493.1 Å3, which represents 64.3% per unit cell volume. The framework of JLU-Liu4 is similar to that of the reported zinc compound.13d However, their synthesis, space group, free 2377

dx.doi.org/10.1021/cg500091k | Cryst. Growth Des. 2014, 14, 2375−2380

Crystal Growth & Design

Article

solvent volume are quite different. Furthermore, the reported zinc compound does not explore the gas adsorption properties. Thermogravimetric Analysis. As shown in Figure S5 (Supporting Information), the TGA curve for JLU-Liu3 shows a total weight loss of 25% between 30 and 280 °C, which corresponds to the loss of guest DMF molecules and coordinated H2O molecules. Upon further heating, a weight loss of 50% between 280 and 350 °C occurs due to the collapse of the framework (calcd: 52%). XRD indicates that over 500 °C, the final product is a pure dense cobalt oxide Co3O4 (JCPDS: 43-1003). Thermogravimetric analysis for JLU-Liu4 shows a weight loss of 22% between 30 and 260 °C, which corresponds to the loss of guest H2O and DMF molecules. Upon further heating, a weight loss of 53% between 260 and 470 °C occurs due to the collapse of the framework (calcd: 56%). XRD indicates that over 500 °C, the final product is a pure dense zinc oxide ZnO (JCPDS: 36-1451). Gas Adsorption Properties. The 3D frameworks of JLULiu3 and JLU-Liu4 underwent a structural transformation upon the removal of the guest molecules in the pores, which was indicated by the shift of the peaks and the change of their intensity in the PXRD. The PXRD patterns of activated samples JLU-Liu3 and JLU-Liu4, after immersing in CHCl3 solvent for 6 h were similar to those of the as-synthesized and simulated samples, which indicated that the structures of JLULiu3 and JLU-Liu4 can be partly restored to their original structures (Figure S3, Supporting Information). JLU-Liu3 and JLU-Liu4 were solvent exchanged with CHCl3 and activated for 10 h. Prior to the sorption studies, the surface area of the activated sample was calculated using the N2 adsorption at 77 K (Figure 3). Interestingly, the N2 adsorption isotherms of the two compounds measured up to 1 atm displayed a particular step in the adsorption (Figure 3a). These two compounds showed very little N2 adsorption at low pressure, and an abrupt increase was observed as the specific gate-opening pressure reached nearly 0.41 and 0.26 atm, which was followed by a quick increase to a maximum uptake of 325 cm3 g−1 and 306 cm3 g−1 at 1 atm. Although the structural transformation could not be proved by the single-crystal X-ray diffraction data, the adsorption behaviors confirmed the reversible framework flexibility of the two compounds: these two sorts of MOFs appeared to be the “closed” form after the removal of the guest molecules and underwent a structural transformation back to the “opened” form above the gate opening pressure.15 In addition, neither of the N2 desorption branches of the two samples traced their adsorption branch and formed large hysteresis loops. To further investigate the porosity of JLU-Liu3 and JLULiu4, the CO2 adsorption/desorption was carried out, and the isotherms at 195 K also displayed an uncommon step and hysteretic desorption. What is more, the two steps curve of the CO2 adsorption isotherms was more obvious. In the first step, the CO2 adsorption amount of JLU-Liu3 was 40 cm3 g−1 at 196 Torr, and in the second step it reached up to 337 cm3 g−1 at 760 Torr. The CO2 adsorption amount of JLU-Liu4 was 68 cm3 g−1 at 118 Torr in the first step, and it reached up to 333 cm3 g−1 at 760 Torr in the second step (Figure 3b). Similar to the N2 adsorption/desorption results, the CO2 desorption isotherms did not trace the adsorption isotherm, either, and formed a large hysteresis loop, via which the gate-opening phenomena were further verified.

Figure 3. (a) N2 adsorption and desorption isotherms of JLU-Liu3 and JLU-Liu4 at 77 K; (b) CO2 isotherms at 195 K (blue dot and solid line: JLU-Liu3, purple diamond and solid line: JLU-Liu4).

To calculate the isosteric heat of adsorption, the CO2 adsorption isotherms of the two materials were tested at 273, 283, and 298 K, but the six CO2 adsorption isotherms did not show stepwise adsorption (Figure 4a,b). It was found that the total CO2 uptake of JLU-Liu3 and JLU-Liu4 are 33 cm−3 g−1 (6.5 wt %, 273 K, 760 Torr), 30 cm−3 g−1 (5.9 wt %, 283 K, 760 Torr), 27 cm−3 g−1 (5.4 wt %, 298 K, 760 Torr), 40 cm−3 g−1 (7.8 wt %, 273 K, 760 Torr), 34 cm−3 g−1 (6.7 wt %, 283 K, 760 Torr) and 27 cm−3 g−1 (5.4 wt %, 298 K, 760 Torr) respectively. At zero-loading, adsorption enthalpy (Qst) values of JLU-Liu3 and JLU-Liu4 were 31 and 49 kJ mol−1 (Figure S6, Supporting Information). The Qst of JLU-Liu4 was similar to that of NH2-MIL-53(Al), and higher than those of ordinary MOFs materials, only a few of them display a Qst of adsorption above 40 kJ mol−1, because most of these materials include exposed cations, functional groups, or amine functionality.16 As we all know, different guest−host interactions result in different CO2 capture ability and enhanced the Qst of adsorption.1b,16 According to structure analysis, we found that there were numerous formic acids coordinated with the Co and Zn atoms in the frameworks of JLU-Liu3 and JLU-Liu4. There were two types of linear-like trinuclear clusters in the framework of JLU-Liu3: one type of trinuclear cluster was coordinated with two formic acids molecules, and the other was coordinated without formic acids molecules. In the structure of JLU-Liu4, there was only one type of linear-like trinuclear cluster, which was coordinated with two formic acids molecules. Meanwhile, the uncoordinated carbonyl groups from formic acid molecules pointed to the channels in two structures, which resulted in stronger guest−host interactions than other MOFs 2378

dx.doi.org/10.1021/cg500091k | Cryst. Growth Des. 2014, 14, 2375−2380

Crystal Growth & Design

Article

Figure 4. CO2 adsorption isotherms of JLU-Liu3 (a) and JLU-Liu4 (b) at 273 K, 283 K, and 298 K. Figure 5. CO2, N2, and CH4 adsorption isotherms for JLU-Liu3 (a) and JLU-Liu4 (b) at 298 K.

materials. As shown in Figure S7, Supporting Information, the carbonyl groups in the structure of JLU-Liu4 were twice those in JLU-Liu3, which made the guest−host interaction the JLULiu4 much stronger than JLU-Liu3. The higher Qst of the two compounds could be attributed to the strong interactions between the CO2 and the carbonyl groups, which tend to interact with CO2.17 To examine the selective separation capability of these flexible MOFs, single-component gas sorption experiments were carried out on CO2, N2, and CH4 (Figure 5). The consequences of N2 and CH4 indicated very little uptake over the entire pressure range. The selectivity ratios of the two compounds for CO2 over CH4 were 23 and 15, and CO2 over N2 were 74 and 97 (Figure S8, Supporting Information), respectively. To the best of our knowledge, MOFs with such a high CO2 selectivity over N2 were rarely reported.18 Such high selective adsorption of CO2 over N2 and CH4 gases could be explained by the effect of the specific intermolecular interaction between CO2 and the framework.19

over N2 are 74 and 97, which render these two novel kinds of MOFs promising candidates for high selective separation of CO2. Further research on the two compounds is in progress to fully understand the structural changes during the sorption process.



ASSOCIATED CONTENT

* Supporting Information S

Crystallographic data in CIF format, table for selected bonds and distances for JLU-Liu3 and JLU-Liu4, powder X-ray diffraction patterns for simulated and as-synthesized samples, IR spectra, thermogravimetric analysis, topology information and tiling for compounds JLU-Liu3 and JLU-Liu4. This information is available free of charge via the Internet at http:// pubs.acs.org/.





CONCLUSIONS In summary, we have successfully constructed two chiral flexible MOFs with ant topology. JLU-Liu3 and JLU-Liu4 MOFs can be switched between the “closed” form and “opened” form via a structural transformation. These two flexible MOFs show very little N2 adsorption at low pressure, but an abrupt increase is observed at the specific gate-opening pressure. Additionally, the CO2 adsorption isotherms also exhibit a more obvious two step curve. Both of the N2 and CO2 desorption branch do not trace the adsorption branch and form a large hysteresis loop. JLU-Liu3 and JLU-Liu4 MOFs show highly selective CO2 uptake over N2 and CH4, and the selectivity ratios for CO2 over CH4 are 23 and 15, and CO2

AUTHOR INFORMATION

Corresponding Authors

*Fax: +86-431-85168624. E-mail: [email protected]. *E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors gratefully acknowledge the financial support of the Natural Science Foundation of China (Grant Nos. 21373095 and 21373122). 2379

dx.doi.org/10.1021/cg500091k | Cryst. Growth Des. 2014, 14, 2375−2380

Crystal Growth & Design



Article

C. Angew. Chem., Int. Ed. 2012, 51, 9804. (f) Zhu, X.; Tian, C.; Mahurin, S. M.; Chai, S.-H.; Wang, C.; Brown, S.; Veith, G. M.; Luo, H.; Liu, H.; Dai, S. J. Am. Chem. Soc. 2012, 134, 10478. (g) Planas, N.; Dzubak, A. L.; Poloni, R.; Lin, L.-C.; McManus, A.; McDonald, T. M.; Neaton, J. B.; Long, J. R.; Smit, B.; Gagliardi, L. J. Am. Chem. Soc. 2013, 135, 7402. (h) Nugent, P.; Belmabkhout, Y.; Burd, S. D.; Cairns, A. J.; Luebke, R.; Forrest, K.; Pham, T.; Ma, S.; Space, B.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M. J. Nature 2013, 495, 80. (12) (a) Ochoa-Fernández, E.; Rønning, M.; Grande, T.; Chen, D. Chem. Mater. 2006, 18, 1383. (b) Du, L.; Lu, Z.; Zheng, K.; Wang, J.; Zheng, X.; Pan, Y.; You, X.; Bai, J. J. Am. Chem. Soc. 2013, 135, 562. (13) (a) Wang, C.; Xie, Z.; deKrafft, K. E.; Lin, W. J. Am. Chem. Soc. 2011, 133, 13445. (b) Gustafsson, M.; Su, J.; Yue, H.; Yao, Q.; Zou, X. Cryst. Growth Des. 2012, 12, 3243. (c) Kent, C. A.; Liu, D.; Meyer, T. J.; Lin, W. J. Am. Chem. Soc. 2012, 134, 3991. (d) Shen, L.; Gray, D.; Masel, R. I.; Girolami, G. S. CrystEngComm. 2012, 14, 5145. (e) Wang, C.; deKrafft, K. E.; Lin, W. J. Am. Chem. Soc. 2012, 134, 7211. (f) Bloch, E. D.; Britt, D.; Lee, C.; Doonan, C. J.; Uribe-Romo, F. J.; Furukawa, H.; Long, J. R.; Yaghi, O. M. J. Am. Chem. Soc. 2010, 132, 14382. (14) Spek, A. J. Appl. Crystallogr. 2003, 36, 7. (15) Mu, B.; Li, F.; Huang, Y.; Walton, K. S. J. Mater. Chem. 2012, 22, 10172. (16) 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. (17) (a) Kazarian, S. G.; Vincent, M. F.; Bright, F. V.; Liotta, C. L.; Eckert, C. A. J. Am. Chem. Soc. 1996, 118, 1729. (b) Beckman, E. J. Chem. Commun. 2004, 1885. (18) (a) An, J.; Fiorella, R. P.; Geib, S. J.; Rosi, N. L. J. Am. Chem. Soc. 2009, 131, 8401. (b) Liu, B.; Li, Y.; Hou, L.; Yang, G.; Wang, Y.-Y.; Shi, Q.-Z. J. Mater. Chem. A. 2013, 1, 6535. (19) Zheng, B.; Bai, J.; Duan, J.; Wojtas, L.; Zaworotko, M. J. J. Am. Chem. Soc. 2011, 133, 748.

REFERENCES

(1) (a) Horcajada, P.; Gref, R.; Baati, T.; Allan, P. K.; Maurin, G.; Couvreur, P.; Férey, G.; Morris, R. E.; Serre, C. Chem. Rev. 2012, 112, 1232. (b) Li, J.-R.; Sculley, J.; Zhou, H.-C. Chem. Rev. 2012, 112, 869. (c) Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Chem. Rev. 2012, 112, 673. (d) Xue, D.-X.; Cairns, A. J.; Belmabkhout, Y.; Wojtas, L.; Liu, Y.; Alkordi, M. H.; Eddaoudi, M. J. Am. Chem. Soc. 2013, 135, 7660. (e) Herm, Z. R.; Wiers, B. M.; Mason, J. A.; van Baten, J. M.; Hudson, M. R.; Zajdel, P.; Brown, C. M.; Masciocchi, N.; Krishna, R.; Long, J. R. Science 2013, 340, 960. (f) Zheng, S.-T.; Bu, J. T.; Li, Y.; Wu, T.; Zuo, F.; Feng, P.; Bu, X. J. Am. Chem. Soc. 2010, 132, 17062. (g) Li, L.; Luo, J.; Wang, S.; Sun, Z.; Chen, T.; Hong, M. Cryst. Growth Des. 2011, 11, 3744. (h) Park, T.-H.; Koh, K.; Wong-Foy, A. G.; Matzger, A. J. Cryst. Growth Des. 2011, 11, 2059. (i) Yun, R.; Lu, Z.; Pan, Y.; You, X.; Bai, J. Angew. Chem., Int. Ed. 2013, 52, 11282. (j) Pramanik, S.; Zheng, C.; Zhang, X.; Emge, T. J.; Li, J. J. Am. Chem. Soc. 2011, 133, 4153. (k) Lu, Z.; Du, L.; Tang, K.; Bai, J. Cryst. Growth Des. 2013, 13, 2252. (l) Hu, Y.-X.; Ma, H.-B.; Zheng, B.; Zhang, W.-W.; Xiang, S.; Zhai, L.; Wang, L.-F.; Chen, B.; Ren, X.-M.; Bai, J. Inorg. Chem. 2012, 51, 7066. (2) (a) Ferey, G.; Serre, C. Chem. Soc. Rev. 2009, 38, 1380. (b) Yang, C.; Wang, X.; Omary, M. A. Angew. Chem., Int. Ed. 2009, 48, 2500. (c) Zornoza, B.; Martinez-Joaristi, A.; Serra-Crespo, P.; Tellez, C.; Coronas, J.; Gascon, J.; Kapteijn, F. Chem. Commun. 2011, 47, 9522. (d) Wei, Y.-S.; Chen, K.-J.; Liao, P.-Q.; Zhu, B.-Y.; Lin, R.-B.; Zhou, H.-L.; Wang, B.-Y.; Xue, W.; Zhang, J.-P.; Chen, X.-M. Chem. Sci. 2013, 4, 1539. (e) Llewellyn, P. L.; Maurin, G.; Devic, T.; Loera-Serna, S.; Rosenbach, N.; Serre, C.; Bourrelly, S.; Horcajada, P.; Filinchuk, Y.; Férey, G. J. Am. Chem. Soc. 2008, 130, 12808. (3) (a) Horcajada, P.; Serre, C.; Maurin, G.; Ramsahye, N. A.; Balas, F.; Vallet-Regí, M. a.; Sebban, M.; Taulelle, F.; Férey, G. J. Am. Chem. Soc. 2008, 130, 6774. (b) Horike, S.; Kishida, K.; Watanabe, Y.; Inubushi, Y.; Umeyama, D.; Sugimoto, M.; Fukushima, T.; Inukai, M.; Kitagawa, S. J. Am. Chem. Soc. 2012, 134, 9852. (4) Wang, S.; Zhao, T.; Li, G.; Wojtas, L.; Huo, Q.; Eddaoudi, M.; Liu, Y. J. Am. Chem. Soc. 2010, 132, 18038. (5) Wang, C.-C.; Tseng, S.-M.; Lin, S.-Y.; Liu, F.-C.; Dai, S.-C.; Lee, G.-H.; Shih, W.-J.; Sheu, H.-S. Cryst. Growth Des. 2007, 7, 1783. (6) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Chem. Soc. Rev. 2009, 38, 1477. (7) (a) Kitaura, R.; Seki, K.; Akiyama, G.; Kitagawa, S. Angew. Chem., Int. Ed. 2003, 42, 428. (b) Horike, S.; Shimomura, S.; Kitagawa, S. Nat. Chem. 2009, 1, 695. (8) (a) Fairen-Jimenez, D.; Moggach, S. A.; Wharmby, M. T.; Wright, P. A.; Parsons, S.; Düren, T. J. Am. Chem. Soc. 2011, 133, 8900. (b) Nijem, N.; Wu, H.; Canepa, P.; Marti, A.; Balkus, K. J.; Thonhauser, T.; Li, J.; Chabal, Y. J. J. Am. Chem. Soc. 2012, 134, 15201. (9) (a) Walton, K. S.; Millward, A. R.; Dubbeldam, D.; Frost, H.; Low, J. J.; Yaghi, O. M.; Snurr, R. Q. J. Am. Chem. Soc. 2008, 130, 406. (b) Lan, A.; Li, K.; Wu, H.; Olson, D. H.; Emge, T. J.; Ki, W.; Hong, M.; Li, J. Angew. Chem., Int. Ed. 2009, 48, 2334. (c) Wu, H.; Reali, R. S.; Smith, D. A.; Trachtenberg, M. C.; Li, J. Chem.Eur. J. 2010, 16, 13951. (d) Prasad, T. K.; Suh, M. P. Chem.Eur. J. 2012, 18, 8673. (10) (a) Serre, C.; Millange, F.; Thouvenot, C.; Noguès, M.; Marsolier, G.; Louër, D.; Férey, G. J. Am. Chem. Soc. 2002, 124, 13519. (b) Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.; Henry, M.; Bataille, T.; Férey, G. Chem.Eur. J. 2004, 10, 1373. (c) Bourrelly, S.; Moulin, B.; Rivera, A.; Maurin, G.; Devautour-Vinot, S.; Serre, C.; Devic, T.; Horcajada, P.; Vimont, A.; Clet, G.; Daturi, M.; Lavalley, J.-C.; Loera-Serna, S.; Denoyel, R.; Llewellyn, P. L.; Férey, G. J. Am. Chem. Soc. 2010, 132, 9488. (11) (a) Franchi, R. S.; Harlick, P. J. E.; Sayari, A. Ind. Eng. Chem. Res. 2005, 44, 8007. (b) D’Alessandro, D. M.; Smit, B.; Long, J. R. Angew. Chem., Int. Ed. 2010, 49, 6058. (c) Liao, P.-Q.; Zhou, D.-D.; Zhu, A.X.; Jiang, L.; Lin, R.-B.; Zhang, J.-P.; Chen, X.-M. J. Am. Chem. Soc. 2012, 134, 17380. (d) Pachfule, P.; Chen, Y.; Jiang, J.; Banerjee, R. Chem.Eur. J. 2012, 18, 688. (e) Wriedt, M.; Sculley, J. P.; Yakovenko, A. A.; Ma, Y.; Halder, G. J.; Balbuena, P. B.; Zhou, H.2380

dx.doi.org/10.1021/cg500091k | Cryst. Growth Des. 2014, 14, 2375−2380