research communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Crystal structure, solvothermal synthesis, thermogravimetric studies and DFT calculations of a five-coordinate cobalt(II) compound based on the N,N-bis­­(2-hy­dr­oxy­eth­yl)glycine anion

CROSSMARK_Color_square_no_text.svg

aSchool of Chemistry and Chemical Engineering, Guangxi University, Nanning, 530004, People's Republic of China, and bGuangxi Experiment Centre of Science and Technology, Guangxi University, Nanning, 530004, People's Republic of China
*Correspondence e-mail: zyl8289@126.com

Edited by A. J. Lough, University of Toronto, Canada (Received 4 August 2016; accepted 14 September 2016; online 23 September 2016)

The reaction of CoCl2·6H2O, N,N-bis­(2-hy­droxy­eth­yl)glycine and tri­ethyl­amine (Et3N) in ethanol solution under solvothermal conditions produced crystals of [N,N-bis­(2-hy­droxy­eth­yl)glycinato]chloridocobalt(II), [Co(C6H12NO4)Cl]. The CoII ion is coordinated in a slightly distorted trigonal–bipyramidal environment which is defined by three O atoms occupying the equatorial plane and the N and Cl atoms in the apical sites. In the crystal, two types of O—H⋯O hydrogen bonds connect the mol­ecules, forming a two-dimensional network parallel to (001). The mol­ecular structure of the title compound confirms the findings of FTIR, elemental analysis, ESI–MS analysis and TG analysis. By using the density functional theory (DFT) (B3LYP) method with 6-31G(d) basis set, the molecular structure has been calculated and optimized.

1. Chemical context

In recent years, coordination compounds have attracted a great deal of inter­est for their structural aesthetics and potential functional applications (Fujita et al., 2004[Fujita, M., Powell, A. & Creutz, C. E. (2004). Creutz Comprehensive Coordination Chemistry II: From Biology to Nanotechnology, Vol. 7, pp. 1-56 Oxford: Elsevier.]). The design of mol­ecular structures is highly influenced by factors such as the coordination nature of the metal ion, the coordin­ating ability and functionality of the organic ligands and the reaction conditions (Zhang et al., 2015[Zhang, W. X., Liao, P. Q., Lin, R. B., Wei, Y. S., Zeng, M. H. & Chen, X. M. (2015). Coord. Chem. Rev. 293-294, 263-278.]; Yin et al., 2015[Yin, Z., Zhou, Y. L., Zeng, M. H. & Kurmoo, M. (2015). Dalton Trans. 44, 5258-5275.]). Hence, the prediction of crystal structure is largely considered to be serendipitous except for simple compounds such as mononuclear mol­ecules. The 3d7configuration of CoII is particularly suited for the construction of metal–organic compounds (Kurmoo, 2009[Kurmoo, M. (2009). Chem. Soc. Rev. 38, 1353-1379.]). One of the inter­esting structural aspects of studying cobalt compared to nickel, iron or manganese is the range of coordination geometries – octa­hedral, tetra­hedral, square–pyramidal, trigonal–bipyramidal and square–planar – which are all stable (Kurmoo, 2009[Kurmoo, M. (2009). Chem. Soc. Rev. 38, 1353-1379.]). There are several coordination modes for the cobalt ion. The common mode is six-coordinate (Bryant et al., 2015[Bryant, M. R., Burrows, A. D., Fitchett, C. M., Hawes, C. S., Hunter, S. O., Keenan, L. L., Kelly, D. J., Kruger, P. E., Mahon, M. F. & Richardson, C. (2015). Dalton Trans. 44, 9269-9280.]; Artetxe et al., 2015[Artetxe, B., Reinoso, S., San Felices, L., Vitoria, P., Pache, A., Martín-Caballero, J. & Gutiérrez-Zorrilla, J. M. (2015). Inorg. Chem. 54, 241-252.]), and only relatively few four-coordinate (Gupta et al., 2015[Gupta, S. K., Kuppuswamy, S., Walsh, J. P., McInnes, E. J. & Murugavel, R. (2015). Dalton Trans. 44, 5587-5601.]) and five-coordinate (Lee et al., 2015[Lee, J. Y., Lee, J. Y. & Lee, H. M. (2015). Inorg. Chem. Commun. 52, 16-19.]) cobalt complexes have been recorded. Generally, five-coordinate compounds have two classical configurations, trigonal–bipyramidal and square–pyramidal, and the extent of each geometry each can be determined by the τ value (Addison et al., 1984[Addison, A. W., Rao, T. N., Reedijk, J., van Rijn, J. & Verschoor, G. C. (1984). J. Chem. Soc. Dalton Trans. pp. 1349-1356.]).

The carboxyl­ate unit is widely used in the synthesis of coordination compounds and is part of commonly used ligands. It is a good bridging group, favouring the formation of products (Zhou et al., 2009[Zhou, Y. L., Meng, F. Y., Zhang, J., Zeng, M. H. & Liang, H. (2009). Cryst. Growth Des. 9, 1402-1410.]). Very recently, we have been investigating CoII compounds constructed from ligands containing carboxyl­ate and hydroxyl groups, which usually form multinuclear and/or polymeric structures and show inter­esting magnetic behavior (Zhou et al., 2009[Zhou, Y. L., Meng, F. Y., Zhang, J., Zeng, M. H. & Liang, H. (2009). Cryst. Growth Des. 9, 1402-1410.]; Zeng et al., 2010[Zeng, M. H., Zhou, Y. L., Wu, M. C., Sun, H. L. & Du, M. (2010). Inorg. Chem. 49, 6436-6442.]). Similarly, herein, we chose N,N-bis­(2-hy­droxy­eth­yl)glycine (bicH3) containing two hydroxyl oxygen atoms, one carboxylate oxygen atom and one nitro­gen atom, which can potentially coordinate to a metal ion as a tetra­dentate ligand (He et al., 1999[He, X. F., Long, L. S., Le, X. Y., Chen, X. M., Ji, L. N. & Zhou, Z. Y. (1999). Inorg. Chim. Acta, 285, 326-331.]). BicH3 contains the properties of both amino acid and amino alcohol as a result of the N-substituted amino, carboxyl, and two hydroxyl groups in the mol­ecule. To the best of our knowledge, the crystal structures of metal–organic compounds with the bicH3 ligand have not been very well explored to date. Potential coordination modes for bicH2, bicH2− and bic3− are shown in Fig. 1[link]. In the course of our ongoing studies on CoII compounds containing ligands with carboxylate moieties, we have directly assembled the title compound [Co(bicH2)Cl], 1, using the flexible tetra­dentate ligand bicH3 and CoCl2·6H2O under solvothermal conditions.

[Scheme 1]
[Figure 1]
Figure 1
Coordination modes for bicH2 (a–e), bicH2− (f–h) and bic3− (i, j).

2. Structural commentary

The mol­ecular structure of the title compound is shown in Fig. 2[link]. The CoII ion is five-coordinated by four atoms from the bicH2 ligand (one carboxyl­ato oxygen atom, two hydroxyl oxygen atoms, one nitro­gen atom) and one terminal chlorine atom in a slightly distorted trigonal–bipyramidal environment (τ = 0.94, τ = |α − β|/60, α and β being the two largest angles around the central atom; values for τ in perfect coordination geometries are 1.0 for trigonal–bipyramid and 0.0 for square–pyramidal). In a similar reported compound which was formed by bicH22− and a CuII ion, a five-coordinate mode was observed (He et al., 1999[He, X. F., Long, L. S., Le, X. Y., Chen, X. M., Ji, L. N. & Zhou, Z. Y. (1999). Inorg. Chim. Acta, 285, 326-331.]); the difference is that one nitro­gen atom of benzimidazole or iso­quinoline has replaced the terminal chloride ion in compound 1. In 1, the bond lengths around the CoII ion are Co1—N1 = 2.1626 (15), Co1—O1 = 2.0482 (13), Co1—O2 = 2.0463 (14), Co1—O3 = 2.0095 (14) and Co1—Cl1 =2.2701 (6) Å. The length of the Co—O(carboxyl­ate) bond is shorter than that of Co—O(hydrox­yl), which may be due to the difference between the electron density of carboxyl­ate oxygen atoms and that of hydroxyl oxygen atoms (He et al., 1999[He, X. F., Long, L. S., Le, X. Y., Chen, X. M., Ji, L. N. & Zhou, Z. Y. (1999). Inorg. Chim. Acta, 285, 326-331.]). According to the total valence–charge balance and the bond lengths, we can conclude that cobalt is in oxidation state +II.

[Figure 2]
Figure 2
The mol­ecular structure of the title compound, showing the atom labeling. Displacement ellipsoids are drawn at the 30% probability level.

3. Supra­molecular features

In the crystal, two types of O—H⋯O hydrogen bonds (Table 1[link]) connect the mol­ecules, forming a two-dimensional network parallel to (001) (Fig. 3[link]). The O—H groups behave as donors to the non-coordinating carboxyl­ate oxygen atom of symmetry-related mol­ecules. The hy­droxy group containing O1 acts as a bifurcated O—H⋯(O,O) donor while caboxylate atom O4 is a bifurcated (O—H,O—H)⋯O acceptor.

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
O2—H2⋯O4i 0.85 (1) 1.79 (1) 2.6271 (19) 165 (2)
O1—H1⋯O4ii 0.79 (3) 1.89 (3) 2.6567 (19) 165 (3)
Symmetry codes: (i) [x-{\script{1\over 2}}, -y+{\script{3\over 2}}, -z]; (ii) -x+1, -y+1, -z.
[Figure 3]
Figure 3
Part of the crystal structure showing the two different O—H⋯O hydrogen bonds as distinct colors, blue for O1—H⋯O4ii bonds and green for O2—H2⋯O4i bonds (symmetry codes as in Table 1[link]).

4. Database survey

A search of the Cambridge Structural Database (CSD; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) was carried out for structures containing the bicH3 ligand. This revealed bicH3 coordinating to a lanthanide metal (Inomata et al., 2001[Inomata, Y., Takei, T. & Howell, F. S. (2001). Inorg. Chim. Acta, 318, 201-206.]), Cd and Na (Katsoulakou et al., 2011[Katsoulakou, E., Konidaris, K. F., Terzis, A., Raptopoulou, C. P., Perlepes, S. P., Manessi-Zoupa, E. & Kostakis, G. E. (2011). Polyhedron, 30, 397-404.]), Cu, Ni and Zn (Thakuria & Das, 2007[Thakuria, H. & Das, G. (2007). Polyhedron, 26, 149-153.]; Liu et al., 2013[Liu, Y., Kuang, D. Z., Feng, Y. L. & Fu, W. W. (2013). Transition Met. Chem. 38, 849-853.]; Lo & Ng, 2010[Lo, K. M. & Ng, S. W. (2010). Acta Cryst. E66, m1485.]), Re, Mn and Fe (Kirillov et al., 2005[Kirillov, A. M., Haukka, M., Kirillova, M. V. & Pombeiro, A. L. (2005). Adv. Synth. Catal. 347, 1435-1446.]; Sun et al., 1997[Sun, Z., Gantzel, P. K. & Hendrickson, D. N. (1997). Polyhedron, 16, 3267-3271.]; Graham et al., 2009[Graham, K., Darwish, A., Ferguson, A., Parsons, S. & Murrie, M. (2009). Polyhedron, 28, 1830-1833.]). A related structure with copper and bromide (Yamaguchi et al., 1991[Yamaguchi, H., Inomata, Y. & Takeuchi, T. (1991). Inorg. Chim. Acta, 181, 31-36.]) shows a very similar mononuclear crystal structure to the title compound. There are only a small number of reports for the ligand coordinating to Co (Funes et al., 2015[Funes, A. V., Carrella, L., Sorace, L., Rentschler, E. & Alborés, P. (2015). Dalton Trans. 44, 2390-2400.]; Zhao & Liu, 2010[Zhao, J.-P. & Liu, F.-C. (2010). Acta Cryst. E66, m848.]; Liu et al., 2015[Liu, Y., Zhou, D., Liu, H.-H. & He, C.-C. (2015). Acta Cryst. E71, m199-m200.]).

5. Synthesis and crystallization

The ligand bicH3 (0.5 mmol) in a ethanol solution (2 mL) was added to a ethanol solution (5 mL) of CoCl2·6H2O (1 mmol). 0.02 mL of tri­ethyl­amine was added dropwise to the mixed solution and stirred for 15 min at room temperature. The reactants were sealed in a 12 mL Teflon-lined autoclave, heated at 413 K for three days and then cooled to room temperature at a rate of 10 K h−1. Purple single crystals (Fig. 4[link]) were obtained along with purple powder. The crystals were picked out, washed with distilled water, and dried in air (yield ca 50.3% based on CoII). Analysis calculated (%) for C6H12ClCoNO4: C 28.30, H 3.93, N 5.50; Found C 28.31, H 3.95, N 5.54%. FTIR data for 1 (KBr, cm−1): 3383(m), 2964(w), 1593(s), 1434(m), 1407(m), 1309(w), 1058(w), 890(w).

[Figure 4]
Figure 4
The optical microscope image of single crystals of compound 1.

6. ESI–MS spectroscopic analysis

The ESI mass spectra were recorded using an LCQ–FLEET mass spectrometer (Thermo). To give further evidence for the inner structure of compound 1, characterization of the mol­ecule in solution was accomplished by ESI–MS experiments. For the methanol solution of 1, the ESI mass spectrum (Fig. 5[link]) exhibits the main ion peak observed at an m/z of 254.93, which can be assigned as [Co(C6H11NO4)Cl]+ (fit: 254.970860). The observed m/z clearly matches the assigned formula as well as the simulated spectra. This suggests that compound 1 produced in solution was stable during the ionization process. ESI–MS can also be used to examine a series of inner-bridge replacement reactions for multinuclear Co compounds (Zhou et al., 2010[Zhou, Y. L., Zeng, M. H., Wei, L. Q., Li, B. W. & Kurmoo, M. (2010). Chem. Mater. 22, 4295-4303.]; Hu et al., 2013[Hu, Y. Q., Zeng, M. H., Zhang, K., Hu, S., Zhou, F. F. & Kurmoo, M. (2013). J. Am. Chem. Soc. 135, 7901-7908.]), which is an important complement to ligand exchange, ion exchange, template exchange and supra­molecular transformations (Chakrabarty et al., 2011[Chakrabarty, R., Mukherjee, P. S. & Stang, P. J. (2011). Chem. Rev. 111, 6810-6918.]; Miras et al., 2009[Miras, H. N., Wilson, E. F. & Cronin, L. (2009). Chem. Commun. pp. 1297-1311.]). The study of the chemistry of coordination compounds by mass spectroscopy is an excellent tool to demonstrate the stability and existence of multinuclear moleculesin solution.

[Figure 5]
Figure 5
ESI mass spectrum of compound 1.

7. TG analysis

The TG analysis was performed on Pyris Diamond TG/DTA. The appearance of the flexible polydentate ligands inspired us to investigate the thermal stability of the network. The crushed single-crystal samples were heated to 1073 K in an N2 atmosphere at a heating rate of 5 K min−1 (Fig. 6[link]). The TGA curve for 1 shows that the framework begins to decompose at 413 K, and the 21.6% remaining weight is assuming to the mass loss percentage of cobalt (cal. 22.9%). One similar compound, [CuCd(bicH2(NO3)Cl2(H2O)]·H2O, constructed with bicH2 is quite unstable and begins to lose lattice water at 393 K (Liu et al., 2013[Liu, Y., Kuang, D. Z., Feng, Y. L. & Fu, W. W. (2013). Transition Met. Chem. 38, 849-853.]). Other compounds synthesized using the same ligand also show mass loss below 413 K (Inomata et al., 2001[Inomata, Y., Takei, T. & Howell, F. S. (2001). Inorg. Chim. Acta, 318, 201-206.]), owing to the loss of coordinating water. However, another reported complex with five-coordinate cobalt begins to disintegrate at 669 K (Lee et al., 2015[Lee, J. Y., Lee, J. Y. & Lee, H. M. (2015). Inorg. Chem. Commun. 52, 16-19.]). In summary, the crystals synthesized using bicH22− tend to decompose at a relatively low temperature.

[Figure 6]
Figure 6
The TG curve of compound 1.

8. DFT calculations

All the calculations were performed by using the GAUSSIAN09 program package (Frisch et al., 2009[Frisch, M. J., et al. (2009). GAUSSIAN09. Gaussian, Inc., Wallingford, CT .]). The mol­ecular structure of the title compound in the ground state was optimized in vacuo without considering the solvent, and a basis set (b1) consisting of a standard LanL2DZ basis set (Dunning & Hay, 1976[Dunning, T. H. Jr & Hay, P. J. (1976). Modern Theoretical Chemistry, Vol. 3, edited by H. F. Schaefer III, p. 1. New York: Plenum.]; Wadt & Hay, 1985[Wadt, W. R. & Hay, P. J. (1985). J. Chem. Phys. 82, 284-298.]; Hay & Wadt, 1985[Hay, P. J. & Wadt, W. R. (1985). J. Chem. Phys. 82, 270-283, 299-310.]) for Co, while the other atoms, C, H, N, O were described by a standard 6-31G(d) set. To investigate the energy differences between the high-spin and low-spin states of the title compound, the ΔE of these two energy states was evaluated using the B3LYP/b1 method (Carabineiro et al., 2008[Carabineiro, S. A., Bellabarba, R. M., Gomes, P. T., Pascu, S. I., Veiros, L. F., Freire, C., Pereira, L. C. J., Henriques, R. T., Oliveira, M. C. & Warren, J. E. (2008). Inorg. Chem. 47, 8896-8911.]; Saraçoğlu & Cukurovali, 2016[Saraçoğlu, H. & Cukurovali, A. (2016). Mol. Cryst. Liq. Cryst. 625, 173-185.]). Vibrational frequencies were calculated for all stationary points to verify that each was a minimum (NIMAG = 0) on the potential energy surface.

To gain an insight of the electronic structures, bonding properties and relative stability of the two different spin ground states (S = 3/2 and 1/2) for compound 1, the calculations in the DFT method were investigated. The optimized geometries calculated for compound 1 are presented in Fig. 7[link]. The structure of 1 presented an almost perfect trigonal–bipyramidal geometry by means of the X-ray diffraction. The chlorine and nitro­gen atoms occupy the axial positions, while the equatorial plane is occupied by three oxygen atoms. According to the energies for the two calculated structures (see Supporting information), the X-ray structure determined for complex 1 should correspond to the complex with high-spin ground state. The result can also be indicated by the mean (δ) and maximum (Δ) absolute deviations obtained for the coordination distances (Co—X). For the high-spin form of complex 1, values of δ = 0.104 Å and Δ = 0.148 Å indicate a reasonable agreement. As for the high-spin ground state and the experimental value, the biggest difference for the bond lengths is found to be 0.148 Å for Co—N, similar to what has been observed in related CoII compounds (Carabineiro et al., 2008[Carabineiro, S. A., Bellabarba, R. M., Gomes, P. T., Pascu, S. I., Veiros, L. F., Freire, C., Pereira, L. C. J., Henriques, R. T., Oliveira, M. C. & Warren, J. E. (2008). Inorg. Chem. 47, 8896-8911.]). The results of the schematic representation of both ground states supported its coordination behavior and the value of ΔE is 13.4 kcal mol−1, which shows that compound 1 can well exist stably. The cartesian coordinates for the two calculated structures are given in the Supporting information.

[Figure 7]
Figure 7
Optimized geometries (B3LYP) for the low-spin (S = 1/2, bottom) and the high-spin (S = 3/2, top) ground states and the relative energy of the two ground states (kcal mol−1).

9. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. H atoms bonded to C atoms were placed in calculated positions with C—H = 0.97 Å and Uiso(H) = 1.2Ueq(C). Hydroxyl hydrogen atoms H1 and H2 were refined independently, H1 with a refined isotropic displacement parameter and H2 with Uiso(H) = 1.5Ueq(O).

Table 2
Experimental details

Crystal data
Chemical formula [Co(C6H12NO4)Cl]
Mr 256.55
Crystal system, space group Orthorhombic, Pbca
Temperature (K) 296
a, b, c (Å) 8.3925 (9), 14.0939 (15), 15.8448 (17)
V3) 1874.2 (3)
Z 8
Radiation type Mo Kα
μ (mm−1) 2.10
Crystal size (mm) 0.84 × 0.27 × 0.24
 
Data collection
Diffractometer Bruker APEXII CCD
Absorption correction Multi-scan (SADABS; Bruker, 2004[Bruker (2004). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.])
Tmin, Tmax 0.512, 0.604
No. of measured, independent and observed [I > 2σ(I)] reflections 15130, 1933, 1854
Rint 0.021
(sin θ/λ)max−1) 0.625
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.025, 0.066, 1.14
No. of reflections 1933
No. of parameters 125
No. of restraints 3
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.51, −0.43
Computer programs: APEX2 and SAINT (Bruker, 2004[Bruker (2004). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.]), SIR2004 (Burla et al., 2007[Burla, M. C., Caliandro, R., Camalli, M., Carrozzini, B., Cascarano, G. L., De Caro, L., Giacovazzo, C., Polidori, G., Siliqi, D. & Spagna, R. (2007). J. Appl. Cryst. 40, 609-613.]), SHELXL2014 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]) and OLEX2 (Dolomanov et al., 2009[Dolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339-341.]).

Supporting information


Computing details top

Data collection: APEX2 (Bruker, 2004); cell refinement: SAINT (Bruker, 2004); data reduction: SAINT (Bruker, 2004); program(s) used to solve structure: SIR2004 (Burla et al., 2007); program(s) used to refine structure: SHELXL2014 (Sheldrick, 2015); molecular graphics: OLEX2 (Dolomanov et al., 2009); software used to prepare material for publication: OLEX2 (Dolomanov et al., 2009).

[N,N-Bis(2-hydroxyethyl)glycinato]chloridocobalt(II) top
Crystal data top
[Co(C6H12NO4)Cl]Dx = 1.818 Mg m3
Mr = 256.55Mo Kα radiation, λ = 0.71073 Å
Orthorhombic, PbcaCell parameters from 9897 reflections
a = 8.3925 (9) Åθ = 2.6–26.4°
b = 14.0939 (15) ŵ = 2.10 mm1
c = 15.8448 (17) ÅT = 296 K
V = 1874.2 (3) Å3Prism, purple
Z = 80.84 × 0.27 × 0.24 mm
F(000) = 1048
Data collection top
Bruker APEXII CCD
diffractometer
1854 reflections with I > 2σ(I)
Graphite monochromatorRint = 0.021
φ and ω scansθmax = 26.4°, θmin = 2.6°
Absorption correction: multi-scan
(SADABS; Bruker, 2004)
h = 1010
Tmin = 0.512, Tmax = 0.604k = 1717
15130 measured reflectionsl = 1919
1933 independent reflections
Refinement top
Refinement on F2Primary atom site location: structure-invariant direct methods
Least-squares matrix: fullHydrogen site location: mixed
R[F2 > 2σ(F2)] = 0.025H atoms treated by a mixture of independent and constrained refinement
wR(F2) = 0.066 w = 1/[σ2(Fo2) + (0.030P)2 + 1.540P]
where P = (Fo2 + 2Fc2)/3
S = 1.14(Δ/σ)max = 0.001
1933 reflectionsΔρmax = 0.51 e Å3
125 parametersΔρmin = 0.43 e Å3
3 restraints
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Co10.41813 (3)0.62528 (2)0.09419 (2)0.02152 (10)
Cl10.15183 (6)0.60446 (4)0.07726 (4)0.04084 (15)
O10.46124 (15)0.51884 (10)0.17919 (8)0.0254 (3)
O20.42621 (16)0.75436 (10)0.15315 (9)0.0315 (3)
H20.3575 (17)0.7980 (10)0.1454 (15)0.047*
O30.48651 (16)0.61778 (10)0.02717 (8)0.0320 (3)
O40.68239 (18)0.63508 (9)0.11936 (9)0.0305 (3)
N10.67257 (18)0.64315 (11)0.10881 (9)0.0226 (3)
C10.6259 (2)0.50535 (15)0.20204 (13)0.0319 (4)
H1A0.64810.43810.20780.038*
H1B0.64690.53560.25590.038*
C20.7323 (2)0.54745 (14)0.13522 (13)0.0308 (4)
H2A0.83980.55330.15710.037*
H2B0.73530.50560.08660.037*
C30.5816 (2)0.79707 (15)0.15957 (15)0.0367 (5)
H3A0.58520.84070.20690.044*
H3B0.60660.83180.10840.044*
C40.6996 (2)0.71678 (15)0.17250 (13)0.0336 (4)
H4A0.80750.74080.16770.040*
H4B0.68650.69010.22850.040*
C50.6287 (2)0.63779 (12)0.04522 (12)0.0240 (4)
C60.7392 (2)0.66736 (15)0.02554 (11)0.0301 (4)
H6A0.84120.63590.01860.036*
H6B0.75710.73530.02260.036*
H10.432 (3)0.468 (2)0.1653 (18)0.054 (9)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Co10.01510 (14)0.02795 (15)0.02151 (15)0.00134 (9)0.00195 (9)0.00217 (9)
Cl10.0166 (2)0.0604 (3)0.0455 (3)0.0060 (2)0.0050 (2)0.0169 (3)
O10.0233 (6)0.0265 (7)0.0265 (7)0.0020 (5)0.0014 (5)0.0011 (5)
O20.0269 (7)0.0294 (7)0.0383 (8)0.0047 (5)0.0040 (6)0.0023 (6)
O30.0217 (7)0.0523 (9)0.0220 (7)0.0095 (6)0.0014 (5)0.0024 (6)
O40.0336 (7)0.0355 (7)0.0223 (7)0.0109 (6)0.0038 (6)0.0020 (5)
N10.0169 (7)0.0310 (8)0.0199 (7)0.0025 (6)0.0026 (6)0.0015 (6)
C10.0263 (9)0.0363 (10)0.0330 (10)0.0024 (8)0.0079 (8)0.0062 (8)
C20.0228 (9)0.0359 (10)0.0337 (10)0.0036 (8)0.0030 (8)0.0022 (8)
C30.0367 (11)0.0290 (10)0.0445 (12)0.0038 (8)0.0073 (9)0.0057 (9)
C40.0280 (10)0.0406 (11)0.0323 (10)0.0047 (8)0.0069 (8)0.0066 (8)
C50.0247 (8)0.0235 (8)0.0238 (9)0.0031 (7)0.0006 (7)0.0009 (7)
C60.0217 (8)0.0443 (11)0.0243 (9)0.0090 (8)0.0001 (7)0.0027 (8)
Geometric parameters (Å, º) top
Co1—Cl12.2701 (6)C1—H1A0.9700
Co1—O12.0482 (13)C1—H1B0.9700
Co1—O22.0463 (14)C1—C21.507 (3)
Co1—O32.0095 (14)C2—H2A0.9700
Co1—N12.1626 (15)C2—H2B0.9700
O1—C11.441 (2)C3—H3A0.9700
O1—H10.79 (3)C3—H3B0.9700
O2—H20.852 (9)C3—C41.518 (3)
O2—C31.440 (2)C4—H4A0.9700
O3—C51.259 (2)C4—H4B0.9700
O4—C51.259 (2)C5—C61.514 (3)
N1—C21.498 (2)C6—H6A0.9700
N1—C41.465 (2)C6—H6B0.9700
N1—C61.473 (2)
O1—Co1—Cl199.03 (4)C2—C1—H1B109.7
O1—Co1—N180.82 (6)N1—C2—C1110.66 (16)
O2—Co1—Cl1101.63 (4)N1—C2—H2A109.5
O2—Co1—O1110.19 (6)N1—C2—H2B109.5
O2—Co1—N179.33 (6)C1—C2—H2A109.5
O3—Co1—Cl199.28 (4)C1—C2—H2B109.5
O3—Co1—O1122.70 (6)H2A—C2—H2B108.1
O3—Co1—O2118.31 (6)O2—C3—H3A110.4
O3—Co1—N180.02 (6)O2—C3—H3B110.4
N1—Co1—Cl1179.02 (5)O2—C3—C4106.80 (16)
Co1—O1—H1115 (2)H3A—C3—H3B108.6
C1—O1—Co1115.54 (11)C4—C3—H3A110.4
C1—O1—H1104 (2)C4—C3—H3B110.4
Co1—O2—H2123.6 (13)N1—C4—C3109.51 (16)
C3—O2—Co1115.72 (12)N1—C4—H4A109.8
C3—O2—H2108.8 (12)N1—C4—H4B109.8
C5—O3—Co1118.44 (12)C3—C4—H4A109.8
C2—N1—Co1104.78 (11)C3—C4—H4B109.8
C4—N1—Co1108.01 (12)H4A—C4—H4B108.2
C4—N1—C2113.16 (15)O3—C5—C6118.30 (16)
C4—N1—C6113.21 (15)O4—C5—O3123.00 (17)
C6—N1—Co1107.81 (11)O4—C5—C6118.69 (17)
C6—N1—C2109.38 (15)N1—C6—C5111.54 (15)
O1—C1—H1A109.7N1—C6—H6A109.3
O1—C1—H1B109.7N1—C6—H6B109.3
O1—C1—C2109.85 (15)C5—C6—H6A109.3
H1A—C1—H1B108.2C5—C6—H6B109.3
C2—C1—H1A109.7H6A—C6—H6B108.0
Co1—O1—C1—C221.0 (2)O3—C5—C6—N115.5 (3)
Co1—O2—C3—C435.2 (2)O4—C5—C6—N1165.10 (16)
Co1—O3—C5—O4179.97 (13)C2—N1—C4—C3156.64 (17)
Co1—O3—C5—C60.7 (2)C2—N1—C6—C592.75 (18)
Co1—N1—C2—C144.47 (17)C4—N1—C2—C173.0 (2)
Co1—N1—C4—C341.12 (19)C4—N1—C6—C5140.06 (17)
Co1—N1—C6—C520.64 (19)C6—N1—C2—C1159.82 (15)
O1—C1—C2—N144.4 (2)C6—N1—C4—C378.2 (2)
O2—C3—C4—N149.9 (2)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O2—H2···O4i0.85 (1)1.79 (1)2.6271 (19)165 (2)
O1—H1···O4ii0.79 (3)1.89 (3)2.6567 (19)165 (3)
Symmetry codes: (i) x1/2, y+3/2, z; (ii) x+1, y+1, z.
 

Acknowledgements

This work was supported by the NSFC (No. 21401030), GXNSF (No. 2014GXNSFBA118048), as well as the Guangxi Experiment Centre of Science and Technology (YXKT2014008). The computational resources are partly provided by Multifunction Computer Center of Guangxi University.

References

First citationAddison, A. W., Rao, T. N., Reedijk, J., van Rijn, J. & Verschoor, G. C. (1984). J. Chem. Soc. Dalton Trans. pp. 1349–1356.  CSD CrossRef Web of Science Google Scholar
First citationArtetxe, B., Reinoso, S., San Felices, L., Vitoria, P., Pache, A., Martín-Caballero, J. & Gutiérrez-Zorrilla, J. M. (2015). Inorg. Chem. 54, 241–252.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationBruker (2004). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBryant, M. R., Burrows, A. D., Fitchett, C. M., Hawes, C. S., Hunter, S. O., Keenan, L. L., Kelly, D. J., Kruger, P. E., Mahon, M. F. & Richardson, C. (2015). Dalton Trans. 44, 9269–9280.  CSD CrossRef CAS PubMed Google Scholar
First citationBurla, M. C., Caliandro, R., Camalli, M., Carrozzini, B., Cascarano, G. L., De Caro, L., Giacovazzo, C., Polidori, G., Siliqi, D. & Spagna, R. (2007). J. Appl. Cryst. 40, 609–613.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCarabineiro, S. A., Bellabarba, R. M., Gomes, P. T., Pascu, S. I., Veiros, L. F., Freire, C., Pereira, L. C. J., Henriques, R. T., Oliveira, M. C. & Warren, J. E. (2008). Inorg. Chem. 47, 8896–8911.  Web of Science CSD CrossRef PubMed CAS Google Scholar
First citationChakrabarty, R., Mukherjee, P. S. & Stang, P. J. (2011). Chem. Rev. 111, 6810–6918.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDolomanov, O. V., Bourhis, L. J., Gildea, R. J., Howard, J. A. K. & Puschmann, H. (2009). J. Appl. Cryst. 42, 339–341.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationDunning, T. H. Jr & Hay, P. J. (1976). Modern Theoretical Chemistry, Vol. 3, edited by H. F. Schaefer III, p. 1. New York: Plenum.  Google Scholar
First citationFrisch, M. J., et al. (2009). GAUSSIAN09. Gaussian, Inc., Wallingford, CT .  Google Scholar
First citationFujita, M., Powell, A. & Creutz, C. E. (2004). Creutz Comprehensive Coordination Chemistry II: From Biology to Nanotechnology, Vol. 7, pp. 1–56 Oxford: Elsevier.  Google Scholar
First citationFunes, A. V., Carrella, L., Sorace, L., Rentschler, E. & Alborés, P. (2015). Dalton Trans. 44, 2390–2400.  CSD CrossRef CAS PubMed Google Scholar
First citationGraham, K., Darwish, A., Ferguson, A., Parsons, S. & Murrie, M. (2009). Polyhedron, 28, 1830–1833.  Web of Science CSD CrossRef CAS Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationGupta, S. K., Kuppuswamy, S., Walsh, J. P., McInnes, E. J. & Murugavel, R. (2015). Dalton Trans. 44, 5587–5601.  CSD CrossRef CAS PubMed Google Scholar
First citationHay, P. J. & Wadt, W. R. (1985). J. Chem. Phys. 82, 270–283, 299–310.  Google Scholar
First citationHe, X. F., Long, L. S., Le, X. Y., Chen, X. M., Ji, L. N. & Zhou, Z. Y. (1999). Inorg. Chim. Acta, 285, 326–331.  CSD CrossRef CAS Google Scholar
First citationHu, Y. Q., Zeng, M. H., Zhang, K., Hu, S., Zhou, F. F. & Kurmoo, M. (2013). J. Am. Chem. Soc. 135, 7901–7908.  CSD CrossRef CAS PubMed Google Scholar
First citationInomata, Y., Takei, T. & Howell, F. S. (2001). Inorg. Chim. Acta, 318, 201–206.  Web of Science CSD CrossRef CAS Google Scholar
First citationKatsoulakou, E., Konidaris, K. F., Terzis, A., Raptopoulou, C. P., Perlepes, S. P., Manessi-Zoupa, E. & Kostakis, G. E. (2011). Polyhedron, 30, 397–404.  Web of Science CSD CrossRef CAS Google Scholar
First citationKirillov, A. M., Haukka, M., Kirillova, M. V. & Pombeiro, A. L. (2005). Adv. Synth. Catal. 347, 1435–1446.  CSD CrossRef CAS Google Scholar
First citationKurmoo, M. (2009). Chem. Soc. Rev. 38, 1353–1379.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLee, J. Y., Lee, J. Y. & Lee, H. M. (2015). Inorg. Chem. Commun. 52, 16–19.  CSD CrossRef CAS Google Scholar
First citationLiu, Y., Kuang, D. Z., Feng, Y. L. & Fu, W. W. (2013). Transition Met. Chem. 38, 849–853.  CSD CrossRef CAS Google Scholar
First citationLiu, Y., Zhou, D., Liu, H.-H. & He, C.-C. (2015). Acta Cryst. E71, m199–m200.  CSD CrossRef IUCr Journals Google Scholar
First citationLo, K. M. & Ng, S. W. (2010). Acta Cryst. E66, m1485.  CSD CrossRef IUCr Journals Google Scholar
First citationMiras, H. N., Wilson, E. F. & Cronin, L. (2009). Chem. Commun. pp. 1297–1311.  CrossRef Google Scholar
First citationSaraçoğlu, H. & Cukurovali, A. (2016). Mol. Cryst. Liq. Cryst. 625, 173–185.  Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSun, Z., Gantzel, P. K. & Hendrickson, D. N. (1997). Polyhedron, 16, 3267–3271.  CSD CrossRef CAS Google Scholar
First citationThakuria, H. & Das, G. (2007). Polyhedron, 26, 149–153.  Web of Science CSD CrossRef CAS Google Scholar
First citationWadt, W. R. & Hay, P. J. (1985). J. Chem. Phys. 82, 284–298.  CrossRef CAS Web of Science Google Scholar
First citationYamaguchi, H., Inomata, Y. & Takeuchi, T. (1991). Inorg. Chim. Acta, 181, 31–36.  CSD CrossRef CAS Google Scholar
First citationYin, Z., Zhou, Y. L., Zeng, M. H. & Kurmoo, M. (2015). Dalton Trans. 44, 5258–5275.  CrossRef CAS PubMed Google Scholar
First citationZeng, M. H., Zhou, Y. L., Wu, M. C., Sun, H. L. & Du, M. (2010). Inorg. Chem. 49, 6436–6442.  Web of Science CSD CrossRef CAS PubMed Google Scholar
First citationZhang, W. X., Liao, P. Q., Lin, R. B., Wei, Y. S., Zeng, M. H. & Chen, X. M. (2015). Coord. Chem. Rev. 293–294, 263–278.  CrossRef CAS Google Scholar
First citationZhao, J.-P. & Liu, F.-C. (2010). Acta Cryst. E66, m848.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationZhou, Y. L., Meng, F. Y., Zhang, J., Zeng, M. H. & Liang, H. (2009). Cryst. Growth Des. 9, 1402–1410.  CSD CrossRef CAS Google Scholar
First citationZhou, Y. L., Zeng, M. H., Wei, L. Q., Li, B. W. & Kurmoo, M. (2010). Chem. Mater. 22, 4295–4303.  CSD CrossRef CAS Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds