Direct Guest Exchange Induced Single-Crystal to Single-Crystal

Direct Guest Exchange Induced Single-Crystal to Single-Crystal ... Publication Date (Web): July 22, 2015 ... Crystal Growth & Design 2017 17 (11), 599...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/crystal

Direct Guest Exchange Induced Single-Crystal to Single-Crystal Transformation Accompanying Irreversible Crystal Expansion in Soft Porous Coordination Polymers Tzu-Yan Ho,†,‡ Sheng-Ming Huang,† Jing-Yun Wu,*,§ Kung-Chung Hsu,‡ and Kuang-Lieh Lu*,† †

Institute of Chemistry, Academia Sinica, Taipei 115, Taiwan Department of Chemistry, National Taiwan Normal University, Taipei 116, Taiwan § Department of Applied Chemistry, National Chi Nan University, Nantou 545, Taiwan ‡

S Supporting Information *

ABSTRACT: Two flexible porous coordination materials, [Mn(pybimc)2]·2H2O·G (G = toluene, 1tol; THF, 1thf), where pybimc = 2-(2′-pyridyl)-benzimidazole-5-carboxylate, featuring identical one-dimensional chain structure have been characterized. Guest exchange studies have exhibited that 1tol cannot be converted to 1thf through direct replacement of guest toluene molecules by THF molecules, but, of particular interest, 1thf is actually converted to 1tol and 1aromatic (where aromatic = o-, m-, p-xylene) upon the exchange of THF to toluene and other aromatic molecules, respectively. This signifies a single-crystal to single-crystal transformation accompanied irreversible crystal expansion. In-depth analyses reveal that the nature of the weak yet sufficiently strong framework−guest C−H···π interactions, rather than the guest size, observed in this system plays a key role in guiding the adsorption of liquid-phase aromatics in the soft crystalline materials.



guest C−H···π interactions, i.e., C(sp2)H···π and C(sp3)H···π, observed in this system revealed the existence of a weak yet sufficiently strong force to influence the guest inclusion behavior of the flexible porous materials (Scheme 1c).

INTRODUCTION Flexible porous coordination polymers (PCPs) have emerged as an interesting and important class of “soft” crystalline materials since they exhibit remarkable framework flexibility in responding to external stimuli.1−3 Dynamic structural distortions such as (ir)reversible crystal expansion and/or contraction upon the change (exchange or entering/exiting) of guest molecules is one of their intriguing properties (Scheme 1a,b).4,5 These types of flexible porous materials would have an open-framework with larger pores for accommodating (or adsorbing) guest molecules and usually show a closed phase with smaller pores when the guest molecules are released or replaced by relatively smaller guest molecules.3b,c,4 In recent years, considerable interest has developed regarding liquid-phase adsorption, especially nonaromatic polar solvents, such as water, alcohols (methanol, ethanol), and related compounds.3b,c,6,7 However, investigations of the adsorption of liquid-phase aromatic solvent guests are extremely rare.8 In general, noncovalent supramolecular interactions such as hydrogen bonding interactions control the collapse/recovery or single-crystal to single-crystal (SCSC) transformations of such soft crystalline materials upon desorption/adsorption.4b,6 Interestingly, the importance of weak C−H···π interactions that may guide the direct guest exchange process as well as affect the expansion and contraction of flexible porous materials has not been addressed to date. Herein we report on the liquid phase guest adsorption behavior of a MnII-based PCP. The host− © XXXX American Chemical Society



RESULTS AND DISCUSSION

Synthesis. Compound [Mn(pybimc) 2 ]·2H 2 O·toluene (1tol), obtained as yellowish crystals, was produced by the hydro(solvo)thermal reaction of Mn(NO3)2·6H2O, 2-(2′pyridyl)-benzimidazole-5-carboxylic acid (pybimcH), H2O, N,N′-dimethylformamide (DMF), and toluene at 160 °C for 72 h (Scheme 2), while crystalline products of [Mn(pybimc)2]· 2H2O·THF (1thf) were formed at room temperature by diffusing a methanol solution of Mn(NO3)2·6H2O into an aqueous solution of pybimcH and KOH using THF as a buffer layer. Description of the Crystal Structure. Single-crystal X-ray diffraction analyses of 1tol and 1thf reveal that both compounds have very similar cell parameters with identical one-dimensional coordination chain structures of [Mn(pybimc)2]. In both compounds, each pybimc ligand binds to two MnII ions in which its 2-pyridyl and benzimidazolyl functions are employed to chelate the first MnII ion with the monodentate carboxylate Received: April 23, 2015 Revised: July 21, 2015

A

DOI: 10.1021/acs.cgd.5b00565 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Scheme 1. (a−c) Illustration of Size- and InteractionControlled (Ir)Reversible Crystal Expansion and/or Contraction of Flexible Porous Materials upon Guest Exchange

Figure 1. (a) Representation of a one-dimensional water-included chain of [Mn(pybimc)2]·2H2O in 1tol and 1thf. The lattice water molecules (cyan) are strongly hydrogen-bonded to the coordination chain of [Mn(pybimc)2]. (b) Partial packing diagram of 1tol viewed down the crystallographic a axis. (c) Partial packing diagram of 1thf viewed down the crystallographic a axis.

and 1thf, respectively,9 and propagate in one direction along the c axis (Figure 1 and Figures S3 and S5 in the SI). Powder X-ray Diffraction (PXRD) and Thermogravimetric (TG) Analysis Studies. The bulky purity of 1tol and 1thf was verified via a comparison of the powder X-ray diffraction (PXRD) patterns of 1tol and 1thf with a calculated PXRD pattern obtained from single-crystal reflection data (Figures S6 and S7 in the SI). Thermogravimetric (TG) analyses indicate that guest solvent molecules are released at temperatures up to 160 °C for 1tol (found 16.8%, calcd. 16.5% based on [Mn(pybimc)2]·2H2O·0.75C7H8) and up to 170 °C for 1thf (found 19.7%, calcd. 19.2% for [Mn(pybimc)2]·3H2O· C4H8O), prior to the decomposition of the remaining solventfree frameworks at ca. 350 °C for both materials (Figure S8 and Table S1 in the SI). This conclusion is consistent with elementary analysis (EA) data. Guest Exchange Studies. For 1tol and 1thf, the identical coodination networks accomondating different lattice guest solvent molecules represent excellent candidates for direct guest exchange studies. When 1tol is immersed in THF, no significant change in the PXRD pattern is observed for the immersed product 1tol→thf (Figure S9 in the SI), implying that the guest toluene molecules in 1tol are not be replaced by THF molecules; in other words, the conversion of “1tol → 1thf” does not occur. On the other hand, when 1thf is immersed in toluene, the PXRD pattern of the immersed product 1thf→tol undergoes drastic shift. The peak (110) at 8.27° is shifted to 8.16°, the peak (111)̅ at 9.26° is shifted to 9.11°, the peak (020) at 11.99° is shifted to 11.29°, and the peak (021) at 14.70° is shifted to 14.22° (Figures 2 and S10 in the SI). The pattern for the immersed product 1thf→tol, however, is well-matched to that of 1tol, indicating that 1thf is actually converted to 1tol through the direct replacement of guest THF molecules by toluene molecules. Further, the drastic PXRD shift of the (020) and (021) reflections, which shows large 2-theta angle difference (Δ(2θ)) of −0.70 and −0.48°, respectively, demonstrates the

group bridging the second. Each MnII ion is surrounded by two sets of chelated pyridyl and benzimidazolyl nitrogens and two monodentate-bonded carboxylate oxygens from four different pybimc ligands. The distorted octahedral geometry of MnN4O2 forms a quasi trischelating coordination environment, the chirality of which naturally exhibits either a lambda (Λ) or a delta (Δ) configuration. Alternating Λ- and Δ-configured MnII ions forms an achiral meso-coordination chain through the linking of pybimc ligands. Two lattice water molecules per formula are strongly hydrogen-bonded to the coordination chain (O−H···O = 2.786(2) and 2.788(2) Å for 1tol and 2.765(5) and 2.789(7) Å for 1thf), thus forming a waterincluded chain of [Mn(pybimc)2]·2H2O (Figure 1). Through N−H···O(water) hydrogen-bonding interactions (N−H···O = 2.730(2) Å for 1tol and 2.689(8) Å for 1thf), neighboring waterincluded chains of [Mn(pybimc)2]·2H2O form a two-dimensional supramolecular sheet supported by hydrogen bonds (Figures S2 and S4 in the Supporting Information, SI). The supramolecular sheets intersect each other in an inclined fashion (the dihedral angle between the mean planes of the sheets is 73.4° for 1tol and 80.6° for 1thf), generating an overall three-dimensional entanglement (Figures S2 and S4 in the SI). The lattice solvent guests (toluene in 1tol and THF in 1thf) are located within the interchain void spaces, occupying a potential solvent accessible volume of ca. 33.0% and ca. 26.3% for 1tol Scheme 2. Self-Assembly Synthesis of 1tol and 1thf

B

DOI: 10.1021/acs.cgd.5b00565 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 4. Weak noncovalent C−H···π/O interactions between the guest solvent molecule (toluene and THF) and the coordination chains of [Mn(pybimc)2] in (a) 1tol and (b) 1thf.

the adsorption of guest solvent molecules in both 1tol and 1thf, among others (such as hydrophilic or hydrophobic interactions and the size of the adsorbate and/or adsorbent)10,11 In 1tol, a toluene molecule is trapped by the host [Mn(pybimc)2]·2H2O chains via C(sp2)−H···π (3.91−4.16 Å) interactions between the aromatic toluene hydrogens and the benzimidazolyl rings and the pyridyl rings of the pybimc ligands, and C(sp3)−H···O (3.84 Å) interactions between the aliphatic toluene hydrogens and the noncoordinated carboxylate oxygens of the pybimc ligands. On the other hand, the corresponding interactions between the guest THF molecules and the host [Mn(pybimc)2]·2H2O chains in 1thf involve C(sp3)−H···π (3.64 Å) attractions between the THF hydrogens and the benzimidazolyl rings of the pybimc ligands, and O···π (3.96 Å) interactions between the THF oxygen and the pyridyl rings of the pybimc ligands. Similar interaction-controlled cases showing distinct framework shrinking and expansion after dehydration were observed for PCPs [Cd(pzdc)(bpee)]·1.5H2O (pzdc = pyrazine-2,3dicarboxylate, bpee = trans-bis(4-pyridyl)ethylene) and [Cd(pzdc)(azpy)]·2H2O (azpy = trans-azobis(4-pyridine)), respectively.4b The C−H···O hydrogen-bonding interactions and lone pair−lone pair electronic repulsion play an important role in guiding the distinct dynamic framework distortion, i.e., shrinking for the former and expansion for the latter, for the two PCPs, respectively, after dehydration. For both PCPs, the original framework completely reformed when the guest-free materials were exposed to water vapor for several hours. During the guest exclusion/inclusion process, the crystallinity of the materials, however, is sustained, and no evidence for disintegration was observed. However, noncovalent C−H···π interactions observed in 1tol and 1thf are essentially different. The nonpolar toluene molecule has a π system and thus gives C(sp2)−H···π interactions to the coordinative host framework in 1tol, while the polar THF molecule gives C(sp3)−H···π interactions in 1tol. Since the electronegativity of the C−H carbon increases with an increase in the s character of a hybrid orbitalthat is, sp > sp2 > sp3, the polarity or the acidity of the C−H bond increases with an increase in the electronegativity of the C−H carbonthat is, C(sp)−H > C(sp2)−H > C(sp3)−H,12 allowing for an order of C−H···π interaction in strength.10 In addition, the numbers of weak C−H···π/O interactions toward the strength of host− guest interactions may also be important and thus should be taken into consideration. As a result, a toluene molecule could provide relatively stronger host−guest interactions than a THF molecule in the present systems, allowing it to have a higher affinity for the host framework. This is consistent with the

Figure 2. Partial PXRD patterns of (a) as-synthesized 1thf; (b) 1thf immersed in toluene (1thf→tol); and (c) as-synthesized 1tol.

existence of a significant, elongated b axis. This result is consistent with the crystal data for 1tol and 1thf, which show that the b axis in 1tol is 7.7% longer than that in 1thf, while the a and c axes in both 1tol and 1thf are nearly the same. However, the inter-chain spaces occupied by lattice solvent guests are aligned along the b axis; therefore, changing the guest solvent molecules without changing the crystal packing would change the inter-chain distance along the b direction and, thus, would have a more noticeable effect on the b axis. In addition to the PXRD data, the TG analysis and EA data provide further support for the conversion of 1thf to 1tol (Figure S11 and Table S1 in the Supporting Information). On the basis of the TG analysis and EA data, the formula of the immersed product 1thf→tol is estimated to be [Mn(pybimc)2]·3H2O· toluene. In other words, these results indicate that the THF molecules in 1thf are replaced by toluene molecules. The above findings clearly show that the conversion of “1thf → 1tol” is achieved, but that “1tol → 1thf” is not, as represented in Figure 3. These results unambiguously demonstrate that the

Figure 3. Representation of irreversible crystal expansion through direct guest exchange.

guest toluene molecule has a stronger affinity for the host framework of [Mn(pybimc)2]·2H2O than a THF molecule. To gain more insights into the host−guest interactions between the guest solvent molecules and the host [Mn(pybimc)2]·2H2O chains in 1tol and 1thf, a detailed analysis of the crystal structures of 1tol and 1thf was carried out. As shown in Figure 4, weak noncovalent interactions herein can be reasonably attributed to C

DOI: 10.1021/acs.cgd.5b00565 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

molecule. The proposed empirical formulas of the immersed products 1thf→o‑xyl, 1thf→m‑xyl, and 1thf→p‑xyl showing successful conversion of “1thf → 1aromatic” are [Mn(pybimc)2]·3H2O· x(aromatic) (x = 0.7−0.9), which are consistent with the EA data for each sample. According to both the TG analysis and the EA data, the immersed product 1thf→mes, in guest exchange was unsuccessful, has an empirical formula of [Mn(pybimc)2]· 2H2O·THF. As a consequence, these results suggests that the crystalline phase of [Mn(pybimc)2]·2H2O·x(solvent) has a high selectivity toward small liquid-phase aromatic molecules (such as toluene, p-, m-, and o-xylene) over THF or large aromatics (such as mesitylene).

results of guest exchange studies in which a small THF molecule could be replaced by a large toluene molecule, but the reverse was not observed. Significantly, this work presents a rare case in which a direct guest solvent exchange occurs, accompanied by an irreversible crystal expansion driven by the adsorption of liquid-phase toluene molecules where the strength of noncovalent C−H···π interactions (C(sp2)−H···π vs C(sp3)−H···π) rather than the molecular size of the guest molecules drive the process. To develop a better understanding on the adsorption behavior of 1thf toward other liquid-phase aromatic solvents, further direct guest exchange studies were performed by immersing 1thf in o-, m-, and p-xylene and mesitylene (1,3,5trimethylbenzene). The PXRD patterns of the so-obtained immersed products of 1thf in o-, m-, and p-xylene, namely, 1thf→o‑xyl, 1thf→m‑xyl, and 1thf→p‑xyl, respectively, are almost identical, in that the (020) and (021) reflections are drastically shifted from high angles of 11.99 and 14.70° in 1thf to low angles of 11.30 and 14.03° in 1thf→o‑xyl, of 11.17 and 14.12° in 1thf→m‑xyl, and of 11.27 and 14.17° in 1thf→p‑xyl (Figure 5 and



CONCLUSION In summary, we report on the successful demonstration of a novel system comprised of one-dimensional soft porous coordination materials exhibiting a SCSC transformation accompanied by crystal expansion upon the exchange of liquid-phase aromatic molecules. More importantly, this case provides a very rare example to represent the influence of the subtle difference between a C(sp2)H···π and a C(sp3)H···π interaction. This is of fundamental chemical importance.



EXPERIMENTAL SECTION

Materials and Instruments. All chemical reagents and solvents were obtained from commercial sources and used as received without further purification. 1H NMR spectra were recorded on Bruker AVA300 FT-NMR spectrometer. Chemical shifts are given in parts per million (ppm) and coupling constants are given in hertz (Hz). Powder X-ray diffraction measurements were recorded at room temperature on a Siemens D-5000 diffractometer at 40 kV, 30 mA for Cu Kα (λ = 1.5406 Å) with a step size of 0.02° in θ and a scan speed of 1 s per step size. Thermogravimetric analyses were performed under nitrogen with a PerkinElmer TGA-7 TG analyzer. Elemental analyses were performed using a PerkinElmer 2400 CHN elemental analyzer. Synthesis of 2-(2′-Pyridyl)-benzimidazole-5-carboxylic Acid (pybimcH). Pyridine-2-carbaldehyde (1.49 g, 10.0 mmol) was slowly added to a solution of 3,4-diaminobenzolic acid (1.56 g, 10.0 mmol) in glacial acetic acid (40 mL). The solution was stirred for 2 h at 80 °C. After cooling to room temperature, analytically pure brown colored products were obtained after filtering, washing with distilled H2O, and drying in air. Yield: 88% (2.11 g, 8.8 mmol). 1H NMR (CD3OD): δ 8.65 (dd, J = 4.5, 1.5 Hz, 1H), 8.28 (s, 1H), 8.22 (dd, J = 5.7, 1.2 Hz, 1H), 7.91−7.85 (m, 2H), 7.60 (d, J = 8.7 Hz, 1H), 8.65 (ddd, J = 7.5, 4.5, 1.2 Hz, 1H) ppm; MS (EI+): m/z 239.1 [M]+; elemental analysis calcd (%) for C13H9N3O2·0.3H2O: C, 63.83; H, 3.96; N, 17.18; found: C, 63.86; H, 4.20; N, 17.12. Synthesis of [Mn(pybimc)2]·2H2O·toluene (1tol). Mn(NO3)2· 6H2O (65.6 mg, 0.23 mmol), pybimcH (48.1 mg, 0.20 mmol), H2O (9 mL), DMF (3 mL), and toluene (0.5 mL) were placed in a Teflon flask which was then placed in a steel bomb. The bomb was placed in an oven maintained at a temperature of 160 °C for 72 h and then cooled to 30 °C. Good quality, deep-yellow block-shaped crystals were obtained after filtering, washing with ethanol and acetone, and drying in air. Yield 70% (47.8 mg, 7.0 × 10−2 mmol). Elemental analysis calcd (%) for C26H16MnN6O4·2H2O·0.75C7H8: C, 58.97; H, 4.12; N, 13.20; found: C, 58.80; H, 4.18; N, 13.05. Synthesis of [Mn(pybimc)2]·2H2O·THF (1thf). A solution of Mn(NO3)2·6H2O (73.5 mg, 0.26 mmol) in MeOH (5 mL) was carefully layered on top of THF (8 mL, middle), and a solution of pybimcH (24.2 mg, 0.10 mmol) and KOH (pH = 8.7) in H2O (5 mL, bottom) at room temperature. The solution was allowed to stand for approximately 2 weeks, resulting in the formation of deep-yellow block-shaped crystals in a yield of 40% (14.1 mg, 2.0 × 10−2 mmol). Elemental analysis calcd (%) for C26H16MnN6O4·3H2O·C4H8O: C, 54.80; H, 4.60; N, 12.78; found: C, 54.72; H, 4.71; N, 12.54.

Figure 5. PXRD patterns of (a) as-synthesized 1thf; (b) 1thf immersed in o-xylene (1thf→o‑xyl); (c) 1thf immersed in m-xylene (1thf→m‑xyl); (d) 1thf immersed in p-xylene (1thf→p‑xyl); and (e) 1thf immersed in mesitylene (1thf→mes).

Table S1 in the Supporting Information). This suggests that the b axis expands, while the a and c axes remain almost the same and a significant crystal expansion occurs through a direct guest exchange. However, the PXRD patterns of 1thf→o‑xyl, 1thf→m‑xyl, and 1thf→p‑xyl are analogous to that of 1tol, implying that the crystal packing and cell parameters are similar to those of 1tol. On the other hand, when 1thf is immersed in mesitylene, the PXRD patterns of the resulting product, 1thf→mes, show no significant 2-theta angle difference in the (020) (Δ(2θ) = −0.10°) and (021) (Δ(2θ) = −0.03°) reflections, implying that the guest THF molecules in 1thf were not replaced by mesitylene molecules. This can be attributed to the molecular size of the respective molecules, since the mesitylene molecule is too large to be included. These results are again supported by TG analyses and EA data of the immersed products 1thf→aromatic, where aromatic = o-xylene, 1thf→o‑xyl; m-xylene, 1thf→m‑xyl; pxylene, 1thf→p‑xyl; and mesitylene, 1thf→mes (Figure S11 and Table S1 in the Supporting Information). An analysis of the TG curves of 1thf→aromatic suggests that each THF molecule in 1thf could be replaced by approximate one aromatic o-, m-, or pxylene molecule but could not be replaced by a mesitylene D

DOI: 10.1021/acs.cgd.5b00565 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

*(K.-L.L.) Fax: +886-2-27831237; e-mail: [email protected]. tw.

Crystal Data Collection and Refinement. Data collections for 1tol and 1thf were performed at 150(2) K on a Bruker SMART CCD diffractometer equipped with a graphite monochromated Mo Kα radiation (λ = 0.71073 Å). Both of the structures were solved by direct methods (SIR9213) to gain starting models, and the structural data were refined using full-matrix least-squares treatment on F 2 (WINGX14 and SHELX-9715) with atomic coordinates and anisotropic thermal parameters for all non-hydrogen atoms. Carbon-bound hydrogen atoms were theoretically added, while oxygen-bound hydrogen atoms were first located on difference Fourier maps, and then fixed at the calculated positions and were included in the final refinement. The hydrogen atoms were refined as riding atoms derived from the parent atoms with the isotropic displacement parameters. Crystal parameters and procedural information corresponding to data collection and structure refinement were given in Table 1. CCDC

Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS We are grateful to the Academia Sinica and the Ministry of Science and Technology, Taiwan, for financial support.

Table 1. Crystallographic Data for Compounds 1tol and 1thf empirical formula Mw crystal system space group a (Å) b (Å) c (Å) β (deg) V (Å3) Z T (K) λ (Å) Dcalc (g cm−3) F000 μ (mm−1) θmin, θmax (deg) refl collected unique refl (Rint) obs refl (I > 2σ(I)) parameters R1a (I > 2σ(I)) wR2b (I > 2σ(I)) R1a (all data) wR2b (all data) GOF on F2 Δρmax, Δρmin (e Å−3) a

1tol

1thf

C33H28MnN6O6 659.55 monoclinic C2/c 18.1409(6) 16.0861(6) 12.4109(4) 124.807(2) 2973.70(18) 4 150(2) 0.71073 1.473 1364 0.502 1.86, 27.14 71032 3295 (0.0800) 2869 211 0.0337 0.0782 0.0408 0.0823 1.029 0.433, −0.319

C30H28MnN6O7 639.52 monoclinic C2/c 18.882(5) 14.835(5) 12.602(5) 124.801(5) 2898.6(17) 4 150(2) 0.71073 1.465 1324 0.514 1.90, 25.34 9854 2602 (0.5256) 986 188 0.0668 0.1289 0.1558 0.1564 0.731 0.782, −1.241

R1 = Σ∥Fo|−|Fc∥/Σ|Fo|. bwR2 = {Σ[w(F2o − F2c )2]/Σ[w(F2o )2]}1/2.

1046942 (1tol) and 1046943 (1thf) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from the Cambridge Crystallographic Data via www.ccdc. cam.ac.uk/data_request/cif.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.5b00565. 1



REFERENCES

(1) (a) Zhang, J.-P.; Liao, P.-Q.; Zhou, H.-L.; Lin, R.-B.; Chen, X.-M. Chem. Soc. Rev. 2014, 43, 5789−5814. (b) Kitagawa, S.; Kitaura, R.; Noro, S.-i. Angew. Chem., Int. Ed. 2004, 43, 2334−2375. (c) Seo, J.; Matsuda, R.; Sakamoto, H.; Bonneau, C.; Kitagawa, S. J. Am. Chem. Soc. 2009, 131, 12792−12800. (d) Uemura, K.; Kitagawa, S.; Kondo, M.; Fukui, K.; Kitaura, R.; Chang, H.-C.; Mizutani, T. Chem. - Eur. J. 2002, 8, 3586−3600. (e) Higuchi, M.; Nakamura, K.; Horike, S.; Hijikata, Y.; Yanai, N.; Fukushima, T.; Kim, J.; Kato, K.; Takata, M.; Watanabe, D.; Oshima, S.; Kitagawa, S. Angew. Chem., Int. Ed. 2012, 51, 8369−8372. (f) Férey, G.; Serre, C. Chem. Soc. Rev. 2009, 38, 1380−1399. (2) (a) Leong, W. L.; Vittal, J. J. Chem. Rev. 2011, 111, 688−764. (b) Vittal, J. J. Coord. Chem. Rev. 2007, 251, 1781−1795. (c) Kitagawa, S.; Uemura, K. Chem. Soc. Rev. 2005, 34, 109−119. (d) Tzeng, B.-C.; Chang, T.-Y.; Sheu, H.-S. Chem. - Eur. J. 2010, 16, 9990−9993. (e) Wu, J.-Y.; Ding, M.-T.; Wen, Y.-S.; Liu, Y.-H.; Lu, K.-L. Chem. Eur. J. 2009, 15, 3604−3614. (f) Ding, M.-T.; Wu, J.-Y.; Liu, Y.-H.; Lu, K.-L. Inorg. Chem. 2009, 48, 7457−7465. (g) Park, I.-H.; Lee, S. S.; Vittal, J. J. Chem. - Eur. J. 2013, 19, 2695−2702. (3) (a) Das, M. C.; Bharadwaj, P. K. J. Am. Chem. Soc. 2009, 131, 10942−10949. (b) Gurunatha, K. L.; Maji, T. K. Inorg. Chem. 2009, 48, 10886−10888. (c) Hu, S.; He, K.-H.; Zeng, M.-H.; Zou, H.-H.; Jiang, Y.-M. Inorg. Chem. 2008, 47, 5218−5224. (d) Biradha, K.; Hongo, Y.; Fujita, M. Angew. Chem., Int. Ed. 2002, 41, 3395−3398. (e) Jung, O.-S.; Kim, Y. J.; Lee, Y.-A.; Park, J. K.; Chae, H. K. J. Am. Chem. Soc. 2000, 122, 9921−9925. (f) Min, K. S.; Suh, M. P. J. Am. Chem. Soc. 2000, 122, 6834−6840. (g) Nanthamathee, C.; Ling, S.; Slater, B.; Attfield, M. P. Chem. Mater. 2015, 27, 85−95. (h) Choi, H.S.; Suh, M. P. Angew. Chem., Int. Ed. 2009, 48, 6865−6869. (4) (a) Henke, S.; Schneemann, A.; Wütscher, A.; Fischer, R. A. J. Am. Chem. Soc. 2012, 134, 9464−9474. (b) Ghosh, S. K.; Bureekaew, S.; Kitagawa, S. Angew. Chem., Int. Ed. 2008, 47, 3403−3406. (c) Maji, T. K.; Uemura, K.; Chang, H.-C.; Matsuda, R.; Kitagawa, S. Angew. Chem., Int. Ed. 2004, 43, 3269−3272. (d) Kitaura, R.; Seki, K.; Akiyama, G.; Kitagawa, S. Angew. Chem., Int. Ed. 2003, 42, 428−431. (e) Biradha, K.; Fujita, M. Angew. Chem., Int. Ed. 2002, 41, 3392− 3395. (f) Uemura, K.; Matsuda, R.; Kitagawa, S. J. Solid State Chem. 2005, 178, 2420−2429. (g) Murdock, C. R.; Lu, Z.; Jenkins, D. M. Inorg. Chem. 2013, 52, 2182−2187. (h) Medishetty, R.; Jung, D.; Song, X.; Kim, D.; Lee, S. S.; Lah, M. S.; Vittal, J. J. Inorg. Chem. 2013, 52, 2951−2957. (5) (a) Chen, C.-L.; Goforth, A. M.; Smith, M. D.; Su, C.-Y.; zur Loye, H.-C. Angew. Chem., Int. Ed. 2005, 44, 6673−6677. (b) Zhang, J.-P.; Lin, Y.-Y.; Zhang, W.-X.; Chen, X.-M. J. Am. Chem. Soc. 2005, 127, 14162−14163. (c) Dybtsev, D. N.; Chun, H.; Kim, K. Angew. Chem., Int. Ed. 2004, 43, 5033−5036. (d) Suh, M. P.; Ko, J. W.; Choi, H. J. J. Am. Chem. Soc. 2002, 124, 10976−10977. (e) Lu, J. Y.; Babb, A. M. Chem. Commun. 2002, 1340−1341. (f) Carlucci, L.; Ciani, G.; Moret, M.; Proserpio, D. M.; Rizzato, S. Angew. Chem., Int. Ed. 2000, 39, 1506−1510. (g) Kepert, C. J.; Hesek, D.; Beer, P. D.; Rosseinsky, M. J. Angew. Chem., Int. Ed. 1998, 37, 3158−3160. (6) Lu, Y.-L.; Wu, J.-Y.; Chan, M.-C.; Huang, S.-M.; Lin, C.-S.; Chiu, T.-W.; Liu, Y.-H.; Wen, Y.-S.; Ueng, C.-H.; Chin, T.-M.; Hung, C.-H.; Lu, K.-L. Inorg. Chem. 2006, 45, 2430−2437. (7) Lama, P.; Aijaz, A.; Neogi, S.; Barbour, L. J.; Bharadwaj, P. K. Cryst. Growth Des. 2010, 10, 3410−3417. (8) (a) Xu, G.; Zhang, X.; Guo, P.; Pan, C.; Zhang, H.; Wang, C. J. Am. Chem. Soc. 2010, 132, 3656−3657. (b) Liu, Q.-K.; Ma, J.-P.;

H NMR spectrum, additional crystal structures, and PXRD and TG diagrams (PDF) Crystallographic data (CIF)

AUTHOR INFORMATION

Corresponding Authors

*(J.-Y.W.) E-mail: [email protected]. E

DOI: 10.1021/acs.cgd.5b00565 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Dong, Y.-B. Chem. - Eur. J. 2009, 15, 10364−10368. (c) Cychosz, K. A.; Wong-Foy, A. G.; Matzger, A. J. J. Am. Chem. Soc. 2008, 130, 6938−6939. (d) Alaerts, L.; Maes, M.; Giebeler, L.; Jacobs, P. A.; Martens, J. A.; Denayer, J. F. M.; Kirschhock, C. E. A.; de Vos, D. E. J. Am. Chem. Soc. 2008, 130, 14170−14178. (9) Spek, A. L. PLATON, A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2005. (10) Liu, T.-F.; Lü, J.; Lin, X.; Cao, R. Chem. Commun. 2010, 46, 8439−8441. (11) Manimaran, B.; Lai, L.-J.; Thanasekaran, P.; Wu, J.-Y.; Liao, R.T.; Tseng, T.-W.; Liu, Y.-H.; Lee, G.-H.; Peng, S.-M.; Lu, K.-L. Inorg. Chem. 2006, 45, 8070−8077. (12) Douglas, B. E.; McDaniel, D. H.; Alexander, J. J. Concepts and Models of Inorganic Chemistry, 3rd ed.; John Wiley & Sons, Inc.: New York, 1994, p 85. (13) Altomare, A.; Cascarano, G.; Giacovazzo, C.; Guagliardi, A. SIR92: A Program for Crystal Structure Solution. J. Appl. Crystallogr. 1993, 26, 343−350. (14) WINGX: Farrugia, L. J. J. Appl. Crystallogr. 1999, 32, 837−838. (15) Sheldrick, G. M. SHELX-97 (including SHELXS and SHELXL); University of Göttingen: Göttingen, Germany, 1997.

F

DOI: 10.1021/acs.cgd.5b00565 Cryst. Growth Des. XXXX, XXX, XXX−XXX