Skip to content
Publicly Available Published by De Gruyter August 26, 2015

Influence of metal ions on the formation of new metal complexes constructed from tetrachlorophthalic acid

  • An-Qi Dai , Qi Yan , Jing Zhong , Sheng-Chun Chen EMAIL logo , Ming-Yang He and Qun Chen EMAIL logo

Abstract

Reaction of 3,4,5,6-tetrachloro-1,2-benzenedicarboxylyic acid (1,2-H2BDC-Cl4) with transitional metal salts at room temperature in mixed DMF/H2O solvent affords three complexes formulated as [Cu(1,2-HBDC-Cl4)2(DMF)2] (1), {[Cd(1,2-HBDC-Cl4)2(H2O)4]·2DMF} (2), and {[Ni(1,2-BDC-Cl4)(H2O)5]·DMF·H2O} (3) (DMF = N,N-dimethylformamide). All these complexes have been characterized by elemental analysis, IR spectroscopy, thermogravimetric analysis, and single-crystal X-ray crystallography. In 1, the CuII ion is four-coordinated with a square-planar geometry formed by two 1,2-HBDC-Cl4 anions and two DMF ligands; in 2, the CdII ion takes an octahedral geometry coordinated by two 1,2-HBDC-Cl4 anions and four aqua ligands; while in 3, the NiII ion is octahedrally coordinated by one 1,2-BDC-Cl4 dianion and five aqua ligands. Intermolecular O–H···O hydrogen bonds and Cl···Cl (or C–H···Cl) interactions provide a significant contribution to stabilizing the three mononuclear structures in the solid state. The results suggest that structural differences among them are attributed to the influence of transition metal ions. The fluorescence of the complexes and of 1,2-H2BDC-Cl4 has been investigated. No significant metal effect has been observed.

1 Introduction

The chemistry of metallosupramolecular species assembled from transition metal ions and organic ligands has attracted great interest not only because of their intriguing structural diversities but also because of their potential applications as new materials [1–8]. In this field, the most effective and facile approach is the appropriate choice of well-designed organic ligands containing modifiable backbones and modes of connectivity, together with metal centers with various coordination preferences [9–12]. Among various carboxylate ligands, the phthalate anions with two carboxylic groups in ortho-position can bind metal ions in many diverse coordination modes [13], leading to the formation of an extensive range of supramolecular coordination systems [14–19]. Phthalic acids with substituents such as methyl [20], nitro [21], and fluoro [22] groups offer great potential for constructing coordination architectures, since the substituents can significantly affect their solid structures, physical and chemical properties by virtue of polarizability, and steric requirement. In this regard, tetrachlorophthalic acid (1,2-H2BDC-Cl4) containing a perchlorinated benzene skeleton is expected to assist in assembling fascinating coordination and supramolecular structures with unique properties. Up to now, main group and lanthanide metal complexes with the anions of 1,2-H2BDC-Cl4 have been reported [23–25]; however, its transition metal complexes are still comparatively less investigated [26].

As a part of our efforts to investigate the design and control of the self-assembly of inorganic–organic hybrid materials based on perchlorinated benzenedicarboxylate ligands [27–29], very recently, we have reported the unusual solvent effect of formamide on the structural variations of chiral and achiral MnII complexes with the anions of 1,2-BDC-Cl4 [30]. A double-strand helical chain structure possessing two single-strand helices of the same handedness was observed when the DMF/H2O solvent system was used. The interesting question arises from this result, whether chiral coordination compounds can be spontaneously resolved with other transition metal ions incorporating the 1,2-BDC-Cl4 anion, and if not so, what would they be like? To further investigate the effect of chlorine interactions on the supramolecular structures of functional crystalline solids, herein, three mononuclear complexes, [Cu(1,2-HBDC-Cl4)2(DMF)2] (1), {[Cd(1,2-HBDC-Cl4)2(H2O)4]·2DMF} (2), and {[Ni(1,2-BDC-Cl4)(H2O)5]·DMF·H2O} (3) (see Scheme 1), were prepared by the reactions of 1,2-H2BDC-Cl4 with CuII, CdII, and NiII salts in DMF/H2O solvent mixtures at room temperature. X-ray structural analyses have shown that all three complexes crystallized in centrosymmetrical space groups (Pbca for 1, Pbcn for 2, and P1̅ for 3), while the aforementioned MnII complex crystallized in the non-centrosymmetrical space group P212121 [30]. The investigation of the packing of the molecules of the three complexes has demonstrated that the co-effect of intermolecular O–H···O hydrogen bonding and weak Cl···Cl (or C–H···Cl) interactions is among the driving forces in the self-assembly of these discrete molecules to generate higher dimensional supramolecular networks. The spectroscopic, thermal, and luminescence properties of 13 are described and discussed.

Scheme 1: Synthetic routes for complexes 1–3.
Scheme 1:

Synthetic routes for complexes 1–3.

2 Results and discussion

2.1 Synthesis and general characterization

Complexes 1–3 were prepared by using 1,2-H2BDC-Cl4 and CuII, CdII, and NiII salts in DMF/H2O under similar conditions. All three complexes are stable to air and moisture and are insoluble in common solvents such as water, alcohol, and acetonitrile. The IR spectra of 1–3 show features attributable to the components of the complexes. The broad bands centered at ca. 3407–3444 cm–1 indicate the O–H stretching of hydroxyl or water. The strong absorption bands at 1718 cm–1 for 1 and 1720 cm–1 for 2 reveal the presence of carboxyl groups, while there is no absorption band around 1720 cm–1 for 3, indicating the complete deprotonation of the 1,2-H2BDC-Cl4 ligand, which is consistent with the crystal structure determination. All three IR spectra display strong peaks in the ranges of 1601–1645 and 1386–1433 cm–1, which can be attributed to the antisymmetric and symmetric stretching vibrations of carboxylate groups. The elemental analysis results of these complexes match well with their molecular formulae evaluated by X-ray diffraction analyses.

2.2 Structural description of complexes 1–3

2.2.1 [Cu(1,2-HBDC-Cl4)2(DMF)2] (1)

Single-crystal X-ray diffraction has revealed that crystals of 1 are orthorhombic with space group Pbca and the asymmetric unit contains half a CuII ion, one 1,2-HBDC-Cl4 anion, and one DMF ligand. As shown in Fig. 1a, the CuII ion is four-coordinated by two oxygen atoms from two 1,2-HBDC-Cl4 ligands with a Cu–O distance of 1.930(1) Å and two oxygen atoms from two DMF ligands with a Cu–O distance of 1.949(1) Å. In this complex, the acid 1,2-H2BDC-Cl4 is mono-deprotonated to the 1,2-HBDC-Cl4 anion and adopts a monodentate terminal coordination mode, giving rise to a discrete structure. Each [Cu(1,2-HBDC-Cl4)2(DMF)2] unit is linked to four neighboring ones by strong intermolecular O3–H3···O2i hydrogen bonding interactions (H···O/O···O distance: 1.807/2.626(2) Å, angle: 178.8(1)°, i = –x– 1/2,y– 1/2, z) between the uncoordinated carboxylate O2 atoms and H atoms of the carboxyl group, affording a supramolecular layer (Fig. 1b). These layers are further assembled into a 3D supramolecular framework via intermolecular Cl1···Cl4ii interactions (Cl···Cl distance = 3.654 Å, ii = –x, y+ 1/2, –z+ 1/2) (Fig. 1c). It should be pointed out that the Cl···Cl distance is shorter than twice Pauling’s van der Waals radius of the chlorine atom (3.76 Å) [31], and is comparable to that stated by Bondi (3.52 Å) [32].

Fig. 1: Views of (a) the coordination environment of the CuII center in 1 (symmetry code: #1 –x, –y, –z), (b) the 2D supramolecular layer with O–H···O hydrogen bonds highlighted as dashed lines, and (c) the 3D supramolecular architecture (Cl···Cl and O–H···O hydrogen bonds are shown as dashed lines).
Fig. 1:

Views of (a) the coordination environment of the CuII center in 1 (symmetry code: #1 –x, –y, –z), (b) the 2D supramolecular layer with O–H···O hydrogen bonds highlighted as dashed lines, and (c) the 3D supramolecular architecture (Cl···Cl and O–H···O hydrogen bonds are shown as dashed lines).

2.2.2 {[Cd(1,2-HBDC-Cl4)2(H2O)4]·2DMF} (2)

X-ray analysis reveals that 2 crystallizes in the space group Pbcn and the asymmetric unit contains half a CdII ion, one 1,2-HBDC-Cl4 anion, two aqua ligands, and one DMF guest molecule. The CdII ion is six-coordinated by two oxygen atoms from two 1,2-HBDC-Cl4 anions and four oxygen atoms from four aqua ligands to constitute an octahedral geometry (Fig. 2a). The Cd–O bond lengths range from 2.282(2) to 2.305(2) Å, and the O–Cd–O bond angles are in the range of 82.4(1)–176.8(1)°. Similar to 1, the mono-deprotonated 1,2-HBDC-Cl4 units of 2 are coordinated to CdII ion via a monodentate terminal coordination mode. The analysis of the molecular packing of 2 in the unit cell shows that there exist multiple hydrogen bond interactions, resulting in a 2D supramolecular network. As illustrated in Fig. 2b, all carboxylate oxygen atoms are involved in the formation of O–H···O hydrogen bonding with water ligands or carboxyl groups (hydrogen bond parameters are listed in Table 1) to result in a chain along the crystallographic c axis. These chains are further stabilized by intermolecular O5–H5A···O5i hydrogen bonding interactions (H···O/O···O distance: 2.53/3.107(3) Å, angle: 128°, i = –x+1,y,z) between coordinated water ligands. With respect to the non-binding DMF guest, it works as a linker to fabricate a supramolecular layer in the crystallographic bc plane connecting the above-mentioned 1D motifs through intermolecular O5–H5B···O7 and C10–H10C···Cl1 interactions (see Table 1 for details), as is depicted in Fig. 2c.

Fig. 2: Views of (a) the coordination environment of the CdII center in 2 (symmetry code: #1 –x + 1, y + 1, –z + 1/2), (b) the hydrogen bonding chain arrangement of 2, and (c) the 2D supramolecular network constructed via O–H···O and C–H···Cl interactions.
Fig. 2:

Views of (a) the coordination environment of the CdII center in 2 (symmetry code: #1 –x + 1, y + 1, –z + 1/2), (b) the hydrogen bonding chain arrangement of 2, and (c) the 2D supramolecular network constructed via O–H···O and C–H···Cl interactions.

Table 1

Hydrogen bond parameters in the crystal structures of 1–3.

ComplexD–H···AH···A (Å)D···A (Å)D–H···A (deg)Symmetry code
1O3–H3···O21.812.626(2)179x– 1/2, y– 1/2,z
2O4–H4···O11.782.581(3)166x,y, z– 1/2
O5–H5A···O22.533.029(3)120–x+1,y,z
O5–H5A···O32.583.155(3)129–x+1,y,z
O5–H5A···O52.533.107(3)128–x+1,y,z
O5–H5B···O71.842.644(3)165x,y– 1, z
O6–H6A···O32.253.032(4)159–x+1,y,z
C10–H10C···Cl12.813.575(5)138
O6–H6B···O22.232.983(4)153
3O6–H6A···O112.012.849(7)172
O6–H6B···O21.832.672(4)170x–1, y,z
O7–H7A···O51.852.703(5)176
O7–H7B···O22.002.835(5)163x+ 1,y,z+ 1
O8–H8A···O51.902.734(5)164x,y,z+ 1
O8–H8B···O41.782.618(5)167x+ 1,y+ 1,z+ 1
O9–H9A···O31.932.780(5)164x+ 1,y+ 1,z+ 1
O9–H9B···O21.992.752(4)145
O10–H10A···O31.992.836(5)167
O10–H10B···O81.982.826(5)176x,y+ 1,z+ 1
O11–H11D···O42.032.849(7)160x–1, y,z
O11–H11E···Cl32.303.121(6)162x+ 1,y+ 1,z

2.2.3 {[Ni(1,2-BDC-Cl4)(H2O)5]·DMF·H2O} (3)

Complex 3 forms monoclinic crystals with space group P1̅. The asymmetric unit consists of one NiII ion, one dianionic 1,2-BDC-Cl4 ligand, five coordinated water molecules, one free water molecule, and one free DMF molecule. In contrast to the mono-deprotonated 1,2-HBDC-Cl4 ligands of 1 and 2, 1,2-H2BDC-Cl4 in 3 is fully deprotonated. One carboxylate group is coordinated to one NiII center with a monodentate coordination mode, while the other is uncoordinated. As shown in Fig. 3a, the NiII ion is six-coordinated by one oxygen atom from one 1,2-BDC-Cl4 dianion and five oxygen atoms from five aqua ligands, taking on a slightly distorted octahedral geometry. The Ni–O bond lengths vary from 2.009(3) to 2.070(3) Å, and the O–Ni–O bond angles range from 84.6(1) to 179.2(1)°. Each mononuclear [Ni(1,2-BDC-Cl4)(H2O)5] unit is connected by a pair of strong intermolecular O10–H10B···O8 hydrogen bonding interactions (see Table 1 for details) between coordinated O10 and O8 water molecules to constitute a dimeric [Ni(1,2-BDC-Cl4)(H2O)5]2 subunit with a Ni···Ni separation of 5.161(1) Å. The interstitial DMF molecules connect the neighboring dimeric units by O7–H7A···O5 and O8–H8A···O5 interactions (see Table 1) to afford a chain-like supramolecular array along the crystallographic b axis (see Fig. 3b). These adjacent 1D motifs are further connected through intermolecular O6–H6B···O2 hydrogen bond interactions (see Table 1) between the coordinated water molecules of O6 and uncoordinated carboxylate O2 atoms of 1,2-BDC-Cl4, leading to the formation of a 2D supramolecular sheet extending along the bc plane (see Fig. 3c).

Fig. 3: Views of (a) the coordination environment of the NiII center in 3, (b) the 1D hydrogen bonding arrangement of 3 containing [Ni(1,2-BDC-Cl4)(H2O)5]2 dimeric subunits, and (c) the 2D supramolecular layered structure.
Fig. 3:

Views of (a) the coordination environment of the NiII center in 3, (b) the 1D hydrogen bonding arrangement of 3 containing [Ni(1,2-BDC-Cl4)(H2O)5]2 dimeric subunits, and (c) the 2D supramolecular layered structure.

In comparison with our previously reported 1D chiral MnII complex [30], complexes 1–3 all crystallize in different achiral space groups and display mononuclear structures. The CuII ion in 1 takes on a square planar coordination geometry completed by two anionic 1,2-HBDC-Cl4 ligands and two DMF ligands, the CdII ion in 2 adopts a distorted octahedral geometry comprising two anionic 1,2-HBDC-Cl4 ligands and four water ligands, while the NiII ion in 3 is octahedrally coordinated by one dianionic 1,2-BDC-Cl4 ligand and five water ligands. The carboxyl groups of 1,2-H2BDC-Cl4 are mono-deprotonated in 1 and 2, which are different from the completely deprotonated ligand in 3. The coordination preference of the metal ions clearly plays an important role in the formation of their complexes since the only difference in the synthetic conditions employed for these complexes is the transition metal components.

2.3 Thermal stability

To investigate the thermal stability of complexes 1–3, thermogravimetric (TGA) measurements were carried out from room temperature to 800 °C and the corresponding curves are shown in Fig. 4. No decomposition of 1 is observed below 95 °C, and the first weight loss of 17.2% in the range of 95 °C–170 °C corresponds to the loss of coordinated DMF ligands (calculated: 17.9%). The residual solid starts to decompose near 230 °C, which is complete at 800 °C. The TGA curve of 2 shows the initial weight loss of 22.9% in the range of 75 °C–190 °C, which can be ascribed to the removal of four water ligands and two lattice DMF molecules (calculated: 23.3%). The weight loss continues until the temperature reaches 600 °C. The remaining weight indicates that the final residue possibly is CdO (calculated: 13.7%, found: 14.5%). In the case of 3, the weight loss of 33.1% (calculated: 32.3%) in the range of 50 °C–210 °C is in accordance with the removal of ten water ligands and two lattice DMF molecules as well as one interstitial water molecule. Subsequently, pyrolysis of the remaining substance is observed upon heating to 275 °C, and further heating to 800 °C reveals a further gradual weight loss of the sample. The final solid holds a weight of 14.6% (calculated: 13.8%) of the total sample which corresponds to NiO.

Fig. 4: Solid-state emission spectra of the ligand 1,2-H2BDC-Cl4 and complexes 1–3.
Fig. 4:

Solid-state emission spectra of the ligand 1,2-H2BDC-Cl4 and complexes 1–3.

2.4 Photoluminescence properties

The emission spectra of complexes 1–3 and of the acid 1,2-H2BDC-Cl4, under excitation of 336 nm in the solid state at room temperature, are shown in Fig. 4. The strongest emission peak for the free acid appears at 474 nm, corresponding to n → π* transitions. For 1–3, the maximum emission peaks are similarly observed in the blue region (1, 476 nm; 2, 475 nm; and 3, 476 nm), which are also attributed to ligand-centered transitions. Weaker shoulders around 510 nm can also be assigned to intraligand transitions because a similar peak at about 508 nm appears for 1,2-H2BDC-Cl4. There is a considerable enhancement of the emission intensity of complex 2 compared with free 1,2-H2BDC-Cl4, which may result from the increased rigidity of the organic ligand when binding to CdII, effectively reducing the loss of the energy through radiationless pathways [33, 34]. This result reveals that no significant effect of metal ions on the fluorescence of the three complexes has been observed.

In summary, three new discrete CuII, CdII, and NiII complexes with perchlorinated phthalate ligands have been synthesized and structurally characterized. The results reveal that the coordination geometries of the transition metal ions play an important role in governing the structural features of these metal-organic coordination architectures. The investigation of the packing of the molecules of such complexes suggests that intermolecular O–H···O hydrogen bonding and Cl···Cl (or C–H···Cl) interactions are among the driving forces in linking the discrete entities into high-dimensional supramolecular networks. The complexes exhibit blue emissions in the solid state at room temperature, not markedly different from that of tetrachlorophthalic acid.

3 Experimental section

All chemicals were commercially available and used as received. Infrared spectra were recorded with a Nicolet ESP 460 Fourier transform spectrometer (Niolet, USA) on KBr pellets in the range of 4000–400 cm–1. Elemental analyses were performed with a Perkin-Elmer PE2400II elemental analyzer (TA, USA). TGA were carried out in the temperature range of 25 °C–800 °C on a Dupont thermal analyzer under N2 atmosphere at a heating rate of 10 °C min–1. Luminescence spectra in the solid state were measured at room temperature on a Varian Cary Eclipse spectrometer (Varian, USA).

3.1 Synthesis of [Cu(1,2-HBDC-Cl4)2(DMF)2] (1)

A solution of CuCl2·2H2O (51.1 mg, 0.3 mmol) in H2O (8 mL) was added to a solution of 1,2-H2BDC-Cl4 (91.2 mg, 0.3 mmol) in DMF (8 mL). After stirring for ca. 15 min, the resultant solution was filtrated and left to stand at room temperature. Blue block-shaped single crystals of 1 suitable for X-ray diffraction were obtained after 1 week by slow evaporation of the solvents in ca. 70% yield (85.6 mg; based on 1,2-H2BDC-Cl4). – Analysis for C22H16Cl8CuN2O10 (%): calcd. C 32.4, N 3.4, H 2.0; found C 33.2, N 3.2, H 1.9. – IR (cm–1): 3423 br, 2959 m, 2927 m, 1718 s, 1643 s, 1577 s, 1433 s, 1395 s, 1344 s, 1239 m, 1123 m, 1056 w, 905 s, 831 m, 667 s, 603 m.

3.2 Synthesis of {[Cd(1,2-HBDC-Cl4)2(H2O)4]· 2DMF} (2)

The same synthetic procedure as that for 1 was used except that CuCl2·2H2O was replaced by CdCl2 (55.0 mg, 0.3 mmol), giving colorless sheet-like single crystals of 2 in ca. 40% yield (56.2 mg, based on 1,2-H2BDC-Cl4). – Analysis for C22H24CdCl8N2O14 (%): calcd C 28.2, N 3.0, H 2.6; found C 27.6, N 3.3, H 2.5. – IR (cm–1): 3444 br, 2958 m, 2931 m, 1720 s, 1610 s, 1573 s, 1426 s, 1386 s, 1261 s, 1120 m, 1082 w, 935 w, 904 m, 828 m, 643 m, 538 m.

3.3 Synthesis of {[Ni(1,2-BDC-Cl4)(H2O)5]2· 2DMF·H2O} (3)

The same synthetic procedure as that for 1 was used except that CuCl2·2H2O was replaced by NiCl2·6H2O (71.3 mg, 0.3 mmol), affording green block-shaped single crystals of 3 in ca. 50% yield (79.9 mg, based on 1,2-H2BDC-Cl4). – Analysis for C22H36Cl8N2Ni2O21 (%): calcd. C 24.4, N 2.6, H 3.5; found C 24.2, N 2.7, H 3.4. – IR (cm–1): 3407 br, 2967 m, 2925 m, 1645 s, 1601 s, 1422 s, 1392 s, 1341 s, 1264 s, 1123 m, 1104 m, 938 w, 812 w, 738 m, 667 m, 621 w.

3.4 X-ray structure determinations

Single-crystal X-ray diffraction measurements were performed on a Bruker Apex II CCD diffractometer at ambient temperature with MoKα radiation (λ = 0.71073 Å). In each case, a semiempirical absorption correction was applied using sadabs [35], and the program saint was used for integration of the diffraction profiles [36]. The structures were solved by direct methods using the shelxs program of the shelxtl package and refined anisotropically for all non-hydrogen atoms by full-matrix least squares on F2 with shelxl [37–40]. Carbon-bound hydrogen atoms were placed in geometrically calculated positions using a riding model. Oxygen-bound hydrogen atoms were firstly localized by difference Fourier maps and then fixed geometrically with isotropic displacement parameters. Crystallographic data and structural refinement parameters are summarized in Table 2, and selected bond lengths and angles are listed in Table 3.

Table 2

Crystallographic data and data collection and refinement details for complexes 13.

Complex123
Empirical formulaC22H16Cl8CuN2O10C22H24Cl8CdN2O14C11H19Cl4NNiO11
Formula weight815.51936.43541.78
Crystal size, mm30.30 × 0.28 × 0.280.26 × 0.22 × 0.200.15 × 0.14 × 0.13
Crystal systemOrthorhombicOrthorhombicTriclinic
Space groupPbcaPbcnP
a, Å10.112(2)26.456(3)7.309(3)
b, Å12.194(2)10.794(2)9.116(4)
c, Å24.098(2)11.956(2)15.481(7)
α, deg909078.620(8)
β, deg909088.577(9)
γ , deg909082.380(9)
V, deg2971.8(5)3414.2(1)1002.3(8)
Z442
Dcalcd., mg m–31.821.821.79
μ, mm–11.51.31.6
F(000), e16281864552
Refl. total/unique152 46/261017 591/30035533/3459
Rint0.05310.05450.0184
R1a/wR2b [I > 2 σ(I)]0.0264/0.06460.0321/0.08480.0445/0.1543
R1a/wR2b (all data)0.0360/0.06680.0365/0.08910.0577/0.2139
GOF (F2)c1.0181.0361.061
Δρfin (max/min), e Å–30.42/–0.660.49/–0.970.71/–0.91

aR1 = Σ||Fo| – |Fc||/Σ|Fo|; bwR2 = [Σw(Fo2Fc2)2w(Fo2)2]1/2, w = [σ2(Fo2) + (AP)2 + BP]–1, where P = (Max(Fo2, 0) + 2Fc2)/3; cGOF = [Σw(Fo2Fc2)2/(nobsnparam)]1/2.

Table 3

Selected bond lengths (Å) and angles (deg) for 13 with estimated standard deviations in parenthesesa.

Complex 1
Cu1–O11.929(2)Cu1–O51.949(2)
O1–Cu1–O1#1180.0O1–Cu1–O588.3(1)O1–Cu1–O5#191.7(1)
O5–Cu1–O5#1180.0
Complex 2
Cd1–O12.305(2)Cd1–O52.282(2)Cd1–O62.295(2)
O1–Cd1–O1#1176.8(1)O1–Cd1–O588.3(1)O1–Cd1–O5#189.3(1)
O1–Cd1–O699.7(1)O1–Cd1–O6#182.3(1)O5–Cd1–O5#187.7(1)
O5–Cd1–O687.3(1)O5–Cd1–O6#1169.5(1)O6–Cd1–O6#199.1(1)
Complex 3
Ni1–O12.051(3)Ni1–O62.023(3)Ni1–O72.009(3)
Ni1–O82.070(3)Ni1–O92.039(3)Ni1–O92.058(3)
O1–Ni1–O684.7(1)O1–Ni1–O793.4(1)O1–Ni1–O8174.0(1)
O1–Ni1–O994.7(1)O1–Ni1–O1087.7(1)O6–Ni1–O788.3(1)
O6–Ni1–O890.1(1)O6–Ni1–O9179.2(1)O6–Ni1–O1090.3(1)
O7–Ni1–O889.3(1)O7–Ni1–O991.3(1)O7–Ni1–O10178.1(1)
O8–Ni1–O990.6(1)O8–Ni1–O1089.5(1)O9–Ni1–O1090.0(1)

aSymmetry codes: for 1, #1: –x, –y, –z; for 2, #1: –x + 1, y, –z + 1/2.

CCDC 1025077, 1025078, and 1025079 contain the supplementary crystallographic data for complexes 13, respectively. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre viawww.ccdc.cam.ac.uk/data_request/cif.


*Corresponding authors: Sheng-Chun Chen and Qun Chen, Jiangsu Key Laboratory of Advanced Catalytic Materials and Technology, Changzhou University, Changzhou 213164, P. R. China, Fax: +86-51986330251, E-mail: (S.-C. Chen); (Q. Chen)

Acknowledgments

We gratefully acknowledge financial support by the National Natural Science Foundation of China (21201026), the Nature Science Foundation of Jiangsu Province (BK20131142), and a project funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions.

References

[1] M. P. Suh, H. J. Park, T. K. Prasad, D. W. Lim, Chem. Rev. 2012, 112, 782.10.1021/cr200274sSearch in Google Scholar

[2] Y. Cui, Y. Yue, G. Qian, B. Chen, Chem. Rev. 2012, 112, 1126.10.1021/cr200101dSearch in Google Scholar

[3] W. Zhang, R.-G. Xiong, Chem. Rev. 2012, 112, 1163.10.1021/cr200174wSearch in Google Scholar

[4] M. Yoon, R. Srirambalaji, K. Kim, Chem. Rev. 2012, 112, 1196.10.1021/cr2003147Search in Google Scholar

[5] J. J. Perry IV, J. A. Perman, M. J. Zaworotko, Chem. Soc. Rev. 2009, 38, 1400.10.1039/b807086pSearch in Google Scholar

[6] J. R. Long, O. M. Yaghi, Chem. Soc. Rev. 2009, 38, 1213.10.1039/b903811fSearch in Google Scholar

[7] S. Kitagawa, R. Matsuda, Coord. Chem. Rev. 2007, 251, 2490.10.1016/j.ccr.2007.07.009Search in Google Scholar

[8] B. Moulton, M. J. Zaworotko, Chem. Rev. 2001, 101, 1629.10.1021/cr9900432Search in Google Scholar

[9] A. Y. Robin, K. M. Fromm, Coord. Chem. Rev. 2006, 250, 2127.10.1016/j.ccr.2006.02.013Search in Google Scholar

[10] L. Pan, X.-Y. Huang, J. Li, Y.-G. Wu, N.-W. Zheng, Angew. Chem. Int. Ed. 2000, 39, 527.Search in Google Scholar

[11] B. F. Abrahams, S. R. Batten, M. J. Grannas, H. Hamit, B. F. Hoskins, R. Robson, Angew. Chem. Int. Ed. 1999, 38, 1475.10.1002/(SICI)1521-3773(19990517)38:10<1475::AID-ANIE1475>3.0.CO;2-3Search in Google Scholar

[12] D. M. L. Goodgame, S. Menzer, A. T. Ross, D. J. Williams, J. Chem. Soc. Chem. Commun. 1994, 22, 2605.10.1039/c39940002605Search in Google Scholar

[13] S. G. Baca, I. G. Filippova, O. A. Gherco, M. Gdaniec, Y. A. Simonov, N. V. Gerbeleu, P. Franz, R. Basler, S. Decurtins, Inorg. Chim. Acta2004, 357, 3419.10.1016/S0020-1693(03)00498-5Search in Google Scholar

[14] D. Sun, Z.-H. Wei, D.-F. Wang, N. Zhang, R.-B. Huang, L.-S. Zheng, Cryst. Growth Des. 2011, 11, 1427.10.1021/cg101562pSearch in Google Scholar

[15] J. Jin, S. Niu, Q. Han, Y. Chi, New J. Chem. 2010, 34, 1176.10.1039/b9nj00565jSearch in Google Scholar

[16] X. Li, M.-Q. Zha, X.-W. Wang, R. Cao, Inorg. Chim. Acta2009, 362, 3357.10.1016/j.ica.2009.03.017Search in Google Scholar

[17] X.-N. Cheng, W. Xue, W.-X. Zhang, X.-M. Chen, Chem. Mater. 2008, 20, 5345.10.1021/cm8012599Search in Google Scholar

[18] A. Thirumurugan, S. Natarajan, Eur. J. Inorg. Chem. 2004, 4, 762.10.1002/ejic.200300594Search in Google Scholar

[19] H.-L. Liu, H.-Y. Mao, H.-Y. Zhang, C. Xu, Q.-A. Wu, G. Li, Y. Zhu, H.-W. Hou, Polyhedron2004, 23, 943.10.1016/j.poly.2003.11.059Search in Google Scholar

[20] G. A. Farnum, A. L. Pochodylo, R. L. LaDuca, Cryst. GrowthDes. 2011, 11, 678.10.1021/cg100972sSearch in Google Scholar

[21] D. Sun, F.-J. Liu, R.-B. Huang, L.-S. Zhang, CrystEngComm2013, 15, 1185.10.1039/C2CE26659HSearch in Google Scholar

[22] Y.-E. Cha, X. Li, S. Song, J. Solid State Chem. 2012, 196, 40.10.1016/j.jssc.2012.07.052Search in Google Scholar

[23] M. K. Rauf, M. A. Saeed, Imtiaz-ud-Din, M. Bolte, A. Badshah, B. Mirza, J. Organomet. Chem. 2008, 693, 3043.Search in Google Scholar

[24] M. Liang, Y.-Q. Sun, D.-Z. Liao, Z.-H. Jiang, S.-P. Yan, P. Cheng, J. Coord. Chem. 2004, 57, 275.10.1080/00958970410001662417Search in Google Scholar

[25] N. Xu, D.-Z. Liao, S.-P. Yan, Z.-H. Jiang, J. Coord. Chem. 2008, 61, 435.10.1080/00958970701357879Search in Google Scholar

[26] S.-Y. Yang, L.-S. Long, Z.-Y. Wu, M.-X. Zhan, R.-B. Huang, L.-S. Zheng, Transition Met. Chem. 2002, 27, 546.10.1023/A:1015644126830Search in Google Scholar

[27] S.-C. Chen, Z.-H. Zhang, K.-L. Huang, Q. Chen, M.-Y. He, A.-J. Cui, C. Li, Q. Liu, M. Du, Cryst. Growth Des. 2008, 8, 3437.10.1021/cg8003905Search in Google Scholar

[28] S.-C. Chen, Z.-H. Zhang, Y.-S. Zhou, W.-Y. Zhou, Y.-Z. Li, M.-Y. He, Q. Chen, M. Du, Cryst. Growth Des. 2011, 11, 4190.10.1021/cg200785gSearch in Google Scholar

[29] S.-C. Chen, Z.-H. Zhang, K.-L. Huang, H.-K Luo, M.-Y. He, M. Du, Q. Chen, CrystEngComm2013, 15, 9613.10.1039/c3ce41108gSearch in Google Scholar

[30] F. Tian, M.-Y. He, K.-L. Huang, Q. Chen, S.-C. Chen, Inorg. Chim. Acta2014, 423, 168.10.1016/j.ica.2014.07.075Search in Google Scholar

[31] L. Pauling, The Nature of the Chemical Bond, Cornell University Press, New York, 1960.Search in Google Scholar

[32] A. Bondi, J. Phys. Chem. 1964, 68, 441.10.1021/j100785a001Search in Google Scholar

[33] H. Yersin, A. Vogler, Photochemistry and Photophysics of Coordination Compounds, Springer-Verlag, Berlin, 1987.10.1007/978-3-642-72666-8Search in Google Scholar

[34] S.-L. Zheng, J.-H. Yang, X.-L. Yu, X.-M. Chen, W.-T. Wong, Inorg. Chem. 2004, 43, 830.Search in Google Scholar

[35] G. M. Sheldrick, SADABS, Program for Empirical Absorption Correction of Area Detector Data, University of Göttingen, Göttingen (Germany) 2002.Search in Google Scholar

[36] SAINT, Software Reference Manual, Bruker Analytical X-ray Instruments Inc., Madison, WI (USA) 1998.Search in Google Scholar

[37] G. M. Sheldrick, SHELXTL NT (version 5.1), Bruker Analytical X-ray Instruments Inc., Madison, WI (USA) 2001.Search in Google Scholar

[38] G. M. Sheldrick, SHELXS/L-97, Programs for Crystal Structure Determination, University of Göttingen, Göttingen (Germany) 1997.Search in Google Scholar

[39] G. M. Sheldrick, Acta Crystallogr.1990, A46, 467.10.1107/S0108767390000277Search in Google Scholar

[40] G. M. Sheldrick, Acta Crystallogr.2008, A64, 112.10.1107/S0108767307043930Search in Google Scholar

Received: 2015-2-7
Accepted: 2015-4-2
Published Online: 2015-8-26
Published in Print: 2015-10-1

©2015 by De Gruyter

Downloaded on 22.5.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2015-0027/html
Scroll to top button