Synthesis, Spectroscopic Studies and X-Ray Diffraction of Heptacoordinated Mn(II) and Co(II) Complexes with Ligands Derived from Carbonohydrazide

Abstract

The ligand 1-(1-(pyridin-2-yl)ethylidene)carbonohydrazide (H4L1) and 1-(pyridin-2-ylmethylene)carbonohydrazide (H4L2) were prepared by reaction of carbonohydrazide with 2-acetylpyridine or pyridine carbaldehyde respectively in a reflux methanol solution. The complexes are prepared by reaction of the ligand with the appropriate metal salt. These complexes are well characterized by elemental analysis, IR and UV spectroscopies and their structure were determined by single X-ray diffraction technic. In the crystal of the dinuclear complex [Mn2(H4L1)2(H2O)4]·Cl4, 1) each Mn(II) center is seven coordinated by two nitrogen atom and one carbonyl atom of the one ligand and one nitrogen atom and one carbonyl oxygen atom of another ligand molecule. The coordination sphere is completed by two water molecules. Each of the carbonyl oxygen atom acts as bridge between the two Mn(II) centers. In the mononuclear complex [Co(H4L2)(NO3)(H2O)2]·(NO3); 2) the Co(II) center is heptacoordinated. The ligand acts in tridentate fashion through two nitrogen atoms and one carbonyl oxygen atom. Two water molecules lie in the apical positions. One nitrate group acts in bidentate manner while the other nitrate group remains uncoordinated. In both complexes the coordination polyhedral are best described as a pentagonal bipyramid. The molecules are linked together in each case by multiple hydrogen bond interaction resulting in a three-dimensional network. The crystallographic data has been deposited in Cambridge Crystallographic Data Centre [CCDC No. 1944387 (complex 1) and 1944386 (complex 2)].

Share and Cite:

Seck, T. , Sy, A. , Lo, D. , Gaye, P. , Sall, M. , Diouf, O. , Diaw, M. and Gaye, M. (2019) Synthesis, Spectroscopic Studies and X-Ray Diffraction of Heptacoordinated Mn(II) and Co(II) Complexes with Ligands Derived from Carbonohydrazide. Open Journal of Inorganic Chemistry, 9, 35-52. doi: 10.4236/ojic.2019.94004.

1. Introduction

Transition metal coordination compounds with Schiff bases continue to attract the attention of researchers because of the easier synthesis, the stability of the molecular structures and the large diverse properties observed [1] - [11] . Indeed catalytic [12] [13] [14] , optical [15] , magnetic [16] [17] , fluorescent [18] [19] and biological activities [8] [20] [21] [22] are reported. The complexes derived from the first row of transition metals can give several models of metalloenzymes [23] [24] and metalloproteins [25] [26] . Heptacoordinated manganese (II) and cobalt (II) structures have also attracted a great deal of interest from chemists in recent years [8] [27] [28] [29] . These kind of structures are present in enzymes: catalase [30] [31] , ribonucleotide manganese reductase [32] and arginase [33] . In addition to their interests in biology, these complexes are likely to give very interesting magnetic properties. In the literature, dinuclear compounds of manganese or cobalt which exhibit ferromagnetic or antiferromagnetic interactions are largely reported [34] [35] [36] . In recent years, we have prepared different symmetrical Schiff base types from the condensation of carbono or thiocarbono hydrazide and aldehydes or ketones which are subsequently used to prepare dinuclear or tetranuclear metal complexes [37] [38] [39] . In the absence of any particular precautions, the reaction which is a nucleophilic addition of the primary amine on the carbonyl group, proceeds on both arms of the carbonohydrazide. By controlling the conditions of the reaction we prepared a ligand from the monocondensation of carbonohydrazide and acetyl pyridine or pyridine carbaldehyde. The topologies of these dissymmetric ligands lead it to act in tridentate or tetradentate fashions leading fascinating structures. In this context, we report the isolation and complete characterization of this ligand and its Mn(II) and Co(II) complexes in which the metal centers are heptacoordinated as showed by crystal X-ray diffraction analysis. The structure of the formed crystals has been by single X-ray diffraction analysis and the data have been deposited in CCDC Cambridge crystallographic data centre.

2. Experimental

2.1. Material and Physical Measurement

Carbonohydrazide (98%), 2-acetylpyridine (99%), 2-pyridinecarboxaldehyde (99%), MnCl2∙4H2O (99%), Co(NO3)2∙6H2O (99.99%) and analytical grade methanol were supplied from Aldrich. All chemicals and solvents were used directly without further purification. Elemental analyses of C, H and N were recorded on a VxRio EL Instrument. The FAB mass spectra were recorded using a Micromass Autospec spectrometer using 3-nitrobenzylalcohol as the matrix. Infrared spectra were obtained on an FTIR Spectrum Two of Perkin Elmer spectrometer in the 4000 - 400 cm−1 region. The UV-Visible spectra was recorded on a Perkin Elmer Lambda UV-Vis spectrophotometer. The molar conductance of 1 × 10−3 M in DMSO solutions of the metal complexes was measured at 25˚C using a WTW LF-330 conductivity meter with a WTW conductivity cell. Room temperature magnetic susceptibilities of the powdered samples were measured using a Johnson Mattey scientific magnetic susceptibility balance (Calibrant: Hg [Co(SCN)4]). The voltammetric measurements were recorded using a Palm Sens3 type potentiostat controlled by PSTrace software. A glassy carbon working electrode with a radius of 2 ± 0.1 mm, a stainless steel wire as a counter-electrode and Ag/AgCl electrode as a reference were used. Cyclic voltammetry experiments were performed in acetonitrile solution 0.1 M of tetrabutylammonium hexafluorophosphate (TBAHFP) as a supporting electrolyte.

2.2. Synthesis of the Ligand 1-(1-(Pyridin-2-yl)Ethylidene)Carbonohydrazide: H4L1

To a solution of carbonohydrazide (3.0 g, 0.333 mmol) in a mixture of 10 mL of distillate water and 30 mL of methanol was added dropwise a solution of 2-acetylpyridine (4.038 g, 0.330 mmol) in 10 mL of methanol. The mixture was stirred under reflux for 4 hours. A white precipitate appears gradually. On cooling, the precipitate was isolated by filtration and successively washed with 2 × 10 mL of hot methanol and dried under P4O10. The filtrated was stored at 4˚COn standing for one week crystals suitable for X-ray analysis were isolated. MP: 222˚C. Yield: 86.4%. Analytical for C8H11N5O: Calc (found) %C = 49.73 (49.43); %H = 5.74 (5.78); %N = 36.25 (36.21). IR (υ, cm−1): 3306; 3086; 1671; 1629; 1578; 1506; 1466, 1141. 1H NMR (dmso-d6, δ (ppm)): 2.36 (s, 3H, -CH3); 4.12 (s, 2H, -NH2); 7.32 - 8.51 (m, 4H, HPy); 8.19 (s, 1H, -(C=O)-NH-NH2); 9.64 (s, 1H, -(C=O)-NH-(C=N)-). 13C NMR (dmso-d6, δ (ppm)):157.32 (CH3-C=N-); 155.30 (C=O); 148.37 (Cipso); 145.45 (CPy); 136.43 (CPy); 123.47 (CPy); 120.13 (CPy); 11.03 (-CH3).

2.3. Synthesis of the Ligand 1-(Pyridin-2-Ylmethylene)Carbonohydrazide: H4L2

To a solution of carbonohydrazide (3.0 g, 0.333 mmol) in a mixture of 10 mL of distillate water and 30 mL of methanol was added dropwise a solution of 2-pyridinecarboxaldehyde (3.5672 g, 0.330 mmol) in 10 mL of methanol. The mixture was stirred at reflux for 4 hours. After cooling to room temperature, the orange solution was stored in the refrigerator and a precipitate appears overnight. The compound was isolated by filtration and washed successively with 2 x 10 mL of cold methanol and 2 × 10 mL of diethyl ether before being dried over P4O10. MP: 165˚C - 167˚C. Yield: 62.1%. Analytical for C7H9N5O: Calc (found) %C = 46.92 (46.90); %H = 5.06 (5.04); %N = 39.09 (39.05). IR (υ, cm−1): 3304; 3197; 1675; 1634; 1609; 1585; 1519; 1494, 1151. 1H NMR (dmso-d6, δ (ppm)): 4.08 (s, 2H, -NH2); 7.30 - 8.52 (m, 4H, HPy); 8.21 (s, 1H, H-C=N-NH-C=O); 10.67 (s, 1H, -(C=O)-NH-NH2-). 13C NMR (dmso-d6, δ (ppm)): 156.27 (CH3-C=N-); 155.500 (C=O); 148.79 (Cipso); 140.46 (CPy); 136.47 (CPy); 123.54 (CPy); 119.50 (CPy) (Scheme 1).

2.4. Synthesis of the Complexes

2.4.1. [Mn2(H4L1)2(H2O)4]∙Cl4 (1)

To a suspension of L1H4 (0.193 g, 1 mmol) in 10 ml of methanol was added MnCl2∙4H2O (0.1979 g, 1 mmol). The resulting yellowish mixture was stirred for 30 min. The resulting yellow suspension was stirred at room temperature during one hour. The suspension was filtered and the filtrate was left at room temperature. On standing for one week crystals suitable for X-ray analysis were isolated.

Scheme 1. Preparation of the ligands and the complexes.

Yield: 74.6 (%). SM (ESI-TOF) (m/z): 194.1 [(H5L)]+; 361 [Mn(H4L)Cl(H2O)3Na]2+; 440 [Mn(H4L)(H3L)]3+; 602.97 [Mn2(H4L)2(Cl)(H2O)4]3+. IR (cm−1): 3276 cm−1; 3186 cm−1; 3142 cm−1; 3070 cm−1; 1658 cm−1; 1595 cm−1; 1557 cm−1; 1470 cm−1; 1362 cm−1; 1185 cm−1; 1154 cm−1; 1027 cm−1; 829 cm−1. UV-vis (Solution, DMF, nm): 219; 357; 421. µeff = 7.72 µB. Λ (Ω−1∙cm2∙mol−1, DMF): 292.

2.4.2. [Co(H4L2)(NO3)(H2O)2]∙(NO3) (2)

To a suspension of L2H4 (0.1792 g, 1 mmol) in 10 ml of methanol was added Co(NO3)2∙6H2O (0.2910 g, 1 mmol) dissolved in 5 mL of methanol. The resulting brown mixture was stirred for 60 min at room temperature. The resulting solution was filtered and the filtrate was left at room temperature. On standing for one week crystals suitable for X-ray analysis were isolated. Yield: 36.8 (%). IR (cm−1): 3227 cm−1; 1637 cm−1; 1578 cm−1; 1540 cm−1; 1455 cm−1; 1402 cm−1; 1380, 1294 cm−1; 1205 cm−1; 1116 cm−1; 1026 cm−1; 936 cm−1. UV-vis (Solution, DMF, nm): 225; 321; 421; 530. µeff = 4.99 µB. Λ (Ω−1∙cm2∙mol−1, DMF): 83.1.

2.5. Crystal Structure Determination and Refinement

Crystals suitable for X-diffraction, of the reported compounds, were grown by slow evaporation of MeOH solution of the complex. Details of the X-rays crystal structure solution and refinement are given in Table 1. Diffraction data were collected using an ENRAF NONIUS Kappa CCD diffractometer with graphite monochromatized MoKα radiation (λ = 0.71073 Å). All data were corrected for Lorentz and polarization effects. No absorption correction was applied. Complex scattering factors were taken from the program package SHELXTL [40] . The structures were solved by direct methods which revealed the position of all non-hydrogen atoms. All the structures were refined on F2 by a full-matrix least-squares procedure using anisotropic displacement parameters for all non-hydrogen atoms [34] . The hydrogen atoms of water molecules and NH groups were located in the Fourier difference maps and refined. Others H atoms (CH and CH3 groups) were geometrically optimized and refined as riding model by AFIX instructions. Molecular graphics were generated using ORTEP-3 [41] .

3. Results and Discussion

3.1. General Study

The infrared spectrum of each molecule ligand shows the presence of two bands in the range 3310 cm−1 and 3080 cm−1 attributable to the vibrations υN-H of the ligand. The vibration bands relative to υC=O and υC=N are respectively located at 1671 cm−1 and 1634 cm−1 for H4L1 and 1675 cm−1 and 1639 cm−1 for H4L2. Bands assignable υC=N and υC=C of the pyridine ring are located between 1460 cm−1 and 1590 cm−1 [42] . The vibration of the N-N bond are pointed at ca.1146 cm−1 for both ligands. The 1H NMR spectra of the ligands recorded in dmso-d6 and show similar signals. The difference is due to the presence of a methyl group on the imino carbon atom for L1H4 instead of imino hydrogenatom as for L2H4. The

Table 1. Crystal data and data collection details of the complexes (1) and (2).

presence of a singlet signal incorporating two protons attributable to NH2 at 4.08 ppm for both spectra in indicative of the mono condensation of the carbono hydrazide with the ketoprecursors. The 13C NMR spectra are similar except the presence of a signal at 11.03 ppm relative to the presence of a methyl group in H4L1. The FAB mass spectrometry of the compound [Mn2(H4L1)2(H2O)4]∙Cl4 gave clear evidence of fragmentation of the complex. The parent ion expected at 710.16 amu was not observed. The observed mass/charge ratio of 602.97 amu, 440 amu, 361 amu and 194.10 amu compared to the calculated exact mass of 603.08 amu, 440.12 amu, 360.63 amu and 194.21 amu are respectively due to the fragments [Mn2(H4L)2(Cl)(H2O)4]3+, [Mn(H4L)(H3L)]3+, [Mn(H4L)Cl(H2O)3Na]2+, and [(H5L)]+. Upon coordination the υC=O and the υC=N were strongly shifted to low frequencies in both spectra. For the Mn(II) complex the υC=O and the υC=N are respectively pointed at 1658 cm−1 and 1595 cm−1 while the spectrum of the Co(II) complex show the υC=O and the υC=N band at 1637 cm−1 and 1578 cm−1. These shifts are indicative of the involvement of the carbono oxygen atom and the nitrogen atom of the azomethine in the complexation with Mn(II) and Co(II) in both cases [43] [44] . In both spectrum broad absorption bands appear at 3276 cm−1 and 3227 cm−1 for the complexes of Mn(II) and Co(II) respectively. These bands are due to the presence of a coordinated water molecule [45] . Additional bands due to the nitrate moieties appears. The sharp and intense band at 1380 cm−1 is attributed to the free nitrate group. The three absorption bands occurring at 1455 cm−1, 1294 cm−1 and 1026 cm−1 respectively attributable to υ1N=O), υ5asNO2) and υ2sNO2) clearly identify a nitrate group coordinated in bidentate fashion [37] . In fact the separation Δυ = υ1 –υ5 is used as criterion of differentiation of the different modes of coordination of the nitrate moiety. The Δυ value of 161 cm−1 is in accordance with the presence of bidentate nitrate group [45] . In the UV-vis spectra of the monohydrazone ligands two bands are observed at 233 and 318 nm for H4L1 and at 235 and 315 nm for H4L2. For both spectra the first band is attributable to the n → π* transitions in aromatic rings while the second band is assignable to the n → π* transition. The electronic spectra of the manganese complex in DMF shows three main bands at 219, 357 and 421 nm and the cobalt (II) complex in DMF exhibits three main bands around 225, 321 and 421 nm and a shoulder around 530 nm. By comparing these bands to the absorption in the spectra of the free ligand, bathochromic shifts is observed for the maxima bands of the free ligand. The band observed at 357 nm for Mn(II) complex and the band pointed at 321 nm for Co(II) complex are assigned to the aromatic ring transition π → π* and the other bands are attributed to the n → π* transitions of C=N chromophore. The values of the conductance of the two complexes in DMF are 192 Ω−1∙cm2∙mol−1 for the manganese (II) complex which is 2:1 electrolyte and 83.1 Ω−1∙cm2∙mol−1 for the cobalt (II) complex which is 1:1 electrolyte in nature [46] . The magnetic susceptibility measurements of the complexes were performed at room temperature. The value of 7.72 µB for the manganese complex is less than the expected value of ca.10 µB for two Mn(II) ions, indicating the presence of two metal atoms per molecule which are in weak antiferromagnetic coupling interaction. For the complex of cobalt (II) the magnetic moment value of 4.99 µB is in the range expected for a mononuclear hepta coordinated cobalt (II) complex [28] .

3.2. Description of the Structures

Suitable single-crystals for X-ray diffraction of the Mn(II) and Co(II) complexes were obtained by slow solvent evaporation at room temperature. Crystal data, collection and refinement parameters are listed in Table 1. Selected bond lengths and angles are summarized in Table 2. Hydrogen bond data for complexes 1 and 2 are gathered in Table 3. Figures 1-4 displays the asymmetric unit and the packing diagrams of the two compounds.

Table 2. Selected bond lengths (Å) and angles (˚) for the complexes (1) and (2).

Symmetry code: (i) −x + 1, −y + 1, −z + 1.

Table 3. Hydrogen bonds for the complexes (1) and (2).

Symmetry codes: (i) −x, −y + 1, −z + 1; (ii) x, y + 1, z; (iii) −x + 1, −y + 1, −z; (iv) −x + 1, −y, −z + 1; (v) −x + 1, −y + 1, −z + 1; (vi) −x,1−y,1−z; (vii) 1 − x, −y, 2 − z.

Figure 1. ORTEP plot (30% probability ellipsoids) showing the molecule structure of the complex 1.

Figure 2. Hydrogen bonds in complex 1.

Figure 3. ORTEP plot (30% probability ellipsoids) showing the molecule structure of the complex 2.

Figure 4. Hydrogen bonds in complex 2.

3.2.1. Complex 1

The [Mn2(H4L1)2(H2O)4]∙4Cl complex crystallizes in the triclinic system with the space group P-1. The view of the structure of the complex 1 is depicted in Figure 1. The asymmetric unit contains two manganese (II) atoms, two organic ligand molecules, four coordinated water molecules and four free chloride anions. Each of the ligand molecules acts in tetradentate fashion through one nitrogen atom from the pyridine, one imino nitrogen atom, one nitrogen atom amino and one carbonyl oxygen atom which acts as bridge between the two manganese (II) atoms. Each of the two metal ions Mn(II) is heptacoordinated and the coordination polyhedron of the Mn(II) center is best described as a distorted pentagonal bipyramid with a MnN3O4 chromophore. Each equatorial planes around each Mn(II) atom is occupied by a pyridine nitrogen atom, an imino nitrogen atom, one oxygen atom from the same ligand, one carbonyl bridged oxygen atom and one amino nitrogen atom from another ligand molecule. The apical positions are occupied by two oxygen atoms of the coordinated water molecules. The angles formed between the metal center and the ligand donor atoms in equatorial position are different from the ideal angle for a regular pentagon (72˚) [N5i-Mn1-N4i = 68.80 (4)˚, O1i-Mn1-N4i = 66.56 (4)˚, N5i-Mn1-N1 = 92.20 (4)˚, O1-Mn1-N1 = 67.72 (4)˚, O1-Mn1-O1i = 65.78 (4)˚] with a sum of 361.1˚ [47] . The apical angle formed by the water oxygen atoms in axial positions around the manganese (II) atom [O2w-Mn1-O3w = 169.82 (5)˚] significantly deviate from the ideal value of 180˚ (Table 2). The angles Mn1-O1-Mn1i and O1-Mn-O1i are respectively 114.22 (4)˚ and 65.78 (4)˚. For the Mn atom the apical position is occupied by the water oxygen atoms Ow2 and OW3 with 2.1606 (11) Å and 2.1454 (11) Å distance. These distances are shorter than those of the equatorial bonds [Mn1-O1 = 2.2135 (10) Å, Mn1-N5i = 2.2968 (12) Å, Mn1-O1i = 2.3219 (9) Å, Mn1-N4i = 2.3247 (11) Å and Mn1-N1 = 2.5223 (13) Å] (Table 2) and are comparable to those found in similar complex [27] . The interatomic distance O1-C1 of 1.2433 (16) Å is close proximity to that observed in the structure of the free ligand which is of the order of 1.231 (3) Å. This fact is indicative of double character bond upon complexation with the Mn(II) ion. Each mononuclear Mn(II) molecule complex is connected to another complex molecule by hydrogen bonding interactions leading to the formation of a supramolecular chain structure propagating along the a axis. The chains are connected by hydrogen bonding interaction of type N-H…Cl, C-H…Cl, Ow-H…Cl, and C-H…N. The uncoordinated chloride anion acts via hydrogen bonds as bridge between the chains leading to the formation of a 3D supramolecular structure (Figure 2, Table 3).

3.2.2. Complex 2

The compound [Co(H4LD)(H2O)2(NO3)]∙NO3 crystallizes in the triclinic system with a P-1 space group. The view of the structure of the complex 2 is depicted in Figure 3. The asymmetric unit is composed by a cobalt (II) atom coordinated to one organic ligand molecule, one bidentated coordinated nitrate anion, two coordinated water molecules and one uncoordinated nitrate ion. The ligand acts in tridentate manner through the oxygen atom of the carbonyl moiety, the imino nitrogen atom and the pyridine nitrogen atom. The cobalt ion is situated in a N2O5 core and the environment is best described as pentagonal bipyramid geometry (Figure 3). The equatorial plane around the Co(II) ion is occupied by the three donor atoms from the organic ligand molecule and the two oxygen atoms of the coordinated nitrate group while the apical positions are occupied by the two oxygen atoms of the coordinated water molecules. The angles formed between the equatorial atoms and the Co(II) ion are: N0aa-Co1-O1 = 71.50 (8)˚, N0aa-Co1-N1 = 74.30 (8)˚, N1-Co1-O6 = 55.96 (8)˚, O5-Co1-O1 = 75.71 (8)˚ and are different from the ideal angle value of 72˚ for a pentagonal basal plan [48] . The angle O4-Co1-O3 formed between the apical atoms and the Co(II) ion deviated from the ideal value of 180˚ with a value of 175.31 (8)˚. The sum of the angles around Co(II) is 360.12˚ and is very close to the ideal sum of 360˚. These deviations are indicative of a distortion of geometry around the cobalt center which is best described as a distorted pentagonal bipyramid. The equatorial distances [Co1-O1 = 2.267 (2) Å, Co1-O5 = 2.212 (2) Å, Co1-O6 = 2.298 (3) Å, Co1-N1 = 2.192 (2) Å and Co1-N0aa = 2.119 (2) Å] are longer that of the axial distances [Co1-O3 = 2.104 (2) Å, Co1-O4 = 2.092 (2) Å] (Table 2) and are comparable to the values reported for a seven coordinated cobalt(II) complex [29] .

In the crystal, the complex molecules are linked by hydrogen bonds, giving rise to a three-dimensional network (Figure 4, Table 3). The structure is built up from pentagonal bipyramids around the Co(II) atom, which are assembled in layers parallel to the bc plane. These layers are interconnected by hydrogen bonds. The coordinating axial water molecule points into the interlayer space and acts as a hydrogen-bond donor towards oxygen atom O-NO2 of the coordinated nitrate (Figure 4) O1W-H1WB…ONO2. Within a layer, the molecules are not interconnected by hydrogen bonds.

3.2.3. Electrochemistry Study [Mn2(H4L)2(H2O)2]∙4Cl

The electrochemical properties of the Mn(II) complex was studied by cyclic voltammeter using a graphite carbon working electrode and a Pt-wire auxiliary electrode in acetonitrile using LiClO4 a the supporting electrolyte. The voltammetric parameters were studied in the scan rate interval 100 - 200 mVs−1. In Figure 5 describes the electrochemical behavior of the complex in acetonitrile medium. The cyclic votammogram of the complex [Mn2(H4L)2(H2O)2]∙4Cl was recorded at −3.00 to +3.00 V (vs Ag/AgCl) with a scan rate of 100 mV/s in DMF solution at room temperature. We note one oxidation and two reduction pics and two oxidation (Figure 5). The cathodic wave shows two distinct reductive peaks at ca. −0.9 v and −1.3 V (ΔE = 0.4 V) which correspond to two electrons charge

Figure 5. Cyclic Votammograms of the complex in CH3CN containing 0.1 M [NBu4]∙[PF6]. Working electrode: a 5 mm diameter of graphitecarbon electrode, scan rate: 100 m∙Vs−1.

Figure 6. Cyclic Votammograms of the complex in CH3CN containing 0.1 M [NBu4]∙[PF6] Scan rate: 25, 150, 100, 200, 300 and 500 mVs−1. Working electrode: a 2 mm diameter of graphite carbon electrode.

transfer process Mn(II)Mn(II) → Mn(II)Mn(I) and Mn(II)Mn(I) → Mn(I)Mn(I). The anodic wave shows two oxidative peaks −1.00 and +1.2 V corresponding to the reversible process Mn(I)Mn(I) → Mn(II)Mn(I) and Mn(II)Mn(I) → Mn(II)Mn(II). A reversible transition Mn(II) → Mn(I) had already been observed for a Mn(II) complex in acetonitrile media [49] . By varying the scans speed the intensities of all the peaks increase with the scanning speed (Figure 6).

4. Conclusion

We describe here one seven coordinated manganese (II) complexes of the tetradentate ligand 1-(1-(pyridin-2-yl)ethylidene)carbonohydrazide and one seven coordinated cobalt(II) of the tridentate 1-(pyridin-2-ylmethylene)carbonohydrazide. The single crystal structure of the Mn(II) complex 1 as well as of the Co(II) complex 2 has revealed a distorted pentagonal-bipyramidalgeometry. The equatorial plane in complex 1 is occupied by N2O and NO from two tetradentate ligand molecules acting as bridges between two Mn(II) ions. The axial positions are occupied by two coordinated water molecules. In the complex 2, the equatorial plane of the bipyramid is occupied by the tridentate N2O ligand and two oxygen atoms from the coordinated nitrate anion which acts in bidentate chelating fashion, the two apical positions being occupied by two coordinated water molecules. The cyclic voltammetry studies in CH3CN show that the manganese (II) complex 2 shows two electrons charge transfer process.

Acknowledgements

The authors thank the FONDATION SONATEL for his financial support http://fondationsonatel.sn/.

Supporting Information

CCDC-1944387 and 1944386 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures/, or by e-mailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44(0)1223-336033.

Conflicts of Interest

The authors declare no conflicts of interest regarding the publication of this paper.

References

[1] Ghosh, M., Layek, M., Fleck, M., Saha, R. and Bandyopadhyay, D. (2015) Synthesis, Crystal Structure and Antibacterial Activities of Mixed Ligand Copper(II) and Cobalt(II) Complexes of a NNS Schiff Base. Polyhedron, 85, 312-319.
https://doi.org/10.1016/j.poly.2014.08.014
[2] Khorshidifard, M., Rudbari, H.A., Kazemi-Delikani, Z., Mirkhani, V. and Azadbakht, R. (2015) Synthesis, Characterization and X-Ray Crystal Structures of Vanadium(IV), Cobalt(III), Copper(II) and Zinc(II) Complexes Derived from an Asymmetric Bidentate Schiff-Base Ligand at Ambient Temperature. Journal of Molecular Structure, 1081, 494-505.
https://doi.org/10.1016/j.molstruc.2014.10.071
[3] Gorczyński, A., Zaranek, M., Witomska, S., Bocian, A., Stefankiewicz, A.R., Kubicki, M., Patroniak, V. and Pawluć, P. (2016) The Cobalt(II) Complex of a New Tridentate Schiff-Base Ligand as a Catalyst for Hydrosilylation of Olefins. Catalysis Communications, 78, 71-74.
https://doi.org/10.1016/j.catcom.2016.02.009
[4] Jing, C., Wang, C., Yan, K., Zhao, K., Sheng, G., Qu, D., Niu, F., Zhu, H. and You, Z. (2016) Synthesis, Structures and Urease Inhibitory Activity of Cobalt(III) Complexes with Schiff Bases. Bioorganic & Medicinal Chemistry, 24, 270-276.
https://doi.org/10.1016/j.bmc.2015.12.013
[5] Pogány, L., Moncol, J., Gál, M., Šalitroš, I. and Boča, R. (2017) Four Cobalt(III) Schiff Base Complexes—Structural, Spectroscopic and Electrochemical Studies. Inorganica Chimica Acta, 462, 23-29.
https://doi.org/10.1016/j.ica.2017.03.001
[6] Wang, X.-W. and Zheng, Y.-Q. (2007) A Dinuclear Copper(II) Complex and a Zigzag Chain Iron(II) Polymer Based on the 4-Antipyrine Derived Schiff Base Ligands: The Hydroxylation and Redox Occurred under the Solvothermal Conditions. Inorganic Chemistry Communications, 10, 709-712.
https://doi.org/10.1016/j.inoche.2007.03.008
[7] Zhang, E., Wei, Y., Huang, F., Yu, Q., Bian, H., Liang, H. and Lei, F. (2018) Synthesis, Crystal Structure and Bioactivity of Manganese Complexes with Asymmetric Chiral Schiff Base. Journal of Molecular Structure, 1155, 320-329.
https://doi.org/10.1016/j.molstruc.2017.11.017
[8] Soliman, S.M., El-Faham, A., Elsilk, S.E. and Farooq, M. (2018) Two Heptacoordinated Manganese(II) Complexes of Giant Pentadentate s-Triazine Bis-Schiff Base Ligand: Synthesis, Crystal Structure, Biological and DFT Studies. Inorganica Chimica Acta, 479, 275-285.
https://doi.org/10.1016/j.ica.2018.04.043
[9] Rathi, S., Maji, A., Singh, U.P. and Ghosh, K. (2019) Manganese (II) Complexes of Tridentate Ligands Having NNN Donors: Structure, DFT Calculations, Superoxide Dismutase, DNA Interaction, Nuclease and Protease Activity Studies. Inorganica Chimica Acta, 486, 261-273.
https://doi.org/10.1016/j.ica.2018.09.081
[10] Sousa, J., Oliveira, D.R., Lomonaco, D., Correia, A.N., Sousa, C.P., Neto, P.L., Paulo, T.F., Mazzetto, S.E., Clemente, C.S. and Mele, G. (2019) Structural, Photophysical and Electrochemical Properties of a Novel Cardanol-Based Salophen Ligand and Its Mn(II) Complex. Journal of Molecular Structure, 1181, 279-286.
https://doi.org/10.1016/j.molstruc.2018.12.111
[11] Gorczyński, A., Pakulski, D., Szymańska, M., Kubicki, M., Bułat, K., Łuczak, T. and Patroniak, V. (2016) Electrochemical Deposition of the New Manganese(II) Schiff-Base Complex on a Gold Template and Its Application for Dopamine Sensing in the Presence of Interfering Biogenic Compounds. Talanta, 149, 347-355.
https://doi.org/10.1016/j.talanta.2015.11.050
[12] Egekenze, R.N., Gultneh, Y. and Butcher, R. (2018) Mn(III) and Mn(II) Complexes of Tridentate Schiff Base Ligands; Synthesis, Characterization, Structure, Electrochemistry and Catalytic Activity. Inorganica Chimica Acta, 478, 232-242.
https://doi.org/10.1016/j.ica.2018.01.027
[13] Abdel-Rahman, L.H., Abu-Dief, A.M., Adam, M.S.S. and Hamdan, S.K. (2016) Some New Nano-Sized Mononuclear Cu(II) Schiff Base Complexes: Design, Characterization, Molecular Modeling and Catalytic Potentials in Benzyl Alcohol Oxidation. Catalysis Letters, 146, 1373-1396.
https://doi.org/10.1007/s10562-016-1755-0
[14] Al-Saeedi, S.I., Abdel-Rahman, L.H., Abu-Dief, A.M., Abdel-Fatah, S.M., Alotaibi, T.M., Alsalme, A.M. and Nafady, A. (2018) Catalytic Oxidation of Benzyl Alcohol Using Nanosized Cu/Ni Schiff-Base Complexes and Their Metal Oxide Nanoparticles, Catalysts, 8, 452.
https://doi.org/10.3390/catal8100452
[15] Zhou, J.-L., Guo, L., Yu, W.-D., Zhang, Z.-H., Wang, Y. and Yan, J. (2019) Impact of Ligand Rotation: Synthesis, Crystal Structures and Third-Order Nonlinear Optical Properties of Mn(II), Cu(II) and Ni(II) Complexes with 5-diethylamino-2-((4-(phenyldiazenyl) phenylimino) methyl) Phenol. Inorganic Chemistry Communications, 99, 189-194.
https://doi.org/10.1016/j.inoche.2018.11.023
[16] Choubey, S., Roy, S., Bhar, K., Ghosh, R., Mitra, P., Lin, C.-H., Ribas, J. and Ghosh, B.K. (2013) Syntheses, Structures, and Magnetic Properties of Terephthalato Bridged Dinuclear Copper(II) and Manganese(II) Complexes with a Tetradentate N-Donor Schiff Base. Polyhedron, 55, 1-9.
https://doi.org/10.1016/j.poly.2013.02.062
[17] Elshafaie, A., Abdel-Rahman, L.H., Abu-Dief, A.M., Hamdan, S.K., Ahmed, A.M. and Ibrahim, E.M.M. (2018) Electric, Thermoelectric and Magnetic Properties of Nickel(II) Imine Nanocomplexes. NANO: Brief Reports and Reviews, 13, Article ID: 1850074.
https://doi.org/10.1142/S1793292018500741
[18] Raju, V., Kumar, R.S., Tharakeswar, Y. and Kumar, S.K.A. (2019) A Multifunctional Schiff-Base as Chromogenic Chemosensor for Mn2+ and Fluorescent Chemosensor for Zn2+ in Semi-Aqueous Environment. Inorganica Chimica Acta, 493, 49-56.
https://doi.org/10.1016/j.ica.2019.04.053
[19] Abu-Dief, A.M., Díaz-Torres, R., Sañudo, E.C., Abdel-Rahman, L.H. and Aliaga-Alcalde, N. (2013) Novel Sandwich Triple-Decker Dinuclear NdIII-(bis-N,N’-p-bromo-salicylideneamine-1,2-diaminobenzene) Complex. Polyhedron, 64, 203-208.
https://doi.org/10.1016/j.poly.2013.04.010
[20] Li, H., Xi, D., Niu, Y., Wang, C., Xu, F., Liang, L. and Xu, P. (2019) Design, Synthesis and Biological Evaluation of Cobalt(II)-Schiff Base Complexes as ATP-Non-competitive MEK1 Inhibitors. Journal of Inorganic Biochemistry, 195, 174-181.
https://doi.org/10.1016/j.jinorgbio.2019.03.022
[21] Abdel-Rahmana, L.H., Abu-Dief, A.M., Aboelez, M.O. and Abdel-Mawgoud, A.A.H. (2017) DNA Interaction, Antimicrobial, Anticancer Activities and Molecular Docking Study of Some New VO(II), Cr(III), Mn(II) and Ni(II) Mononuclear Chelates Encompassing Quaridentate Imine Ligand. Journal of Photochemistry & Photobiology, B: Biology, 170, 271-285.
https://doi.org/10.1016/j.jphotobiol.2017.04.003
[22] Abdel-Rahman, L.H., Abu-Dief, A.M., Shehata, M.R., Atlam, F.M. and Abdel-Maw-goud, A.A.H. (2019) Some New Ag(I), VO(II) and Pd(II) Chelates Incorporating Tridentate Imine Ligand: Design, Synthesis, Structure Elucidation, Density Functional Theory Calculations for DNA Interaction, Antimicrobial and Anticancer Activities and Molecular Docking Studies. Applied Organometallic Chemistry, 33, e4699.
https://doi.org/10.1002/aoc.4699
[23] Pordea, A., Mathis, D. and Ward, T.R. (2009) Incorporation of Biotinylated Manganese-Salen Complexes into Streptavidin: New Artificial Metalloenzymes for Enantioselective Sulfoxidation. Journal of Organometallic Chemistry, 694, 930-936.
https://doi.org/10.1016/j.jorganchem.2008.11.023
[24] Rondot, L., Girgenti, E., Oddon, F., Marchi-Delapierre, C., Jorge-Robin, A. and Ménage, S. (2016) Catalysis without a Headache: Modification of Ibuprofen for the Design of Artificial Metalloenzyme for Sulfide Oxidation. Journal of Molecular Catalysis A: Chemical, 416, 20-28.
https://doi.org/10.1016/j.molcata.2016.02.015
[25] McKee, V. (1993) Macrocyclic Complexes as Models for Nonporphine Metalloproteins. Advances in Inorganic Chemistry, 40, 323-410.
https://doi.org/10.1016/S0898-8838(08)60186-5
[26] Horrocks, W.D. and Burlone, D.A. (1979) Metalloprotein Models, Location of the Magnetic Axes in Low-Symmetry Complexes. Single Crystal Electron Paramagnetic Resonance, Magnetic Susceptibility Anisotropy, and Angular Overlap Ligand Field Calculations on a Complex Containing the Distorted Tetrahedral CoIIN2O2 Coordination Unit, Bis(N-isopropylsalicyclaldiminato)cobalt(II). Inorganica Chimica Acta, 35, 165-175.
https://doi.org/10.1016/S0020-1693(00)93435-2
[27] Bhar, K., Sutradhar, D., Choubey, S., Ghosh, R., Lin, C.-H., Ribas, J. and Ghosh, B.K. (2013) Hexa- and Hepta-Coordinated Manganese(II) Dicyanamide Complexes Containing a Tetradentate N-Donor Schiff Base: Syntheses, Composition Tailored Architectures and Magnetic Properties. Journal of Molecular Structure, 1051, 107-114.
https://doi.org/10.1016/j.molstruc.2013.07.029
[28] Martinez-Bulit, P., Garza-Ortíz, A., Mijangos, E., Barrón-Sosa, L., Sánchez-Bartéz, F., Gracia-Mora, I., Flores-Parra, A., Contreras, R., Reedijk, J. and Barba-Behrens, N. (2015) 2,6-Bis(2,6-diethylphenyliminomethyl)pyridine Coordination Compounds with Cobalt(II), Nickel(II), Copper(II), and Zinc(II): Synthesis, Spectroscopic Characterization, X-Ray Study and in Vitro Cytotoxicity. Journal of Inorganic Biochemistry, 142, 1-7.
https://doi.org/10.1016/j.jinorgbio.2014.09.007
[29] Gökçe, C., Dilek, N. and Gup, R. (2015) Seven Coordinated Cobalt(II) Complexes with 2,6-Diacetylpyridine Bis(4-acylhydrazone) Ligands: Synthesis, Characterization, DNA-Binding and Nuclease Activity. Inorganica Chimica Acta, 432, 213-220.
https://doi.org/10.1016/j.ica.2015.03.040
[30] Baffert, C., Collomb, M.-N., Deronzier, A., Kjærgaard-Knudsen, S., Latour, J.-M., Lund, K.H., McKenzie, C.J., Mortensen, M., Nielsen, L.P. and Thorup, N. (2003) Biologically Relevant Mono- and Di-Nuclear Manganese II/III/IV Complexes of Mononegative Pentadentate Ligands. Dalton Transaction, 9, 1765-1772.
https://doi.org/10.1039/b300823a
[31] Waldo, G.S. and Penner-Hahn, J.E. (1995) Mechanism of Manganese Catalase Peroxide Disproportionation: Determination of Manganese Oxidation States during Turnover. Biochemistry, 34, 1507-1512.
https://doi.org/10.1021/bi00005a006
[32] Zahran, Z.N., Chooback, L., Copeland, D.M., West, A.H. and Richter-Addo, G.B. (2008) Crystal Structures of Manganese- and Cobalt-Substituted Myoglobin in Complex with NO and Nitrite Reveal Unusual Ligand Conformations. Journal of Inorganic Biochemistry, 102, 216-233.
https://doi.org/10.1016/j.jinorgbio.2007.08.002
[33] Türkoğlu, S. and Özer, İ. (1992) Possible Involvement of Manganese in the Catalytic Mechanism of Bovine Liver Arginase. International Journal of Biochemistry, 24, 937-939.
https://doi.org/10.1016/0020-711X(92)90100-F
[34] Shiga, T. and Oshio, H. (2007) Syntheses, Structures and Magnetic Properties of Mixed-Valence Pentanuclear [Mn3IIMn2III] and Hexanuclear [Co4IICo2III] Complexes Derived from 3-Formylsalicylic Acid. Polyhedron, 26, 1881-1884.
https://doi.org/10.1016/j.poly.2006.09.026
[35] Nayak, M., Hazra, S., Lemoine, P., Koner, R., Lucas, C.R. and Mohanta, S. (2008) Self-Assembled [2×1+1×2] Heterotetranuclear CuII3MnII/CuII3CoII and [2×2+1×3] Heptanuclear CuII7 Compounds Derived from N,N’-o-phenylenebis(3-ethoxysali-cylaldimine): Structures and Magnetic Properties. Polyhedron, 27, 1201-1213.
https://doi.org/10.1016/j.poly.2007.12.010
[36] Bar, A.K., Pichon, C. and Sutter, J.-P. (2016) Magnetic Anisotropy in Two- to Eight-Coordinated Transition-Metal Complexes: Recent Developments in Molecular Magnetism. Coordination Chemistry Reviews, 308, 346-380.
https://doi.org/10.1016/j.ccr.2015.06.013
[37] Dieng, M., Diouf, O., Gaye, M., Sall, A.S., Pérez-Lourido, P., Valencia, L., Caneschi, A. and Sorace, L. (2013) Polynuclear Nickel(II) Complexes with Salicylaldimine Derivative Ligands. Inorganica Chimica Acta, 394, 741-746.
https://doi.org/10.1016/j.ica.2012.09.037
[38] Sow, M.M., Diouf, O., Gaye, M., Sall, A.S., Pérez-Lourido, P., Valencia-Matarranz, L., Castro, G., Caneschi, A. and Sorace, L. (2013) Synthesis, Spectral Characterization and X-Ray Crystal Structure of Fe(III) and Co(III) Complexes with an Acyclic Schiff Base Ligand. Inorganica Chimica Acta, 406, 171-175.
https://doi.org/10.1016/j.ica.2013.07.018
[39] Sow, M.M., Diouf, O., Gaye, M., Salam-Sall, A., Castro, G., Pérez-Lourido, P., Valencia, L., Caneschi, A. and Sorace, L. (2013) Sheets of Tetranuclear Ni(II) [2 × 2] Square Grids Structure with Infinite Orthogonal Two-Dimensional Water-Chlorine Chains. Crystal Growth & Design, 13, 4172-4176.
https://doi.org/10.1021/cg400885f
[40] Sheldrick, G.M. (1997) SHELXTL 97. Program for the Refinement of Crystal Structures. University of Göttingen, Göttingen.
[41] Farrugia, L.J. (1997) It ORTEP-3 for Windows—A Version of It ORTEP-III with a Graphical User Interface (GUI). Journal of Applied Crystallography, 30, 565.
https://doi.org/10.1107/S0021889897003117
[42] Singh, A.K., Pandey, O.P. and Sengupta, S.K. (2013) Synthesis, Spectral and Antimicrobial Activity of Zn(II) Complexes with Schiff Bases Derived from 2-Hydrazino-5-[Substituted phenyl]-1,3,4-Thiadiazole and Benzaldehyde/2-Hydroxyacetophe-none/Indoline-2,3-Dione. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy, 113, 393-399.
https://doi.org/10.1016/j.saa.2013.04.045
[43] Abdel-Rahman, L.H., Abu-Dief, A.M., El-Khatib, R.M. and Abdel-Fatah, S.M. (2016) Sonochemical Synthesis, DNA Binding, Antimicrobial Evaluation and in Vitro Anticancer Activity of Three New Nano-Sized Cu(II), Co(II) and Ni(II) Chelates Based on Tridentate NOO Imine Ligands as Precursors for Metal Oxides. Journal of Photochemistry and Photobiology B: Biology, 162, 298-308.
https://doi.org/10.1016/j.jphotobiol.2016.06.052
[44] Abdel-Rahman, L.H., Ismail, N.M., Ismael, M., Abu-Dief, A.M. and Ahmed, E.A. (2017) Synthesis, Characterization, DFT Calculations and Biological Studies of Mn(II), Fe(II), Co(II) and Cd(II) Complexes Based on a Tetradentate ONNO Donor Schiff Base Ligand. Journal of Molecular Structure, 1134, 851-862.
https://doi.org/10.1016/j.molstruc.2017.01.036
[45] Kumar, D.S. and Alexander, V. (1995) Macrocyclic Complexes of Lanthanides in Identical Ligand Frameworks Part 1. Synthesis of Lanthanide(III) and Yttrium(III) Complexes of an 18-Membered Dioxatetraaza Macrocycle. Inorganica Chimica Acta, 238, 63-71.
https://doi.org/10.1016/0020-1693(95)04687-5
[46] Geary, W.J. (1971) The Use of Conductivity Measurements in Organic Solvents for the Characterisation of Coordination Compounds. Coordination Chemistry Reviews, 7, 81-122.
https://doi.org/10.1016/S0010-8545(00)80009-0
[47] Sarr, M., Diop, M., Thiam, E.I., Barry, A.H., Gaye, M. and Retailleau, P. (2018) Crystal Structure of Aquachlorido(nitrato-κ2O,O’)[1-(pyridin-2-yl-κN)-2(pyridin-2-ylmethylidene-κN)hydrazine-κN2]manganase(II). Acta Crystallographica Section E: Crystallographic Communications, 74, 450-453.
https://doi.org/10.1107/S2056989018003493
[48] Schleife, F., Rodenstein, A., Kirmse, R. and Kersting, B. (2011) Seven-Coordinate Mn(II) and Co(II) Complexes of the Pentadentate Ligand 2,6-Diacetyl-4-Carboxy-methyl-Pyridine Bis(benzoylhydrazone): Synthesis, Crystal Structure and Magnetic Properties. Inorganica Chimica Acta, 374, 521-527.
https://doi.org/10.1016/j.ica.2011.02.064
[49] Compton, R.G., Barghout, R., Eklund, J.C. and Fisher, A.C. (1993) Organometallic Photoelectrochemistry: Oxidation of fac-Tricarbonylchloro[bis(diphenylphosphino) methane] Manganese(I). Journal of Physical Chemistry, 97, 1661-1664.
https://doi.org/10.1021/j100110a031

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.