Cu(II) Coordination Polymers Based on Flexible 5,5

Apr 17, 2015 - 3 features a (4,4)-connected 3-fold interpenetrating 3D pts topology network with the Schläfli symbol (42·84)2. 4 possesses two binuc...
1 downloads 19 Views 6MB Size
Article pubs.acs.org/crystal

Co(II)/Mn(II)/Cu(II) Coordination Polymers Based on Flexible 5,5′(hexane-1,6-diyl)-bis(oxy)diisophthalic Acid: Crystal Structures, Magnetic Properties, and Catalytic Activity Lu Liu, Chao Huang, Lin Zhang, Ran Ding, Xiaonan Xue, Hongwei Hou,* and Yaoting Fan The College of Chemistry and Molecular Engineering, Zhengzhou University, Zhengzhou, Henan, 450001, P. R. China S Supporting Information *

ABSTRACT: To systematically explore the impact of coordination complexes on the synthesis of 2-imidazoline and 1,4,5,6-tetrahydropyrimidine derivatives, five Co(II)/Mn(II)/Cu(II) architectures, formulated as {[Co(L)0.5(H2O)2]·CH3OH·H2O}n (1), {[Co(L)0.5(pbib)]· 4H2O}n (2), [Mn(L)0.5(Hatz)0.5(H2O)]n (3), {[Cu(L)0.5(phen)2][Cu(L)0.5(phen)2]·0.5L·5H2O}n (4), and {[Cu(L)0.5(2,2′-bpy)(H2O)]· H2O}n (5) (H4L = 5,5′-(hexane-1,6-diyl)-bis(oxy)diisophthalic acid, pbib = 1,4-bis(imidazol-1-ylmethyl)benzene, Hatz = 1H-1,2,4-triazol-3amine, phen = 1,10-phenanthroline, 2,2′-bipy = 2,2′-bipyridine), have been designed and synthesized. 1 presents a (4,4)-connected 2D sql net with its point (Schläfli) symbol of (44·62)2, which is finally extended to a 3D supramolecular framework by π···π stacking interactions. 2 has a 3D (4,4)-connected new topology net with a point symbol of (86)2. 3 features a (4,4)-connected 3-fold interpenetrating 3D pts topology network with the Schläfli symbol (42·84)2. 4 possesses two binuclear molecules, and these adjacent binuclear units are further stretched to a 2D infinite packing structure through two distinct types of π···π stacking interactions. 5 is a 2D layer structure with the (8)(84·122) topology. The magnetic studies of 1 and 3 elucidate that both of them signify antiferromagnetic interactions. 4 and 5 have been justified to be available heterogeneous catalysts for the synthesis of 2-imidazoline and 1,4,5,6tetrahydropyrimidine derivatives.



INTRODUCTION The 2-imidazoline and 2-substituted 1,4,5,6-tetrahydropyrimidine derivatives have often been used for herculean building blocks in molecules with biological and pharmacological activity.1−3 They demonstrate significant activity against several diseases involving inflammation,4 hyperglycemia,5 hypercholesterolemia,6 hypertension,7 and cancer,8 and they can also act as antiviral,9 antidepressant,10 antibacterial,11 anthelmintics,12 and fungicides.13 The wide requirement for versatile 2-imidazoline and 1,4,5,6-tetrahydropyrimidine derivatives in diverse fields has speeded up the development of different synthetic routes. Among them, the cascade cycloaddition of a nitrile with a diamine that is mediated by various catalysts, which would be a direct approach to synthesis of imidazoline and tetrahydropyrimidine rings in conformity to the guidelines of green chemistry,1,3 has gained especial attention from the academic community. Moreover, a variety of catalysts have also been applied for this purpose, such as H2S,14,15 p-toluenesulfonic acid,14,15 sulfur,16 1,3-dibromo-5,5-dimethylhydantoin,17 thioacetamide,18 and supported 12-tungstophosphoric acid.19 Nevertheless, there are some drawbacks in most of these methods, in other words, harsh conditions such as corrosive reagents, high temperature, long reaction times, and large amounts of toxic solid supports.20 Thus, the replacement of these traditional, environmentally hazardous and corrosive © 2015 American Chemical Society

catalysts by an appropriate and friendly catalysts in order to evolve clean technology with short reaction time, high catalytic activity, recyclability, and simple workup is a prerequisite, worthwhile endeavor and justifiably also favored by chemists.21−23 The metal−organic frameworks (MOFs) with boundless network structures constructed from organic bridging ligands and inorganic connecting nodes have been appearing as very outstanding materials for separation, gas storage, sensing, heterogeneous catalysis, and drug delivery.24−59 At present, the research on MOFs as catalysts is one of the hot topics. In comparison with ordinary homogeneous catalysts, MOFs as heterogeneous catalysts will have many marvelous superiorities, comprising separation and recovery, disposal of spent catalysts, and so on.60 Practically, the “design”61−63 and fine nature of MOFs are fundamental to decide the possible catalytic properties and homologous activity.64−66 It is self-evident that the catalytic activity of MOFs would strongly rely on the characteristics of both the linking ligand and metal ion as well as additive ligands emerging in the structure. The multicarboxylate ligands, as serviceable building blocks in the Received: January 5, 2015 Revised: April 5, 2015 Published: April 17, 2015 2712

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design

Article

1739(vw), 1618(w), 1582(w), 1539(m), 1459(w), 1397(s), 1321(w), 1264(w), 1250(w), 1136(w), 1116(w), 1035(m), 993(w), 936(w), 888(w), 802(m), 784(m), 729(m), 679(w), 638(w), 560(w), 512(w), 468(w). Synthesis of {[Co(L)0.5(pbib)]·4H2O}n (2). A mixture of Co(Ac)2· 4H2O (0.0498 g, 0.1 mmol), L (0.0446 g, 0.1 mmol), pbib (0.0476 g, 0.1 mmol), NaOH (0.0160 g, 0.4 mmol), and H2O (8 mL) was placed in a 25 mL Teflon-lined stainless steel container, which was sealed and then heated at 130 °C for 72 h, followed by slow cooling to ambient temperature at a rate of 5 °C h−1. Purple crystals of 2 were obtained in a yield of 55% (based on Co). Anal. Calcd for C25H31CoN4O9 (%): C, 50.85; H, 5.29; N, 9.48. Found: C, 50.87; H, 5.28; N, 9.46. IR (KBr, cm−1): 3431(m), 3129(w), 1740(vw), 1615(w), 1562(m), 1454(w), 1400(s), 1322(w), 1268(w), 1234(w), 1105(m), 1088(m), 1046(m), 950(w), 809(m), 780(w), 736(w), 660(w), 629(w), 560(w). Synthesis of [Mn(L)0.5(Hatz)0.5(H2O)]n (3). A mixture of MnCl2· 4H2O (0.0197 g, 0.1 mmol), L (0.0223 g, 0.05 mmol), Hatz (0.0084 g, 0.1 mmol) (Hatz = 1H-1,2,4-triazol-3-amine), CH3CN (5 mL), and H2O (5 mL) was placed in a 25 mL Teflon-lined stainless steel container. The mixture was sealed and heated at 160 °C for 72 h. After cooling to ambient temperature at a rate of 5 °C h−1, colorless barshaped crystals of 3 were collected (yield: 60%, based on Mn). Anal. Calcd for C12H12MnN2O6 (%): C, 43.00; H, 3.60; N, 8.35. Found: C, 43.01; H, 3.63; N, 8.34. IR (KBr, cm−1): 3425(m), 3147(m), 2947(w), 1739(w), 1614(m), 1556(m), 1456(w), 1396(s), 1320(w), 1263(m), 1132(w), 1038(m), 927(w), 890(w), 780(s), 716(m), 634(w), 561(w), 514(w), 455(w). Synthesis of {[Cu(L)0.5(phen)2][Cu(L)0.5(phen)2]·0.5L·5H2O}n (4). A mixture of Cu(NO3)2·4H2O (0.0241 g, 0.1 mmol), L (0.0223 g, 0.05 mmol), phen (0.009 g, 0.05 mmol), and NaOH (0.15 mmol) in H2O (10 mL) was kept in a 25 mL Teflon-lined stainless steel vessel at 130 °C for 72 h. After the mixture was cooled to room temperature at a rate of 5 °C h−1, blue rod-shaped crystals suitable for X-ray diffraction were obtained with a yield of 42% (based on Cu). Anal. Calcd for C81H71Cu2N8O20 (%): C, 60.66; H, 4.46; N, 6.98. Found: C, 60.64; H, 4.47; N, 6.95. IR (KBr, cm−1): 3429(m), 3131(m), 1740(vw), 1612(m), 1567(m), 1518(w), 1401(s), 1262(w), 1224(w), 1106(w), 1047(w), 848(m), 782(m), 723(s), 561(w). Synthesis of {[Cu(L)0.5(2,2′-bpy)(H2O)]·H2O}n (5). A mixture of Cu(NO3)2·3H2O (0.0241 g, 0.1 mmol), L (0.0223 g, 0.05 mmol), 2,2′bpy (0.0078 g, 0.05 mmol), NaOH (0.1 mmol), and H2O (10 mL) was heated at 130 °C for 72 h in a 25 mL Teflon-lined stainless steel vessel. After the mixture was cooled, blue bar-shaped crystals suitable for X-ray diffraction were separated and then washed with the mother solution and air-dried (yield, 53% based on Cu). Anal. Calcd for C21H21CuN2O7 (%): C, 52.88; H, 4.43; N, 5.87. Found: C, 52.86; H, 4.45; N, 5.84. IR (KBr, cm−1): 3431(m), 3132(m), 2943(m), 2857(w), 1607(w), 1577(s), 1451(w), 1377(s), 1259(w), 1159(m), 1125(m), 1045(m), 918(w), 879(w), 776(s), 730(s), 673(w), 638(w), 553(w), 488(w). Typical Procedure for the Synthesis of 2-Imidazoline (9a− 9d) and 1,4,5,6-Tetrahydropyrimidine Derivatives (10a−10d) with Catalysts 1−5.75,76 Nitrile (1.0 mmol, 1.0 equiv, 6a−6d), ethylenediamine or 1,3-diaminopropane (0.5 mL, 7 or 8), and catalyst (0.1 mmol, 0.1 equiv based on metal ions, complexes 1−5) were mixed together and refluxed in toluene. After 4 h, the most of the cool solvent was filtrated and evaporated in vacuo; then, the residue was partitioned between chloroform (25 mL) and H2O (25 mL). The organic phase was washed with saturated NaCl aqueous, dried by Na2SO4, and then filtered and evaporated in vacuo. The crude product was purified by column chromatography on silica gel (EtOAc:MeOH, 3:1). Column chromatography was performed with silica gel (200− 300 mesh). 1H NMR spectra were recorded on 300 MHz instrument and 13C NMR spectra on 75 MHz instrument (in the Supporting Information). Crystal Data Collection and Refinement. The data of the 1−5 were collected at room temperature on a Rigaku Saturn 724 CCD diffractomer with Mo−Kα radiation (λ = 0.71073 Å). Absorption corrections were accomplished by utilizing multiscan program. The data were gathered according to Lorentz and polarization effects. The

fabrication of organic−inorganic materials with wishful topologies because of their rich coordination modes, have rooted in the hearts of chemistry workers. Meanwhile, 1,2,4-triazole derivatives unite the coordination geometry of both imidazoles and pyrazoles and display a strong and representative property of serving as bridging ligands between two metal ions. In the bridging capacity, the 1,2,4-triazole ligands exhibit a wonderful coordination diversity. This property and their powerful σdonor properties make them very tempting for the design of fresh polynuclear complexes with amazing properties.67 1,10Phenanthroline (phen) is a hydrophobic, rigid planar, electronpoor heteroaromatic system, and its nitrogen atoms are nicely placed to act cooperatively in cation binding.68 As a classic bidentate chelating ligand for transition-metal ions, it plays a momentous role in the development of coordination chemistry. 69 2,2′-Bipyridine (2,2′-bipy), as a chelating component, can form stable five-membered chelate rings upon coordination to a metal ion. The use of this kind of bridging ligands will potentially enhance metal−metal interactions, and the forming complexes will also own great stability.70 Besides, 2,2′-bipy and phen have also been proved to be extremely significant ligands for metal centers, and the acquired complexes always present absorbing chemical and physical properties.71 Given this, in this article, the reactions of Co(II)/Mn(II)/ Cu(II) salts and H4L were performed in the presence/absence of auxiliary ligands 1,4-bis(imidazol-1-ylmethyl)benzene (pbib), 1H-1,2,4-triazol-3-amine (Hatz), 1,10-phenanthroline (phen), and 2,2′-bipyridine (2,2′-bipy). Five different structural complexes, {[Co(L)0.5(H2O)2]·CH3OH·H2O}n (1), {[Co(L)0.5(pbib)]·4H2O}n (2), [Mn(L)0.5(Hatz)0.5(H2O)]n (3), {[Cu(L)0.5(phen)2][Cu(L)0.5(phen)2]·0.5L·5H2O}n (4), and {[Cu(L)0.5(2,2′-bpy)(H2O)]·H2O}n (5), have successfully been obtained to execute the experiments on the synthesis of 2-imidazoline and 1,4,5,6-tetrahydropyrimidine derivatives. The thermal stability has been surveyed on complexes 1−5. Additionally, magnetic susceptibility measurements of 1 and 3 have also been researched detailedly.



EXPERIMENTAL SECTION

Materials and Physical Measurements. All of the chemicals were commercially available except for H4L72 and pbib,73 which were prepared on the basis of a modified approach from the literature. The Fourier transform infrared (FT-IR) spectra were calendared on a Bruker-ALPHA spectrophotometer adopting KBr pellets in the scale of 400−4000 cm−1. Thermogravimetric analyses were put into effect with a Netzsch STA 449C thermal analyzer by a 10 °C min−1 heating rate in the air flow. A FLASH EA 1112 analyzer was used for measuring the content of C, H, and N. A PANalytical X’Pert PRO diffractometer using Cu Kα1 radiation was employed to collect powder X-ray diffraction patterns of the samples. Variable-temperature magnetic susceptibilities were carried out on a SQUID MPMS XL-7 instrument with phase-pure crystalline samples under the applied field of 1 kOe in the temperature region of 2−300 K. The diamagnetic corrections were conducted utilizing Pascal’s constants.74 Synthesis. Synthesis of {[Co(L)0.5(H2O)2]·CH3OH·H2O}n (1). Co(NO3)2·6H2O (0.0582 g, 0.1 mmol), L (0.0446 g, 0.1 mmol), and NaOH (0.0160 g, 0.4 mmol) were dispersed in a mixture of CH3OH and H2O in the molar ratio of 1:4 when sealed in a Teflon-lined autoclave (25 mL capacity), which was then heated at 160 °C for 72 h in a 25 mL Teflon-lined stainless steel vessel. After the reaction system was cooled to room temperature by a rate of 5 °C h−1, purple blockshaped crystals of 1 were collected. (yield, 85% based on Co). Anal. Calcd for C12H19CoO9 (%): C, 39.35; H, 5.22. Found: C, 39.34; H, 5.25. IR (KBr, cm−1): 3433(m), 3166(m), 2953(vw), 2871(vw), 2713

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design

Article

Figure 1. (a) Perspective view of the coordination environment of Co(II) ion in 1 (hydrogen atoms are omitted for clarity). Symmetry code: A = 1 − x, 1 − y, 2 − z; B = x, 1 + y, z. (b) The 2D layer structure of 1 viewed along the a-axis direction. (c) View of the 3D supramolecular structure of 1 formed by π−π stacking interactions between benzene rings of adjacent H4L. (d) The 2D network with (44·62)2 topology in 1.

Chart 1. Diverse Coordination Modes of H4L in Complexes 1−5

structures were settled by immediate methods and refined based upon F2 with a full-matrix least-squares technique employing the SHELXL97 crystallographic software package.77 All of the non-hydrogen atoms were refined by means of anisotropy. The hydrogen atoms of ligands were allocated to idealized positions using a riding model and then refined isotropically. Experimental details of the X-ray analyses for 1−5 are generalized in Table S1 (in the Supporting Information). Selected bond lengths and bond angles of 1−5 are listed in Table S2 (in the Supporting Information). Crystallographic data for 1−5 have been deposited at the Cambridge Crystallographic Data Centre with CCDC reference numbers 1041879−1041883.

Co(II) ions locates in an octahedral coordination sphere, which is anchored by two oxygens (O1 and O2) coming from one μ1η1:η1-chelate carboxylate and two oxygens (O5A and O6B) from two μ2-η1:η1-bridge carboxylates at the equatorial position, as well as two oxygens (O3 and O4) deriving from two coordinated water molecules at the axial position. The Co1−O bond lengths vary from 2.003(2) to 2.263(2) Å. The bond angles around Co1(II) ions fall in the range of 59.67(6)− 176.79(9)°. Each L4− anion acts as a μ6-bridge to link six Co(II) ions through its four carboxylate groups, and the coordination mode of (κ1-κ1)-(κ2)-(κ1-κ1)-(κ2)-μ6 for L4− is found (Mode I in Chart 1). The Co1 ion and symmetry-related Co1C (symmetry codes: C, 1 − x, 2 − y, 2 − z) ion are bridged by two μ2-η1:η1 bridging carboxylate groups to provide a binuclear [Co2(CO2)2] SBU with the nonbonding Co···Co distance of 4.232 Å. Each binuclear SBU is surrounded by four L ligands, and each L ligand is connected to four binuclear [Co2(CO2)2] units. In this



RESULTS AND DISCUSSION Crystal Structure of {[Co(L)0.5(H2O)2]·CH3OH·H2O}n (1). The results of crystallographic analysis demonstrate that the asymmetric unit of 1 possesses one crystallographically independent CoII atom, half an L ligand, two coordinated water molecules, one free CH3OH molecule, and one lattice water molecule. As illustrated in Figure 1a, the six-coordinate 2714

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design

Article

Figure 2. (a) Coordination environment of Co(II) ion in 2 with hydrogen atoms omitted for clarity. Symmetry code: A = 0.5 + x, 1.5 − y, 0.5 + z. (b) The pbib with trans-conformation links the adjoining Co(II) ions into a 1D chain. (c) View of the 2D Co (II)/L4− net fabricated by alternately left- and right-handed helical chains along the bc plane. (d) Perspective view of the 3D coordination framework accomplished by a 2D Co−L4− sheet and 1D Co−pbib chain. (e) Schematic view of the 3D topological network for 2.

groups on the opposite side of the central benzene ring. Ndonor···N−Csp3···Csp3 torsion angles of pbib-I and pbib-II are 69.936° and 108.016°, severally. The two adjacent Co(II) ions are alternately linked by pbib-I and pbib-II to give rise to a 1D infinite chain along the c axis (Figure 2b). The Co···Co separation across pbib-I and pbib-II bridges are 13.464 and 14.428 Å, separately. The four carboxyl groups of the L4− ligand all adopt monodentate bridging coordination modes connecting four Co atoms (Mode II in Chart 1). Most eye-catchingly, the medial and lateral oxygen atoms from the L4− ligand coordinate with metal ions by turns. By this way, each Co(II) ion is linked by L4− anions into a 2D layer with alternately arranged left- and right-handed helical chains along the a axis (Figure 2c). Both of the helical pitches are 16.914(3) Å, corresponding to the length of b axis. The combination of the 2D Co−L4− sheet and 1D Co−pbib chain by sharing the common Co(II) ions completes the fabrication of the 3D metal−organic framwork of 2 (Figure 2d). The O−H···O hydrogen bonding interactions between 2 and guest H2O molecules are also delineated in Figure S2 (Supporting Information) (O6−H···O7 = 2.815 Å; O6−H···O9 = 2.820 Å; O7−H···O6 = 2.815 Å; O7−H···O1 = 2.717 Å; O8−H···O9 = 2.823 Å; O8−H···O7 = 2.890 Å; O8−H···O3 = 2.880 Å). To further comprehend the structure of 2, topological analysis by reducing the complicated structure to a simple node and linker was referred. Each Co(II) ion is attached to four equivalent nodes through two L4− ligands and two pbib ligands, and each L4− anion links four Co(II) atoms. Thus, the Co(II) ion and L4− anion can be regarded as 4-connecting nodes and 4-connecting linkers, separately. Moreover, the pbib-I and pbibII can be assigned as linear linkers, singly. On the basis of the simplification principle,79 the 3D framework of 2 can be described as a (4,4)-connected new topology net with a point symbol of (86)2 (Figure 2e). Crystal Structure of [Mn(L)0.5(Hatz)0.5(H2O)]n (3). The Xray structural determination manifests that complex 3 is a 3-fold

way, the L anions link the binuclear units to create a 2D layered framework lying on the bc plane (Figure 1b). The adjacent 2D layers are finally extended to a 3D supramolecular framework by π···π stacking interactions between the two aromatic phenyl rings with a centroid−centroid distance of about 3.670 Å (Figure 1c). The O−H···O hydrogen bonding interactions between 1 and guest CH3OH, H2O molecules are also portrayed in Figure S1 (Supporting Information) (O3−H··· O9 = 2.586 Å; O3−H···O8 = 2.699 Å; O4−H···O1 = 2.752 Å; O4−H···O6 = 2.907 Å; O8−H···O9 = 2.190 Å; O4−H···O8 = 3.073 Å; O3−H···O7 = 3.032 Å). From the standpoint of structural topology, the L anions and the binuclear [Co2(CO2)2] SBU can be seen as 4-connected nodes, respectively. Thus, the whole 2D framework of 1 can be represented as a (4,4)-connected net with its point (Schläfli) symbol of (44·62)2, which is referred to as sql topology (Figure 1d). Crystal Structure of {[Co(L)0.5(pbib)]·4H2O}n (2). When pbib was introduced into the reaction system, complex 2 was obtained. Single-crystal X-ray diffraction analysis declares that 2 crystallizes in a monoclinic system with the P21/c space group. As portrayed in Figure 2a, the structure of 2 contains one crystallographically unique Co ion, half an L ligand, one pbib ligand, and four lattice water molecules. Each distorted tetrahedral Co1 center is coordinated by two carboxylate oxygen atoms (O2 and O4A) from two different L ligands (Co1−O2 = 1.989(3) Å, Co1−O4A = 2.016(3) Å) and two nitrogen donors (N1 and N3) from two distinct pbib ligands (Co1−N1 = 2.024(4) Å, Co1−N3 = 2.033(4) Å). The distortion of the tetrahedron can be expressed by the calculated value of the τ4 parameter introduced by Houser78 to depict the geometry of a four-coordinate metal system, which is 0.91 for Co1 (for ideal tetrahedron τ4 = 1). The bond angles around Co atoms range from 106.64(2)° to 116.92(2)°. In 2, there are two kinds of independent pbib, both of which exhibit a symmetrical trans-conformation, with both imidazole 2715

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design

Article

Figure 3. (a) Coordination environment around the Mn(II) centers in 3. Hydrogen atoms and solvent molecules are omitted for clarity. Symmetry code: A = 0.5 − x, 1.5 − y, 1 − z; B = −0.5 + x, 1.5 − y, −0.5 + z. (b) The 2D sheet structure constructed from Mn(II) centers and parts of L4− anions, which contains the binuclear [Mn2(CO2)2] unit. (c) Schematic view of the 3D architecture of 3. (d) Schematic representation of the 3-fold interpenetrated topology nets for 3.

Figure 4. (a) Perspective view of the coordination environment of Cu(II) ion in 4 (hydrogen atoms are omitted for clarity). (b) L4−-I and four phen ligands connect two Cu1(II) ions to form a binuclear unit I; L4−-II and four phen ligands join with two Cu2(II) ions to yield a binuclear unit II. (c) View of the 1D infinite chain formed by π−π stacking interactions of the type A. (d) The π−π stacking interactions of the type A and type B interconnect separate binuclear structures, affording a 2D infinite packing structure.

interpenetrated 3D network. The asymmetric unit of 3 possesses one crystallographically independent MnII atom, half an L ligand, half a Hatz ligand, and one coordinated water molecule. In accordance with Figure 3a, the Mn1 ion is in a five-coordinated mode (MnO4N), in which Mn1 is equatorially bonded to carboxylic oxygen atoms (O2, O3A, and O4B) from three different L4− ligands (Mn1−O2 = 2.066(3) Å, Mn1−O3A

= 2.070(3) Å, Mn1−O4B = 2.077(3) Å). The two axial sites on the metal are occupied by one N atom from one pbbm ligand (Mn1−N1 = 2.246(4) Å) and one carboxyl O from one coordinated water molecule (Mn1−O5 = 2.187(3) Å). The structural index parameter (τ)80 is 0.58, indicating that the geometry around Mn(II) is best described as a trigonalbipyramidal distorted square-based pyramid (TBPDSBP).80,81 2716

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design

Article

Figure 5. (a) Coordination environment of Cu(II) ion in 5 with hydrogen atoms omitted for clarity. Symmetry code: A = 2.5 − x, 0.5 + y, 0.5 − z. (b) The 2D layer structure of 5 fabricated from L4− and Cu(II) ion. (c) Schematic description of the 2D network with (8)(84·122) topology for 5.

Depicted in Figure 4a, the Cu1 center is five-coordinated by one oxygen atom (O2) belonging to one L4− ligands as well as four nitrogen atoms (N1, N2, N3, N4) from two phen ligands. The structural index parameter (τ)82−85 is 0.48, indicating that the geometry around Cu(II) is best described as a trigonalbipyramidal distorted square-based pyramid (TBPDSBP).82−86 The distance of Cu1−O2 is 1.943(2) Å, and Cu1−N bond distances lie in the range of 1.994(2)−2.211(2) Å. Although the Cu2 center is also five-coordinated, it takes a distorted squarepyramidal geometry (τ = 0.29).87−89 The distances of Cu2−O7 are 1.960(2) Å, while the Cu2−N bond lengths range from 1.994(2) to 2.211(2) Å. Ligands L4− takes two different coordination modes, and there are only two carboxylate groups participating in the coordination with Cu(II) ions. Two carboxylate groups of the first type (L4−-I) both adopt monodentate bridging modes (Mode IV in Chart 1); however, two benzene rings are not on the same plane owing to the tortuosity of the flexible long chain comprising a −(CH2)6− spacer and two ether oxygen atoms. The two carboxylate groups of the second type (L4−-II) both adopt monodentate bridging modes (Mode V in Chart 1), but two benzene rings are nearly on the same plane. L4−-I and four phen ligands connect two Cu1(II) ions to form a binuclear unit I (cyan molecules); meanwhile, L4−-II and four phen ligands join with two Cu2(II) ions to yield a binuclear unit II (pink molecules) (Figure 4b). Between the adjacent binuclear units I and II, an arresting feature is that both types of π···π interactions are observed. One type (named as A) is π···π stacking interactions between the phen ring consisting of N1 and N2 in the binuclear unit I as well as a benzene ring originating from the L ligand in the binuclear unit II, with a center-to-center separation of 3.616 Å (red bars). Another type (named as B) is π···π stacking interactions between the phen ring making up of N3 and N4 in the binuclear unit I as well as a phen ring composed of N5 and N6 in the binuclear unit II, with two kinds of center-to-center separations of 3.583 Å (blue bars) and 3.699 Å (yellow bars). These π···π stacking interactions of

The bond angles around Mn1 range from 83.03(1)° to 167.29(1)°. In this structure, the L4− ligand exhibits one interesting binding mode: (κ1-κ1)-(κ1)-(κ1-κ1)-(κ1)-μ6 (Mode III in Chart 1), in which four carboxylic groups serve as syn− anti-μ2-η1:η1, μ1-η1:η0, syn−anti-μ2-η1:η1, and μ1-η1:η0 modes in a clockwise direction, severally, to bridge six Mn(II) ions. The Mn1 ion and symmetry-related Mn1C (symmetry codes: C, −x, −y, 0.5 − z) ion are bridged by two μ2-η1:η1 bridging carboxylate groups to furnish a binuclear [Mn2(CO2)2] unit (Figure 3b). Within the bimetallic subunit, the intermetallic distance of Mn···Mn is about 3.625 Å. Then, the binuclear [Mn2(CO2)2] units are further linked by L ligands to yield an intriguing 3D open framework that possesses largely cavities approximately 10.49 Å × 13.59 Å in size (Figure 3c). Noticeably, the Hatz ligand does not play an indispensable role in the construction of the final 3D architecture. The O− H···O hydrogen bonding interactions between 3 and guest H2O molecules are also depicted in Figure S3 (Supporting Information) (O5−H···O6 = 2.812 Å; O5−H···O1 = 2.725 Å). Better insight into the tanglesome 3D architecture can be fulfilled by topology analysis. The binuclear [Mn2(CO2)2] unit links four L ligands, and each L4− ligand connects four binuclear Mn units. Thereby, the binuclear Mn units can be regarded as 4-connected nodes, while the independent L4− ligands act as 4-connected linkers. As a result, the whole structure can be rationalized as a (4,4)-connected pts topology network with the Schläfli symbol (42·84)2. The three identical pts nets are further interpenetrated in a 3-fold mode (Figure 3d). Crystal Structure of {[Cu(L)0.5(phen)2][Cu(L)0.5(phen)2]· 0.5L·5H2O}n (4). Crystallographic analysis makes clear that 4 crystallizes in the triclinic space group P1̅ and exhibits a binuclear structure owing to the influence of chelating phen. The asymmetric unit of complex 4 consists of two independent [Cu(L)0.5(phen)2] units, half an L ligand, and five lattice water molecules. Each independent [Cu(L)0.5(phen)2] unit contains one Cu(II) atom, half an L ligand, and two phen ligands. 2717

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design

Article

Figure 6. χmT vs T plot and χm−1 vs T plot for 1 and 3 at H = 1 kOe from 2 to 300 K.

dihedral angles (θ) between the two phenyl rings of L4− is almost all 0°, except for that of L4−-I in 4. The six carbon aliphatic backbone and two ether oxygen atoms are all in a plane. For 1, each L4− anion coordinates to metal ions in a (κ1κ1)-(κ2)-(κ1-κ1)-(κ2)-μ6 coordination mode (Mode I of Chart 1), linking the binuclear units [Co2(CO2)2] to build a 2D layered framework. In 2, the four carboxyl groups of the L4− ligand all adopt monodentate bridging coordination modes connecting to metal ions (Mode II of Chart 1), leading to a 2D layer with alternately arranged left- and right-handed helical chains. When pbib is introduced into the synthetic procedure of 2, the 1D Co(II)/pbib infinite chain as pillars supports the 2D Co(II)/L4− layer, forming a 3D metal−organic framwork. For 3, the L4− ligand coordinates with the central metals in a (κ1κ1)-(κ1)-(κ1-κ1)-(κ1)-μ6 binding mode (Mode III of Chart 1), yielding a 3-fold interpenetrated 3D framework. Only two carboxylate groups of L4− in 4 participate in the coordination with Cu(II) ions, showing two types of discriminating coordination fashions (L4−-I and L4−-II) (Modes IV and V of Chart 1). L4−-I and L4−-II join with Cu(II) ions to yield two different binuclear units in the exsitence of phen ligands. In 5, four carboxylate groups of L4− all take part in the coordination with Cu(II) ions, and they all exhibit monodentate bridging modes (Mode VI of Chart 1), which differs from that of 2, affording a 2D infinite sheetlike structure. PXRD Patterns and Thermogravimetric Analyses. The PXRD patterns of complexes 1−5 are shown in Figure S6 in the Supporting Information. The PXRD patterns of the five complexes determined by experiment are in line with the simulated ones of single crystals, which indicate that each of the five complexes is pure phase. To appraise the stability of the coordination architectures, thermogravimetric analyses (TGA) of 1−5 were carried out (Figure S7, Supporting Information). For 1, the first step weight loss, attributed to the gradual release of one lattice water molecule and one isolated CH3OH molecule, is observed in the range of 80−138 °C (obsd, 14.11%; calcd, 13.65%). The second step weight loss from 340 to 521 °C corresponds to the decomposition of two coordinated water molecules and L4−. The TGA curve of complex 2 exhibits that it loses weight from 42 to 177 °C, corresponding to the decomposition of four coordinated water molecules (obsd, 12.31%; calcd, 12.20%). The decomposition of L4− and pbib is observed in the range of 324−458 °C. As for complex 3, the initial weight losses in the range of 149−248 °C correspond to the loss of one coordinated water molecule (obsd, 5.54%; calcd, 5.37%). The further weight losses in the

the type A lead to the formation of a 1D infinite chain (Figure 4c). To our surprise, these π···π stacking interactions of the type A and type B interconnect separate binuclear structures, resulting in a 2D infinite packing structure (Figure 4d). The O−H···O hydrogen bonding interactions between 4 and guest H2O molecules are also exhibited in Figure S4 (Supporting Information) (O13−H···O8 = 2.474 Å; O16−H···O19 = 2.520 Å; O16−H···O1 = 2.617 Å; O17−H···O14 = 2.829 Å; O17− H···O15 = 3.035 Å; O18−H···O1 = 3.026 Å; O20−H···O11 = 2.658 Å). Crystal Structure of {[Cu(L)0.5(2,2′-bpy)(H2O)]·H2O}n (5). The crystal structure determination uncovers that complex 5 crystallizes in the monoclinic crystal system with space group of P21/c. As depicted in Figure 5a, the asymmetric unit consists of one independent CuII atom, half an L ligand, one 2,2′-bipy ligand, one associated water molecule, and one lattice water molecule. The Cu(II) atom shows five coordination with ligation from two carboxylate oxygen atoms (O2 and O4A) belonging to two L ligands (Cu1−O2 = 1.945(2) Å, Cu1−O4A = 1.961(2) Å), one oxygen atom (O5) of one coordinated H2O molecule (Cu1−O5 = 2.371(3) Å), and two nitrogen donors (N1 and N2) originating from one 2,2′-bipy ligand (Cu1−N1 = 1.981(3) Å, Cu1−N2 = 2.021(3) Å). This gives rise to a distorted square-pyramidal coordination geometry (τ5 = 0.13). For 5, four carboxylate groups of the L4− anion all show the monodentate bridging modes (Mode VI in Chart 1), but different from that of the L4− anion in 2. The reason is that the inboard oxygen atom from the L4− anion bonds with metal ions in 5. The Cu(II) atoms are connected by L4− anions to afford a 2D infinite sheetlike structure propagating in the crystallographic c direction (Figure 5b). The O−H···O hydrogen bonding interactions between 5 and guest H2O molecules are also revealed in Figure S5 (Supporting Information) (O5−H··· O1 = 2.626 Å; O5−H···O2 = 3.121 Å; O7−H···O2 = 3.207 Å; O7−H···O3 = 2.906 Å). Topological analysis was executed to have an insight into the structure of 5. If the Cu(II) atoms are regarded as 2-connected nodes and L4− anions are defined as 4-connected linkers, the 2D framework of 5 can be described as (8)(84·122) topology (Figure 5c). Effect of Diverse Coordination Modes of H4L and NDonor Coligands on the Structures of Complexes 1−5. From the structure description above, we find that the L4− anion exhibits diverse coordination modes in complexes 1−5 (Chart 1), connecting to six (complexes 1 and 3) or four (complexes 2 and 5) or two (complex 4) metal ions. The 2718

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design

Article

range of 320−521 °C represent the decomposition of the whole framework. The TGA curve of 4 shows the first step weight loss (obsd: 5.25%) at 44−141 °C, corresponding to the loss of five lattice water molecules (calcd: 5.62%). The removal of the organic components occurs in the range of 250−565 °C. The TGA data of complex 5 display a two-step weight loss. The first weight loss of 3.60% is observed in the temperature range of 74−152 °C, which corresponds to the loss of one lattice water molecule (calcd: 3.77%). The second step weight loss from 265−513 °C corresponds to the loss of one coordinated water molecule and the decomposition of the coordination framework. Magnetic Properties Measurements. The magnetic properties of 1 and 3 were investigated over the temperature range of 2−300 K. The χmT value of polymer 1 at 300 K is 5.85 cm3 K mol−1, which is significantly larger than the spin only value of 3.75 cm3 K mol−1 for two uncoupled spins Co(II) ions (g = 2.0 and S = 3/2). As the temperature decreases, the value of χmT continuously decreases and reaches a minimum of 2.62 cm3 K mol−1 at 2 K (Figure 6, left). The feature might be caused by antiferromagnetic exchange interactions as well as the effect of spin−orbit coupling. The plot of χm−1 versus T for 1 is consistent with the Curie−Weiss law in the temperature range of 20−300 K, with C = 6.02 cm3 K mol−1 and θ = −12.19 K. In this case, in order to obtain an estimate of the strength of the antiferromagnetic exchange interaction, a simple phenomenological equation was carried out.90−92

B = 1 + 3e 2J / kT + 5e 6J / kT + 7e12J / kT + 9e 20J / kT + 11e 30J / kT χm =

1 − χd (2zJ ′/Ng 2β 2)

Best fitting results lead to the following parameters for 3: J = −1.01 cm−1, zJ′ = −0.02 cm−1, g = 2.02, and R = 6.03 × 10−3. (The agreement factor R = ∑[(χm ) obs − (χm ) calcd] 2 / ∑[(χm)obs]2). The Catalytic Capacity. The catalytic reaction of cascade cycloaddition utilizing diamines and aromatic nitriles is one of the most efficient methods toward the synthesis of 2imidazoline and 2-substituted 1,4,5,6-tetrahydropyrimidine derivatives. The employment of the complexes as heterogeneous catalysts in the liquid phase is highly desirable for the cascade cycloaddition reaction, because the usage of orthodox homogeneous catalysts for this reaction would spark off several dilemmas, such as trouble in separation and recovery and disposal of spent catalysts. In light of this inspiration, the cascade cycloaddition reaction, in which the aforesaid complexes 1−5 were used as prospective heterogeneous catalysts, has been investigated. What is frustrating is that complexes 1−3 show quite low catalytic capacities. Only complexes 4 and 5 are effective catalysts for the cascade cyclozation of aromatic nitriles (6a−6c) with ethane diamine to produce 2-imidazolines (9a−9c). The catalytic results of 4 and 5 are listed in Table 1. Homoplastically, 4 and 5 are also

χm T = A exp( −E1/kT ) + B exp(−E2 /kT )

Table 1. Synthesis of 2-Imidazolines with Different Catalysts

In this equation, A + B equals the Curie constant, and E1 and E2 stand for the activation energies corresponding to the spin− orbit coupling and the antiferromagnetic exchange interaction, respectively. The best fit of the experimental data gives that A + B = 6.06 cm3 K mol−1, E1/k = 34.67 K, −E2/k = −0.87 K, and R = 5.42 × 10−5, respectively. According to the Ising chain approximation [χmT ∝ exp(+J/2kT)], the exchange interaction J is −1.21 cm−1, and the value induces that the antiferromagnetic exchange interaction is very weak. For complex 3, the experimental χmT value at 300 K is 8.71 cm3 K mol−1, which is slightly lower than that anticipated for a noninteracting pair of S = 5/2 ions (8.75 cm3 K mol−1 per Mn2) (Figure 6, right). As the temperature lowers to 2 K, the χmT value decreases first slowly and then rapidly. This behavior reveals that antiferromagnetic interactions are operative in 3. The temperature dependence of the reciprocal molar magnetic susceptibility data (1/χm) as a function of temperature is linear above 10 K, obeying the Curie−Weiss law of 1/χm = (T − θ)/C with the Curie constant C = 9.28 cm3 mol−1 K and the Weiss constant θ = −14.74 K over the temperature range of 10−300 K. The negative θ value also evinces the presence of antiferromagnetic interactions in 3. In 3, the main magnetic interactions can be considered to occur in the binuclear [Mn2(CO2)2] units linked by two μ2η1:η1-carboxylate groups. In order to quantitatively evaluate the magnetic interactions in 3, for similar binuclear Mn(II) complexes, the following equation93 is induced from Hamiltonian Ĥ = −2JŜ1·Ŝ2:94 χd =

χd

entry 1 2 3 4 5 6 7 8

Ar

product

catalyst

yield (%)

Ph

9a

4-NCC6H4

9b

4-pyridyl

9c

1-naphthyl

9d

4 5 4 5 4 5 4 5

28 55 40 71 50 83 10 26

effective for the synthesis of 2-substituted 1,4,5,6-tetrahydropyrimidines (Table 2). Specifically, 4 is only binuclear structure. In its structure, there are two kinds of separate metal ions, namely, Cu1(II) ion and Cu2(II) ion. Both of them are fivecoordinated. Within this catalytic system, 4 can act as a resultful catalyst. This is more possibly attributed to the exsitence of unsaturated coordination sites. As for 5, the presence of coordinated water molecules is a crucial factor. Because water molecules are good leaving groups, they may depart during the catalytic process, leaving unsaturated copper sites (fourcoordinated) to be exposed to aromatic nitriles, which will launch this reaction more easily.76 Besides, 5 is a plicated 2D layer, and the metal center always appears on the raised position. The special characteristics of 5 also will be favorable to the contact of substrates with the catalytic active sites in the

2Ng 2β 2 A kT B

A = e 2J / kT + 5e 6J / kT + 14e12J / kT + 30e 20J / kT + 55e 30J / kT 2719

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design

Article

mechanism of the synthetic reaction of 2-imidazoline and 1,4,5,6-tetrahydropyrimidine derivatives using 4 or 5 as catalyst provisionally (Scheme 1). First, the aromatic nitrile is activated

Table 2. Synthesis of 1,4,5,6-Tetrahydropyrimidines with Different Catalysts

Scheme 1. Suggested Mechanism for the Synthesis of 2Imidizoline and 1,4,5,6-Tetrahydropyrimidine Derivatives Catalyzed by Cu-Based Complexes

entry 1 2 3 4 5 6 7 8

product

catalyst

yield (%)

Ph

Ar

10a

4-NCC6H4

10b

4-pyridyl

10c

1-naphthyl

10d

4 5 4 5 4 5 4 5

34 40 48 65 35 54 10 22

catalytic process, which further could expedite the reaction. Evidently, the catalytic activity of 5 is higher than that of 4, giving a conversion of 55, 71, and 83% for 9a−9c, respectively. This situation may result from the particular construction of 5 that is superior to the structure of 4. Although there also exist coordinated water molecules in 1 and 3, their catalytic capacities are extremely low. We know that 1 is a flat 2D layer, and all of the atoms were almost in the same plane. Even if the coordinated water molecules of the metal center in 1 go away in the catalytic process, its catalytic activity will also be low because the substrates may not easily find the catalytic active sites. As for 3, it reveals a 3-fold interpenetrated 3D network. Such a interpenetrating framework can gravely decrease or even dispell the interior cavity of the MOFs, which will be adverse to catalytic reaction by reason for its incompetence to convey the substrates through the small channels. 2 displays a close-knit 3D framework with alternately arranged left- and right-handed helical chains. Its feature could minish the open channel sizes and deter the heterogeneous catalytic reaction. It seems as if the unsaturated coordination sites of Co(II) ions in 2 play no role in the reaction. We speculate that the effect of helical structures could occupy one dominating status. Viewed from another perspective, only Cu(II) complexes have an advantage in catalytic reactions, whereas the other two Co(II) complexes and one Mn(II) complex do not accelerate the reaction. It is clear that the metal centers may also have a momentous impact on the catalytic activities, in addition to the coordination environments of center metal ions and the different structures of complexes. When 1-naphthyl nitrile (entries 7−8 in Tables 1 and 2) was used as original material and 5 serves as heterogeneous catalyst, the yields of 9d and 10d are a little low. It may be due to the fact that the size of the substrate 1-naphthyl nitrile is a bit big, which probably goes against the reaction. It also further illustrates that the substrate size has a certain influence on the catalyzed reaction. Furthermore, taking the reaction of 6c and 7 (entry 6 in Table 1), for example, 4 and 5 as heterogeneous catalysts could be recycled and reused at least two times, and the yield of 9c somewhat descends. The excellent match between the PXRD of the 5 recovered from the catalytic reaction by centrifugation and that of the primordial solid of 5 proved that 5 is stable under the catalytic condition (Figure S8, Supporting Information). Meantime, we also speculate the

by the catalyst (4 or 5) to give I. Then, the diamines attacks I to produce II. Cyclization of II affords the final product. For our catalytic experiment, although the catalytic activity is not better or not poorer than reported ones,14−19 4 and 5 as heterogeneous catalysts have some of their own particular strengths, such as reusability, easy separation, environmental friendliness.



CONCLUSIONS In brief, five Co(II)/Mn(II)/Cu(II) coordination polymers have been prepared under (solvo)hydrothermal conditions, by using H4L with the introduction of rationally selected N-donor ancillary ligands. They exhibit different frameworks with beautiful topologies. The magnetic studies on 1 and 3 illuminate that both of them reveal antiferromagnetic interactions. 4 and 5 have turned out to be practicable heterogeneous catalysts for the synthesis of 2-imidazoline and 1,4,5,6-tetrahydropyrimidine derivatives.



ASSOCIATED CONTENT

S Supporting Information *

X-ray crystallographic data, selected bond lengths and bond angles, powder X-ray patterns, TGA curves for 1−5, and 1H NMR and 13C NMR data for 9a−9d and 10a−10d. The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.5b00016.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Fax: (86)371-67761744 (H.H.). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS

This work was financially supported by the National Natural Science Foundation (No. 21371155) and the Research Fund for the Doctoral Program of Higher Education of China (20124101110002). 2720

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design



Article

(34) Spokoyny, A.-M.; Kim, D.; Sumrein, A.; Mirkin, C.-A. Chem. Soc. Rev. 2009, 38, 1218−1227. (35) Uemura, T.; Yanai, N.; Kitagawa, S. Chem. Soc. Rev. 2009, 38, 1228−1236. (36) Wang, Z.; Cohen, S. M. Chem. Soc. Rev. 2009, 38, 1315−1329. (37) Ma, L.; Abney, C.; Lin, W. Chem. Soc. Rev. 2009, 38, 1248− 1256. (38) Murray, L.-J.; Dincâ, M.; Long, J.-R. Chem. Soc. Rev. 2009, 38, 1294−1314. (39) Farrusseng, D.; Aguado, S.; Pinel, C. Angew. Chem., Int. Ed. 2009, 48, 7502−7513. (40) Lee, J.; Farha, O.-K.; Roberts, J.; Scheidt, K.-A.; Nguyen, S.-T.; Hupp, J.-T. Chem. Soc. Rev. 2009, 38, 1450−1459. (41) Habib, H.-A.; Sanchiz, J.; Janiak, C. Dalton Trans. 2008, 1734− 1744. (42) Wisser, B.; Lu, Y.-R.; Janiak, C. Z. Anorg. Allg. Chem. 2007, 633, 1189−1192. (43) Corma, A.; García, H.; Llabrés i Xamena, F.-X. Chem. Rev. 2010, 110, 4606−4655. (44) Hasegawa, S.; Horike, S.; Matsuda, R.; Furukawa, S.; Mochizuki, K.; Kinoshita, Y.; Kitagawa, S. J. Am. Chem. Soc. 2007, 129, 2607− 2614. (45) Chen, Y.-Q.; Qu, Y.-K.; Li, G.-R.; Zhuang, Z.-Z.; Chang, Z.; Hu, T.-L.; Xu, J.; Bu, X.-H. Inorg. Chem. 2014, 53, 8842−8844. (46) Cho, S.; Ma, B.; Nguyen, T.-S.; Hupp, J.-T.; Albrecht-Schmitt, T.-E. Chem. Commun. 2006, 2563−2565. (47) Cho, S.; Gadzikwa, T.; Afshari, M.; Nguyen, T.-S.; Hupp, J.-T. Eur. J. Inorg. Chem. 2007, 4863−4867. (48) Hong, D.; Hwang, Y.-K.; Serre, C.; Férey, G.; Chang, J. Adv. Funct. Mater. 2009, 19, 1537−1552. (49) Xuan, Z.-H.; Zhang, D.-S.; Chang, Z.; Hu, T.-L.; Bu, X.-H. Inorg. Chem. 2014, 53, 8985−8990. (50) Habib, H.-A.; Hoffmann, A.; Hoppea, H.-A.; Janiak, C. Dalton Trans. 2009, 1742−1751. (51) Zheng, B.; Bai, J.; Duan, J.; Wojtas, L.; Zaworotko, M.-J. J. Am. Chem. Soc. 2011, 133, 748−751. (52) Hu, A.; Ngo, H.-L.; Lin, W. J. Am. Chem. Soc. 2003, 125, 11490−11491. (53) Huang, S.-L.; Jia, A.-Q.; Jin, G.-X. Chem. Commun. 2013, 49, 2403−2405. (54) Wu, C.; Lin, W. Angew. Chem., Int. Ed. 2007, 46, 1075−1078. (55) Li, H.; Han, Y.-F.; Lin, Y.-J.; Guo, Z.-W.; Jin, G.-X. J. Am. Chem. Soc. 2014, 136, 2982−2985. (56) Hu, A.; Ngo, H.-L.; Lin, W. Angew. Chem., Int. Ed. 2003, 42, 6000−6003. (57) Huang, S.-L.; Lin, Y.-J.; Andy Hor, T.-S.; Jin, G.-X. J. Am. Chem. Soc. 2013, 135, 8125−8128. (58) Schoedel, A.; Boyette, W.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M.-J. J. Am. Chem. Soc. 2013, 135, 14016−14019. (59) James, S.-L. Chem. Soc. Rev. 2003, 32, 276−288. (60) Shi, D.-B.; Ren, Y.-W.; Jiang, H.-F.; Cai, B.-W.; Lu, J.-X. Inorg. Chem. 2012, 51, 6498−6506. (61) Ferey, G. Chem. Soc. Rev. 2008, 37, 191−214. (62) Habib, H.-A.; Sanchiz, J.; Janiak, C. Inorg. Chim. Acta 2009, 362, 2452−2460. (63) Elsaidi, S.-K.; Mohamed, M.-H.; Wojtas, L.; Chanthapally, A.; Pham, T.; Space, B.; Vittal, J.-J.; Zaworotko, M.-J. J. Am. Chem. Soc. 2014, 136, 5072−5077. (64) D’Vries, R.-F.; Iglesias, M.; Snejko, N.; Gutiérrez-Puebla, E.; Monge, M.-A. Inorg. Chem. 2012, 51, 11349−11355. (65) Ma, X.-H.; Yuan, W.-B.; Bell, S.-E.-J.; James, S.-L. Chem. Commun. 2014, 50, 1585−1587. (66) Ma, X.-H.; Lim, G.-K.; Harris, K.-D.-M.; Apperley, D.-C.; Horton, P.-N.; Hursthouse, M.-B.; James, S.-L. Cryst. Growth Des. 2012, 12, 5869−5872. (67) Haasnoot, J. G. Coord. Chem. Rev. 2000, 200−202, 131−185. (68) Bencinia, A.; Lippolisb, V. Coord. Chem. Rev. 2010, 254, 2096− 2180. (69) Sammes, P.-G.; Yahioglu, G. Chem. Soc. Rev. 1994, 23, 327−334.

REFERENCES

(1) Dunbar, P.-G.; Durant, G.-J.; Fang, Z.; Abuh, Y.-F.; El-Assadi, A.A.; Ngur, D.-O.; Periyasamy, S.; Hoss, W.-P.; Messer, W.-S. J. Med. Chem. 1993, 36, 842−847. (2) Messer, W.-S.-J.; Abuh, Y.-F.; Ryan, K.; Shepherd, M.-A.; Schroeder, M.; Abunada, S.; Sehgal, R.; El-Assadi, A.-A. Drug. Dev. Res. 1997, 40, 171−184. (3) Dolinkin, A.-O.; Chernov’yants, M.-S. Pharm. Chem. J. 2010, 44, 99−106. (4) Kahlon, D.-K.; Lansdell, T.-A.; Fisk, J.-S.; Tepe, J.-J. Bioorg. Med. Chem. 2009, 17, 3093−3103. (5) Meidute-Abaraviciene, S.; Mosen, H.; Lundquist, I.; Salehi, A. Acta Physiol. 2009, 195, 375−383. (6) Li, H. Y.; Drummond, S.; De Lucca, I.; Boswell, G.-A. Tetrahedron 1996, 52, 11153−11162. (7) Masajtis-Zagajewska, A.; Majer, J.; Nowicki, M. Hypertens. Res. 2010, 33, 348−353. (8) Sun, M.; Wu, X.-Q.; Chen, J.-Q.; Cai, J.; Cao, M.; Ji, M. Eur. J. Med. Chem. 2010, 45, 2299−2306. (9) Zhou, S.-M.; Kern, E.-R.; Gullen, E.; Cheng, Y.-C.; Drach, J.-C.; Matsumi, S.; Mitsuya, H.; Zemlicka, J. J. Med. Chem. 2004, 47, 6964− 6972. (10) Weinhardt, K.; Wallach, M.-B.; March, M. J. Med. Chem. 1985, 28, 694−698. (11) Linger, C.; Azadi, P.; Macleod, J.-K.; Dell, A.; Abdallah, M.-A. Tetrahedron Lett. 1992, 33, 1737−1740. (12) McFarland, J.-W.; Howes, H.-L. J. Med. Chem. 1972, 15, 365− 368. (13) Garcia, M.-B.; Grilli, S.; Lunazzi, L.; Mazzanti, A.; Orelli, L.-R. J. Org. Chem. 2001, 66, 6679−6684. (14) Marxer, A. J. Am. Chem. Soc. 1957, 79, 467−472. (15) Oxley, P.; Short, W.-H. J. Chem. Soc. 1947, 497−505. (16) Mirkhani, V.; Moghadam, M.; Tangestaninejad, S.; Kargar, H. Tetrahedron Lett. 2006, 47, 2129−2132. (17) Hojati, S.-F.; Mohammadpoor-Baltork, I.; Maleki, B.; Gholizadeh, M.; Shafiezadeh, F.; Haghdoust, M. Can. J. Chem. 2010, 88, 135−141. (18) Dash, P.; Kudav, D.-P.; Parihar, J.-A. J. Chem. Res. 2004, 7, 490− 491. (19) Mohammadpoor-Baltork, I.; Moghadam, M.; Tangestaninejad, S.; Mirkhani, V.; Hojati, S.-F. Polyhedron 2008, 27, 750−758. (20) Mohammadpoor-Baltork, I.; Moghadam, M.; Tangestaninejad, S.; Mirkhani, V.; Hojati, S.-F. Catal. Commun. 2008, 9, 1153−1156. (21) Mohammadpoor-Baltorka, I.; Moghadama, M.; Tangestaninejada, S.; Mirkhania, V.; Eskandaria, Z.; Salavatib, H. J. Iran. Chem. Soc. 2011, 8, S17−S27. (22) Ferguson, M.; Giri, N.; Huang, X.; Apperley, D.; James, S.-L. Green Chem. 2014, 16, 1374−1382. (23) James, S.-L.; Frišcǐ ć, T. Chem. Soc. Rev. 2013, 42, 7494−7496. (24) Phan, A.; Doonan, C.-J.; Uribe-Romo, F.-J.; Knobler, C.-B.; O’Keeffe, M.; Yaghi, O.-M. Acc. Chem. Res. 2010, 43, 58−67. (25) Guo, Z.; Cao, R.; Wang, X.; Li, H.; Yuan, W.; Wang, G.; Wu, H.; Li, J. J. Am. Chem. Soc. 2009, 131, 6894−6895. (26) Li, J.; Kuppler, R.-J.; Zhou, H. Chem. Soc. Rev. 2009, 38, 1477− 1504. (27) Zheng, A.-X.; Si, J.; Tang, X.-Y.; Miao, L.-L.; Yu, M.; Hou, K.-P.; Wang, F.; Li, H.-X.; Lang, J.-P. Inorg. Chem. 2012, 51, 10262−10273. (28) Xie, L.; Lin, J.; Liu, X.; Wang, Y.; Zhang, W.; Zhang, J.; Chen, X. Inorg. Chem. 2010, 49, 1158−1165. (29) Li, H.-X.; Zhao, W.; Li, H.-Y.; Xu, Z.-L.; Wang, W.-X.; Lang, J.P. Chem. Commun. 2013, 49, 4259−4261. (30) Dai, F.; Dou, J.; He, H.; Zhao, X.; Sun, D. Inorg. Chem. 2010, 49, 4117−4224. (31) Li, K.-H.; Olson, D.-H.; Seidel, J.; Emge, T.-J.; Gong, H.-W.; Zeng, H.-P.; Li, J. J. Am. Chem. Soc. 2009, 131, 10368−10369. (32) Liu, D.; Ren, Z.; Li, H.; Lang, J.-P.; Li, N.; Abrahams, B.-F. Angew. Chem., Int. Ed. 2010, 49, 4767−4770. (33) Düren, T.; Bae, Y.; Snurr, R.-Q. Chem. Soc. Rev. 2009, 38, 1237− 1247. 2721

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722

Crystal Growth & Design

Article

(70) Sumby, C.-J. Coord. Chem. Rev. 2011, 255, 1937−1967. (71) Gómez-Lor, B.; Gutiérrez-Puebla, E.; Iglesias, M.; Monge, M.A.; Ruiz-Valero, C.; Snejko, N. Chem. Mater. 2005, 17, 2568−2573. (72) Berl, V.; Schmutz, M.; Krische, M.-J.; Khoury, R.-G.; Lehn, J.-M. Chem.Eur. J. 2002, 8, 1227−1244. (73) Meng, X.-R.; Song, Y.-L.; Hou, H.-W.; Han, H.-Y.; Xiao, B.; Fan, Y.-T.; Zhu, Y. Inorg. Chem. 2004, 43, 3528−3536. (74) König, E. Magnetic Properties of Coordination and Organometallic Transition Metal Compounds; Springer: Berlin, 1966. (75) Ishihara, M.; Togo, H. Tetrahedron 2007, 63, 1474−1480. (76) An, S.-J.; Yin, B.; Liu, P.; Li, X.-G.; Li, C.; Li, J.-L.; Shi, Z. Synthesis 2013, 45, 2525−2532. (77) Sheldrick, G.-M. Acta Crystallogr., Sect. A 2008, 64, 112−122. (78) Yang, L.; Powell, D.-R.; Houser, R.-P. Dalton Trans. 2007, 955− 964. (79) Balaban, A.-T. From Chemical Topology to Three-Dimensional Geometry; Plenum Press: New York, 1997. (80) Addison, A.-W.; Rao, T.-N.; Reedijk, J.; Rijn, J.-V.; Verschoor, G.-C. J. Chem. Soc., Dalton Trans. 1984, 1349−1356. (81) Murphy, G.; Nagle, P.; Murphy, B.; Hathaway, B. J. Chem. Soc., Dalton Trans. 1997, 2645−2652. (82) Gu, Z.-G.; Xu, Y.-F.; Zhou, X.-H.; Zuo, J.-L.; You, X.-Z. Cryst. Growth Des. 2008, 8, 1306−1312. (83) Wang, X.-L.; Qin, C.; Wang, E.-B.; Su, Z.-M. Chem.Eur. J. 2006, 12, 2680−2691. (84) Fan, J.; Yee, G.-T.; Wang, G.-B.; Hanson, B.-E. Inorg. Chem. 2006, 45, 599−608. (85) Lan, Y.-Q.; Li, S.-L.; Su, Z.-M.; Shao, K.-Z.; Ma, J.-F.; Wang, X.L.; Wang, E.-B. Chem. Commun. 2008, 58−60. (86) Sun, L.-X.; Qi, Y.; Che, Y.-X.; Batten, S.-R.; Zheng, J.-M. Cryst. Growth Des. 2009, 9, 2995−2998. (87) Long, L.-S.; Cai, J.-W.; Ren, Y.-P.; Tong, Y.-X.; Chen, X.-M.; Ji, L.-N.; Huang, R.-B.; Zheng, L.-S. J. Chem. Soc., Dalton Trans. 2001, 845−849. (88) Wang, M.-C.; Sue, L.-S.; Liau, B.-C.; Ko, B.-T.; Elango, S.; Chen, J.-H. Inorg. Chem. 2001, 40, 6064−6068. (89) Liu, L.; Liu, Y.-H.; Han, G.; Wu, D.-Q.; Hou, H.-W.; Fan, Y.-T. Inorg. Chim. Acta 2013, 403, 25−34. (90) Rueff, J. M.; Masciocchi, N.; Rabu, P.; Sironi, A. Eur. J. Inorg. Chem. 2001, 2843−2848. (91) Wang, X.-Y.; Sevov, S.-C. Inorg. Chem. 2008, 47, 1037−1043. (92) Liang, L.-L; Ren, S.-B.; Wang, J.; Zhang, J.; Li, Y.-Z.; Du, H.-B.; You, X.-Z. CrystEngComm 2010, 12, 2669−2671. (93) Carlin, R.-L. Magnetochemistry; Springer: Berlin, 1986. (94) Bleaney, B.; Bowers, K.-D. Proc. R. Soc. London 1952, 214, 451− 465.

2722

DOI: 10.1021/acs.cgd.5b00016 Cryst. Growth Des. 2015, 15, 2712−2722