Next Article in Journal
Diamagnetism of Bulk Graphite Revised
Next Article in Special Issue
Synthesis, Structure and Magnetic Study of a Di-Iron Complex Containing N-N Bridges
Previous Article in Journal
Biomolecular EPR Meets NMR at High Magnetic Fields
Previous Article in Special Issue
Synthesis, Crystal Structures, and Magnetic Properties of Lanthanide (III) Amino-Phosphonate Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ni(II) Dimers of NNO Donor Tridentate Reduced Schiff Base Ligands as Alkali Metal Ion Capturing Agents: Syntheses, Crystal Structures and Magnetic Properties

by
Monotosh Mondal
1,2,
Maharudra Chakraborty
1,
Michael G. B. Drew
3 and
Ashutosh Ghosh
1,*
1
Department of Chemistry, University College of Science, University of Calcutta, 92, A.P.C. Road, Kolkata 700 009, India
2
Department of Chemistry, Haldia Government Colleg, Debhog, Purba Medinipur 721657, India
3
School of Chemistry, The University of Reading, P.O. Box 224, Whiteknights, Reading RG6 6AD, UK
*
Author to whom correspondence should be addressed.
Magnetochemistry 2018, 4(4), 51; https://doi.org/10.3390/magnetochemistry4040051
Submission received: 22 October 2018 / Revised: 13 November 2018 / Accepted: 14 November 2018 / Published: 21 November 2018
(This article belongs to the Special Issue A Themed Issue in Honor of Late Professor Samiran Mitra)

Abstract

:
Three trinuclear Ni(II)-Na(I) complexes, [Ni2(L1)2NaCl3(H2O)]·H2O (1), [Ni2(L2)2NaCl3(H2O)] (2), and [Ni2(L3)2NaCl3(OC4H10)] (3) have been synthesized using three different NNO donor tridentate reduced Schiff base ligands, HL1 = 2-[(3-methylamino-propylamino)-methyl]-phenol, HL2 = 2-[(3-methylamino-propylamino)-methyl]-4-chloro-phenol, and HL3 = 2-[(3-methylamino-propylamino)-methyl]-6-methoxy-phenol that had been structurally characterized. Among these complexes, 1 and 2 are isostructural in which dinuclearNi(II) units act as metalloligands to bind Na(I) ions via phenoxido and chlorido bridges. The Na(I) atom is five-coordinated, and the Ni(II) atom possesses hexacordinated distorted octahedral geometry. In contrast, in complex 3, two -OMe groups from the dinuclear Ni(II) unit also coordinate to Na(I) to make its geometry heptacordinated pentagonal bipyramidal. The magnetic measurements of complexes 13 indicate ferromagnetic interactions between dimeric Ni(II) units with J = 3.97 cm−1, 4.66 cm−1, and 5.50 cm−1 for 13, respectively, as is expected from their low phenoxido bridging angles (89.32°, 89.39°, and 87.32° for 13, respectively). The J values have been calculated by broken symmetry DFT method and found to be in good agreement with the experimental values.

Graphical Abstract

1. Introduction

During the past few decades, synthesis and characterizations of polynuclear transition metal complexes and their magnetic and electronic properties have received much attention from chemists due to their potential applications in magnetic ordering and catalytic and biological mimicking [1,2,3,4,5,6,7,8,9]. A common strategy for the synthesis of such polynuclear complexes is to bind the metal centers by a single atom bridge. The Schiff bases derived from diamines and salicylaldehyde derivatives are very useful for this purpose, as the phenoxido oxygen atom can connect two or three metal centers very efficiently. The literature shows that exploiting this property, various N2O2 or N2O4 donor Schiff base ligands have been used widely for the synthesis of 3d-3d′, 3d-4f/5f, and 3d-ns heterometallic complexes [10,11,12,13,14,15,16]. On the contrary, N2O donor tridentate Schiff base ligands are known to form mostly homonuclear complexes [17,18,19]. As, for example, with Ni(II), these N2O donor ligands form diphenoxido bridged dinuclear complexes [20,21,22]. These complexes attracted significant attention due to their importance in the study of magnetic coupling between the metal centers through the phenoxido bridges [23,24,25]. It is now well understood that the ferro-to antiferromagnetic crossover angle for diphenoxido bridged Ni(II) is 93–94° and most of these reported complexes are antiferromagnetically coupled, because they are stable at bond angles higher than the crossover angle [26,27,28,29]. It is also known that the ferromagnetically coupled complexes with the lower Ni-O-Ni bridging angle can be obtained by introducing an additional bridging atom between the Ni(II) centers in the complexes of NNO donor ligands [23,30,31,32,33]. However, to date, unlike the mononuclear complexes of N2O2 or N2O4 donor Schiff base ligands, the dinuclear complexes of these NNO donor ligands are not known to act as metalloligands when coordinating with a second metal ion.
Herein, we report the synthesis of three new heterometallic Ni(II)-Na(I) complexes, [Ni2(L1)2NaCl3(H2O)]·H2O (1), [Ni2(L2)2NaCl3(H2O)] (2), and [Ni2(L3)2NaCl3(OC4H10)] (3), by using three different NNO donor tridentate reduced Schiff bases ligands (HL1 = 2-[(3-methylamino-propylamino)-methyl]-phenol, HL2 = 2-[(3-methylamino-propylamino)-methyl]-4-chloro-phenol, and HL3 = 2-[(3-methylamino-propylamino)-methyl]-6-methoxy-phenol). The complexes are characterized by single crystal X-ray crystallography, electronic spectra, IR spectra and elemental analyses. In 1 and 2, the sodium ion is attached to the dinuclear Ni(II) complex with the help of phenoxido and chlorido bridges, whereas in 3, two -OMe groups form additional bridges between the dinuclear unit and Na(I). The magnetic properties of the complexes have been studied both experimentally and using broken symmetry DFT method. The couplings are found to be ferromagnetic, and consistent with the low phenoxido bridging angles between the NI(II) centers. The experimentally obtained coupling parameters match well with the theoretically calculated values.

2. Results and Discussion

2.1. Syntheses of the Complexes

Three reduced tridentate Schiff base ligands, HL1, HL2, and HL3, were synthesized by condensation of N-methyl-1,3-propanediamine with salicylaldehyde, 5-chlorosalicylaldehyde, and 2-hydroxy-3-methoxybenzaldehyde in 1:1 molar ratios and subsequent reduction with sodium borohydride in methanol. We obtained complexes (13) accidentally in low yield when the methanolic solutions of these as-synthesized ligands were reacted directly with NiCl2·6H2O after reduction. The sodium ions that were present in the solution (from sodium borohydride) were captured by the dinuclear complexes. Later, in order to improve the yield, we isolated the ligands and reacted them with nickel(II) chloride in 1:1 molar ratios in the presence of sodium chloride in methanol-water solvent mixture (vide experimental section, Scheme 1).

2.2. IR and UV–Vis Spectra of the Complexes

Complexes 13 show a moderately strong, sharp peak at 3266 cm−1, 3270 cm−1, and 3282 cm−1, respectively, due to N-H stretching vibration, indicating that the imine group(C=N) of the Schiff base ligands is reduced. Absence of any sharp peak at 1620–1650 cm−1 for stretching vibration of imine group(C=N) also indicates that Schiff base is reduced (Figures S1–S3). The electronic spectra (Figure S4) of complexes 1, 2, and 3 are recorded in methanolic solution. The spectra show bands at 925.2 and 644.7 nm for 1, 933.9 and 641.0 nm for 2, and 973.7 and 638.7 nm for 3, which can be assigned to the spin-allowed transitions 3T1g(F)←3A2g and 3T2g3A2g, respectively.

2.3. Description of the Crystal Structures

The structures of complexes 1 and 2, both with the formula [Ni2(L1−2)2NaCl3(OH2)], as shown in Figure 1 (A for 1 and B for 2), contain crystallographic C2 symmetry and have very similar structures. Each nickel atom is six-coordinated with a distorted octahedral environment. In the equatorial plane are three donor atoms, O(11), N(19), and N(23), of one ligand together with the oxygen atom O(11)$1 ($1=1−x, y, 3/2−z) of a second ligand. In axial positions there are two chlorine atoms, namely, Cl(2) on the two-fold axis, which bridges the two nickel atoms; and Cl(1), which bridges to a sodium ion, also on a two-fold axis. The sodium ion is additionally bonded to two O(11) atoms and a water molecule O(1) on the two fold axis, and thus the geometry around sodium is five-coordinated distorted trigonal bipyramidal.
The dimensions in the nickel and sodium coordination spheres in 1 and 2 are very similar and are as expected. The Ni-O(11) bonds of 2.049(2), 2.053(2) Å are shorter than the Ni-O(11)$1 bonds of 2.164(2), 2.164(2) Å. The Ni-N bonds fall in the range of 2.079(2)–2.085(3) Å. Bonds to Cl(1) are slightly longer at 2.4782(10), 2.4626(8) Å than those to Cl(2) of 2.4300(10), 2.4497(7) Å. The sodium ions are bonded to O(11) at 2.388(3), 2.400(2) Å, and Cl(2) at 2.7532(11), 2.7525(8) Å, and O(1) at 2.293(6), 2.296(6) Å in 1 and 2, respectively. In both structures, the water oxygen atom O(1) had high thermal parameters and the hydrogen atoms could not be located. In 1, the atom was refined with reduced occupancy, as was an adjacent solvent oxygen atom. For both structures, there is a hydrogen bond from N(19) to Cl(2) ($2 = ½−x, ½−y, 1−z) with dimensions H…Cl 2.61 2.56 Å, N…Cl 3.331(3), 3.332(2) Å, and N-H…Cl 130, 136° for 1 and 2, respectively (Table 1). The Ni…Ni and Ni…Na distances are 2.963(1), 2.966(1) Å and 3.168(2), and 3.186(2) Å in 1 and 2, respectively. In contrast, the structure of 3 that is shown in Figure 2 is different in that the ligand contains an -OMe group that is bonded to the sodium ion. However, the molecule is also located on a two-fold axis, and the arrangement around the two nickel atoms is similar to that found in 1 and 2. The only significant difference in the bond lengths around nickel is to be found in Ni-O(11)$1 ($1 = 1−x, y, 3/2−z), which is considerably longer at 2.3178(11) Å. However the sodium ion, also on a two-fold axis, is now seven-coordinated, with a distorted pentagonal bipyramidal structure being bonded to 2*O(11), 2*O(131), and O(1) from the (C2H5)2O ligand in the equatorial plane at bond lengths 2.4164(15), 2.4265(13), and 2.775(4) Å, respectively. The bond to O(1) is therefore particularly weak. The five atoms in the equatorial plane show a r.m.s. deviation of 0.219 Å with the sodium ion on the two-fold axis perforce in the plane. The two chlorine atoms in axial positions form longer bonds at 2.8930(5) Å than are found in 1 and 2. As in 1 and 2, only one of the N-H forms a hydrogen bond, in this case N(19)-H…Cl(2)$2 ($2=−x+1/2, −y+3/2, −z+1) with H…Cl 2.62 N…Cl 3.3454(14) Å N-H…Cl 131o. The Ni…Ni and Ni…Na distances are slightly longer than in 1 and 2 at 3.011(1), 3.280(1) Å. Selected bond lengths and bond angles are summarized in Table S1.

2.4. Magnetic Properties

The variation of the product of molar magnetic susceptibility (χM) and temperature (T) with respect to temperature (T) for NiII dimers 1, 2, and 3 is presented in Figure 3, Figure 4 and Figure 5. It can be seen that the experimental χMT values for three NiII dimers at room temperature (~300 K) lie around 2.24–2.73 cm3 K mol−1, which in good agreement with the expected value for two non-interacting NiII, S = 1 centers (the theoretical spin-only value is 2.0 cm3 K mol−1, g = 2). For all the complexes, variation of χMT values shows similar type of behavior on lowering the temperature: the χMT values gradually increase upon lowering of temperatures and attain a maximum at ca. 10–12 K. On further decreasing the temperature, a steep fall of the values is observed. This type of behavior is typical of a system that shows dominant intramolecular ferromagnetic exchange coupling, and the steep drop in χMT values at lower temperatures can be attributed to zero-field splitting (ZFS) of the ground state (S = 2) and/or intermolecular interactions between the dimers.
The obtained magnetic data of the three compounds were analyzed by using the HDVV Hamiltonian H = −2JS1S2. We fitted the magnetic data of complexes 1, 2, and 3 for a simple S = 1 dimer model following the equation (1), which is derived from the HDVV Hamiltonian [4], in which x = 2 J / kT .
χ M T   =   N g 2 β 2 T k ( T θ ) × 2 e x + 10 e 3 x 1 + 3 e x + 5 e 3 x  
The Weiss constant was taken (θ) as an additional parameter to include the intermolecular interaction term (ZJ′) as well as the anisotropy term (D). The parameters N, β, and k in Equation (1) have their usual meanings, and J is the coupling parameter.
The χMT versus T curves were fitted by least-squares method minimizing the function R = Σ [(χMT)exp.− (χMT)calcd.]2/Σ(χMT)exp2. The best-fitting parameters are found to be J = 3.97(8) cm−1, g = 2.08(2), and θ = −1.16 K with R = 1.2 × 10−5 (for complex 1); J = 4.66(2) cm−1, g = 2.14(9), and θ = −1.49 K with R = 1.52 × 10−5 (for complex 2); and J = 5.50(1) cm−1, g = 2.30(2), and θ = −1.29 K with R = 2.3 × 10−6 (for complex 3) (Table 2). Here, it is worth mentioning that consideration of only the anisotropy term (D) in the spin Hamiltonian yielded poor data fitting at low temperature.
Thus, besides intra-dimer coupling, ZFS and inter-dimer coupling are also present in the systems; however, their correct evaluation is not possible due to their close relationship [34,35].
Isothermal magnetization measurements studies (Figures S5–S7) suggest that the ferromagnetic coupling of all complexes leads to S = 2 ground spin states. However, the magnetization values at 5T are smaller than the expected values of ca. 4µB, because there are inter-dimer antiferromagnetic interactions and/or zero-field splitting [26].

2.5. Theoretical Magnetic Calculation

In this present work, we have calculated the magnetic exchange parameters (J) by broken symmetry DFT calculations using B3LYP functional as described in experimental section. The Hamiltonian for the systems was taken as H = −2JS1S2, in which J is the coupling constants for the pair of NiII centers having S1 and S2 spin vectors for all the dimers.
The magnetic coupling constant (J) between a pair of metallic centers can be estimated by the following relationship [26]:
E BS -   E HS S max ( S max   +   1 )   =   J  
For complex 1, the calculated coupling constant (J) by broken symmetry DFT method is 4.25 cm−1 showing ferromagnetic coupling interaction in agreement with the experimental results. However, the DFT method slightly overestimates the J value. For complex 2, the estimated value is 3.56 cm−1, which also indicates the presence of ferromagnetic interaction between the NiII centers, but the theoretical result slightly underestimates the coupling constant. For complex 3, the calculated value of coupling constant from broken symmetry DFT method is found to be 5.50 cm−1, which is in good agreement with the experimental value.
In order to further investigate the mechanism of exchange pathway, the spin density distribution has been analyzed. Molecular orbital theory suggests that the spin delocalization transfers the spin density from the magnetic metal centers to their corresponding adjacent ligand atoms. The Mulliken spin populations of complexes 1, 2, and 3 in their broken symmetry states are represented in Table 3, Table 4 and Table 5, respectively (spin density orbital plots are shown in Figures S8–S10).
The positive sign of spin density represents the α spin states, and the negative sign represents the β spin states. From the tables, it can be seen that two NiII centers have same amount of spin populations but the sign is opposite. Moreover, each NiII center has highest value of spin density among all the atoms present in a compound. Thus, it is obvious that NiII centers are indeed the magnetic centers. It can be also seen from the tables that about ~17% of total spin density is distributed among the other atoms present in the complexes.
If the crystallographic structure of each compound is examined, one could find two oxygen atoms and one chlorine atom bridging the two magnetic metal centers along with one Cl-Na-Cl moiety. However, we can see from the spin density table that there is nearly zero delocalization on Na and bridging Cl atom for all the complexes, and only the significant amount of spin density is showing on the bridging oxygen atoms from phenoxido bridge. Negligible overall spin population on the bridging chloride atoms is the result of partial α and β spin density overlap due to the spin polarization in broken symmetry state (Figures S8–S10) [36,37,38]. Thus, it may be concluded that the ferromagnetic exchange interactions between two NiII centers are dominated by phenoxido bridges for all the complexes [39].

2.6. Comparison of Structural and Magnetic Parameters

The dimeric units of all three NiII complexes have a C2 axis of symmetry, and the two NiII centers are connected through a pair of phenoxido bridges and a chlorido bridge along with a Cl-Na-Cl bridging unit in all the complexes (Figure 6). The core structures are similar for all the complexes, but they have slightly different Ni-O-Ni, Ni-Cl-Ni bond angles and Ni⋯Ni distances. From the structures of these complexes, it is expected that the overall magnetic exchange interaction between the NiII centers are mediated by two phenoxido and one chlorido bridges. However, from the spin density calculation, no significant overall spin density is observed on the chloride bridge. Thus, only the phenoxido bridges play crucial role in the coupling parameters (J).
The sign and magnitude of the magnetic coupling constants of binuclear oxido/phenoxido bridging Ni(II) complexes have been investigated both experimentally and theoretically and correlated with the structural parameters by various groups [24,40,41,42,43]. It is now understood that ferromagnetic coupling is observed when the bridging angle is around 90° as a result of the orthogonality of magnetic orbitals. As the Ni-O-Ni angle increases, the magnitude of coupling constant decreases and tends to become antiferromagnetic at around 95°–96°. For the present complexes, the two phenoxido bridging angles are close to 90°, (89.32(8)°, 89.39(6)°, and 87.32(4)° for 13, respectively). Hence, the coupling is ferromagnetic. It is to be noted here that several ferromagnetically coupled diphenoxido bridged dinuclear Ni(II) were synthesized by introducing an additional water bridge between the Ni(II) centers. In the present complexes, this additional bridging atom is chloride, and it has a similar effect on magnetic coupling as the water bridge, i.e., it stabilizes the molecule below the crossover angle, and consequently the coupling becomes ferromagnetic.

3. Experimental Section

3.1. Starting Materials

Salicylaldehyde, 5-chlorosalicylaldehyde, 2-hydroxy-3-methoxybenzaldehyde and N-methyl-1,3-propanediamine, and sodium borohydride were purchased from Spectrochem, India and were of reagent grade. They were used without further purification. The other reagents and solvents were of commercially available reagent quality, unless otherwise stated.

3.2. Synthesis of the Ligands 2-[(3-Methylamino-propylamino)-methyl]-phenol (HL1), 2-[(3-Methylamino-propylamino)-methyl]-4-chloro-phenol (HL2), 2-[(3-Methylamino-propylamino)-methyl]-6-methoxy-phenol (HL3)

The ligand, HL1, was synthesized [44] by refluxing a solution of salicylaldehyde (0.52 mL, 5 mmoL) and N-methyl-1,3-propanediamine (0.52 mL, 5 mmoL) in methanol (30 mL) for 1 h. The methanolic solution was subsequently cooled to 0°C and solid sodium borohydride (210 mg, 6 mmoL) was added slowly with constant stirring. After completion of the addition, the resulting reaction mixture was acidified with concentrated HCl (5 mL) and then evaporated to dryness. The reduced Schiff-base ligand (HL1) was extracted from the solid residue with methanol. This methanolic solution was used for further reaction. HL2 and HL3 were synthesized, maintaining a similar procedure by using 5-chlorosalicylaldehyde (782.85 mg, 5 mmoL) and 2-hydroxy-3-methoxybenzaldehyde (760.75 mg, 5 mmoL), respectively, instead of salicylaldehyde.

3.3. Synthesis of the Complexes [Ni2(L1)2NaCl3(H2O)]·H2O (1), [Ni2(L2)2NaCl3(H2O)] (2), and [Ni2(L3)2NaCl3(OC4H10)] (3)

NiCl2·6H2O (647.95 mg, 5 mmoL), dissolved in 20 mL methanol, was added to methanolic solution of previously prepared HL1 as described above with constant stirring. An aqueous solution of (2 mL) of NaCl (146.1 mg, 2.5 mmoL) was slowly added with constant stirring after 15 min. The resulting green solution was kept in air for overnight. On slow evaporation of the resulting filtrate, green single crystals of complex 1 were obtained for X-ray diffraction. Complexes 2 and 3 were synthesized following the same procedure, using HL2 and HL3, respectively. Green single crystals of 2 were obtained by the same procedure as stated above. For complex 3, crystals suitable for X-ray diffraction were obtained by layer diffusion of diethylether to the acetonitrile solution of 3.
Complex (1): Yield: 1.139 g; (70%). Anal. Calcd. For C22H36Cl3N4NaNi2O3 (651.31): C, 40.57; H, 5.57; N, 8.60. Found: C, 40.72; H, 5.68; N, 8.75. IR (KBr pellet, cm−1): ν(N-H), 3266 cm−1, ν(C-N),1592 cm−1.
Complex (2): Yield: 1.35 g; (75%). Anal. Calcd. For C22H34Cl5N4NaNi2O3 (720.19): C, 36.69; H, 4.76; N, 7.78 Found: C, 36.81; H, 4.90; N, 7.94. IR (KBr pellet, cm−1): ν(N-H), 3270 cm−1, ν(C-N),1591 cm−1.
Complex (3): Yield: 1.381 g; (72%). Anal. Calcd. For C28H48Cl3N4NaNi2O5 (767.46): C, 43.82; H, 6.30; N, 7.30. Found: C, 43.96; H, 6.45; N, 7.51. IR (KBr pellet, cm−1): ν(N-H), 3282 cm−1, ν(C-N),1569 cm−1.

3.4. Physical Measurements

Elemental analyses (C, H, and N) were performed using a Perkin-Elmer 2400 series II elemental analyzer. IR spectra in KBr pellets (4500–500 cm−1) were recorded using a Perkin-Elmer RXI FT-IR spectrophotometer. Electronic spectra (1500–250 nm) were recorded in a Hitachi U-3501 spectro-photometer. The magnetic measurements of all the three NiII dimers (1, 2, and 3) were investigated with a Quantum Design superconducting quantum interference device vibrating sample magnetometer (SQUID-VSM). Powdered polycrystalline samples were used for all measurements under a DC magnetic field. The molar paramagnetic susceptibilities (χM) were measured at a constant magnetic field at 0.075 T under a decreasing temperature range of 300 to 2.5 K. Isothermal magnetizations measurements were performed at 2 K up to 5 Tesla for all the complexes. The measured susceptibilities were corrected according to the literature values of Pascal’s table due to the diamagnetic contributions [45]. For complex 3, all the magnetic calculations were carried out by deducting the molecular weight of readily volatile diethyl ether molecule from the actual molecular weight of the sample.

3.5. X-ray Crystallographic Data Collection and Refinement

Collected diffractable single crystals of complexes 1, 2, and 3 were mounted on a Bruker-AXS SMART APEX II diffractometer equipped with a graphite monochromator and Mo-Kα (λ = 0.71073 Å) radiation. The crystals were positioned at 60 mm from the CCD. 360 frames were measured with a counting time of 5 s. The structures were solved using direct methods with the Shelxs97 program [46]. The non-hydrogen atoms were refined with anisotropic thermal parameters. The hydrogen atoms bonded to carbon were included in geometric positions and given thermal parameters equivalent to 1.2 times those of the atom to which they were attached. The structures were refined on F2 using Shelxl16/6 on F2 [47]. In 1, there were two oxygen atoms refined with reduced occupancy but the attached hydrogen atoms could not be located. For 2, the squeeze option in Platon was used [48]. Details of the crystallographic data are summarized in Table 6. CCDC-1874587 (1), CCDC-1874588 (2), and CCDC-1874589 (3) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

3.6. Computational Methodology

The NiII⋯NiII coupling constants (J) of these complexes were calculated theoretically by broken symmetry DFT as proposed by Ruiz et al. [49,50,51]. All the coordinates of atoms were obtained from experimental X-ray structures and used in calculations without further optimization of the structures. The hybrid B3LYP functional [52,53,54] along with Ahlrichs type triple zeta with polarization function def2-tzvp basis set [55] has been employed in all calculations as implemented in the ORCA package (version 3.0.3) [56]. The zero’th-order regular approximation (ZORA) has been also incorporated to describe scalar relativistic effect [57]. To speed up the calculations with desired accuracy, RIJCOSX approximation with auxiliary def2-TZVP/J coulomb fitting basis set and tight SCF convergence criteria (Grid 4) have also been incorporated [58].

4. Conclusions

In this article, three dinuclear Ni(II) complexes (1, 2, and 3) have been synthesized using the reduced form of three different NNO donor Schiff base ligands along with chloride coligand. In all the complexes, the two Ni(II) centers are connected through two phenoxido bridges and one chlorido bridge. The additional chloride bridge reduced the Ni-O-Ni bond angles below the crossover angle, and hence ferromagnetic coupling between the Ni(II) centers was observed. These complexes are the first examples of diphenoxido bridged dinuclear Ni(II) compounds in which an additional chlorido bridge is formed between the Ni(II) centre. Moreover, these complexes provide unprecedented examples in which the dimeric Ni(II) complexes act as metalloligands to capture an alkali metal ion with the help of bridging phenoxido groups. The magnetic properties of the complexes have been rationalized by theoretical calculations. The theoretically calculated coupling parameters obtained via broken symmetry DFT methods match well with those found experimentally. Spin density calculations show that the additional chlorido bridge connecting the Ni(II) centers does not contribute to the coupling parameters and the oxo-bridges are solely responsible for the coupling by carrying the spin densities between the magnetic metal centers.

Supplementary Materials

The following are available online at https://www.mdpi.com/2312-7481/4/4/51/s1. Figures S1–S3: IR spectrums of complexes 13, Figure S4: UV-Vis spectrums of complexes 13, Figures S5–S7: Magnetic field dependence of molar magnetizations for complexes 13 at 2 K. Figures S8–S10: SD in the BS state of complexes 13, Table S1: Molecular dimensions (distances, Å, angles, °) in 1, 2, and 3.

Author Contributions

M.M. participated in the preparations, characterizations, X-ray structural analysis and magnetic data collections. M.C. analyzed, interpreted the magnetic data. M.G.B.D involved in structural analysis. A.G. designed the study and wrote the manuscript.

Funding

This research received no external funding.

Acknowledgments

M.C. acknowledges the financial support provided by University Grants Commission, India (F.4-2/2006 (BSR)/CH/15-16/0160, Dated 22April 2016) through the D.S. Kothari Post-Doctoral Fellowship (DSKPDF). A.G. thanks the Sophisticated Analytical Instrument Facility (SAIF), IIEST, Shibpur, Howrah for providing the single crystal X-ray diffractometer facility. We also thank the CRNN, University of Calcutta, Kolkata, India, for magnetic measurement. A.G. also thanks University Grants Commission (UGC), New Delhi for funding the CAS-V, Department of Chemistry, University of Calcutta.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Deuss, P.J.; Heeten, R.; Laan, W.; Kamer, P.C. Bioinspired catalyst design and artificial metalloenzymes. Chem. Eur. J. 2011, 17, 4680–4698. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, W.; Thorp, H.H. Bond valence sum analysis of metal-ligand bond lengths in metalloenzymes and model complexes. 2. Refined distances and other enzymes. Inorg. Chem. 1993, 32, 4102–4105. [Google Scholar] [CrossRef]
  3. Suh, J. Model studies of metalloenzymes involving metal ions as Lewis acid catalysts. Acc. Chem. Res. 1992, 25, 273–279. [Google Scholar] [CrossRef]
  4. Kahn, O. Molecular Magnetism; VCH Publishers, Inc.: New York, NY, USA, 1993. [Google Scholar]
  5. Barrios, A.M.; Lippard, S.J. Interaction of urea with a hydroxide-bridged dinuclear nickel center: An alternative model for the mechanism of urease. J. Am. Chem. Soc. 2000, 122, 9172–9177. [Google Scholar] [CrossRef]
  6. Lanznaster, M.; Neves, A.; Bortoluzzi, A.J.; Szpoganicz, B.; Schwingel, E. New FeIIIZnII Complex Containing a Single Terminal Fe-O(phenolate) Bond as a Structural and Functional Model for the Active Site of Red Kidney Bean Purple Acid Phosphatase. Inorg. Chem. 2002, 41, 5641–5643. [Google Scholar] [CrossRef] [PubMed]
  7. Borovik, A.S.; Papaefthymiou, V.; Taylor, L.F.; Anderson, O.P.; Que, L., Jr. Models for iron-oxo proteins. Structures and properties of FeIIFeIII, ZnIIFeIII, and FeIIGaIII complexes with (.mu.-phenoxo) bis (.mu.-carboxylato) dimetal cores. J. Am. Chem. Soc. 1989, 111, 6183–6195. [Google Scholar] [CrossRef]
  8. Belle, C.; Gautier-Luneau, I.; Karmazin, L.; Pierre, J.L.; Albedyhl, S.; Krebs, B.; Bonin, M. Regio-Directed Synthesis of a ZnIIFeIII Complex from an Unsymmetrical Ligand and Its Relevance to Purple Acid Phosphatases. Eur. J. Inorg. Chem. 2002, 2002, 3087–3090. [Google Scholar] [CrossRef]
  9. Mayans, J.; Font-Bardia, M.; Di Bari, L.; Arrico, L.; Zinna, F.; Pescitelli, G.; Escuer, A. From Mesocates to Helicates: Structural, Magnetic and Chiro-Optical Studies on Nickel (II) Supramolecular Assemblies Derived from Tetradentate Schiff Bases. Chem. Eur. J. 2018, 24, 7653–7663. [Google Scholar] [CrossRef] [PubMed]
  10. Gheorghe, R.; Cucos, P.; Andruh, M.; Costes, J.P.; Donnadieu, B.; Shova, S. Oligonuclear 3d–4f Complexes as Tectons in Designing Supramolecular Solid-State Architectures: Impact of the Nature of Linkers on the Structural Diversity. Chem. Eur. J. 2006, 12, 187–203. [Google Scholar] [CrossRef] [PubMed]
  11. Bhowmik, P.; Harms, K.; Chattopadhyay, S. Formation of polynuclear copper (II)–sodium (I) heterometallic complexes derived from salen-type Schiff bases. Polyhedron 2013, 49, 113–120. [Google Scholar] [CrossRef]
  12. Biswas, S.; Naiya, S.; Drew, M.G.B.; Estarellas, C.; Frontera, A.; Ghosh, A. Trinuclear and tetranuclear adduct formation between sodium perchlorate and copper(II) complexes of salicylaldimine type ligands: Structural characterization and theoretical investigation. Inorg. Chim. Acta 2011, 366, 219–226. [Google Scholar] [CrossRef]
  13. Sakamoto, S.; Fujinami, T.; Nishi, K.; Matsumoto, N.; Mochida, N.; Ishida, T.; Sunatsuki, Y.; Re, N. Carbonato-Bridged NiII2LnIII2 (LnIII = GdIII, TbIII, DyIII) Complexes Generated by Atmospheric CO2 Fixation and Their Single-Molecule-Magnet Behavior: [(μ4-CO3)2{NiII (3-MeOsaltn)(MeOH or H2O)LnIII(NO3)}2] solvent [3-MeOsaltn = N,N′-Bis (3-methoxy-2-oxybenzylidene)-1, 3-propanediaminato]. Inorg. Chem. 2013, 52, 7218–7229. [Google Scholar] [PubMed]
  14. Ghosh, S.; Biswas, S.; Bauza, A.; Barceló-Oliver, M.; Frontera, A.; Ghosh, A. Use of metalloligands [CuL](H2L = salen type di-Schiff bases) in the formation of heterobimetallic copper (II)-uranyl complexes: Photophysical investigations, structural variations, and theoretical calculations. Inorg. Chem. 2013, 52, 7508–7523. [Google Scholar] [CrossRef] [PubMed]
  15. Thurston, J.H.; Tang, C.G.Z.; Trahan, D.W.; Whitmire, K.H. Toward Rational Control of Metal Stoichiometry in Heterobimetallic Coordination Complexes: Synthesis and Characterization of Pb(Hsal)2 (Cu(salen*))2, [Pb(NO3)(Cu(salen*)) 2](NO3),Pb(OAc)2(Cu(salen*)), and [Pb(OAc)(Ni(salen*))2](OAc). Inorg. Chem. 2004, 43, 2708–2713. [Google Scholar] [CrossRef] [PubMed]
  16. Kilic, A.; Tas, E.; Deveci, B.; Yilmaz, I. Synthesis, electrochemical and in situ spectroelectrochemical studies of new transition metal complexes with two new Schiff-bases containing N2O2/N2O4 donor groups. Polyhedron 2007, 26, 4009–4018. [Google Scholar] [CrossRef]
  17. Biswas, R.; Ida, Y.; Baker, M.L.; Biswas, S.; Kar, P.; Nojiri, H.; Ishida, T.; Ghosh, A. A New Family of Trinuclear Nickel (II) Complexes as Single-Molecule Magnets. Chem. Eur. J. 2013, 19, 3943–3953. [Google Scholar] [CrossRef] [PubMed]
  18. Hung, W.C.; Lin, C.C. Preparation, characterization, and catalytic studies of magnesium complexes supported by NNO-tridentate Schiff-base ligands. Inorg. Chem. 2008, 48, 728–734. [Google Scholar] [CrossRef] [PubMed]
  19. Li, X.; Lah, M.S.; Pecoraro, V.L. Vanadium complexes of the tridentate Schiff base ligand N-salicylidene-N′-(2-hydroxyethyl) ethylenediamine: Acid-base and redox conversion between vanadium (IV) and vanadium (V) iminophenolates. Inorg. Chem. 1988, 27, 4657–4664. [Google Scholar] [CrossRef]
  20. Mondal, M.; Guha, P.M.; Giri, S.; Ghosh, A. Deactivation of catecholase-like activity of a dinuclear Ni (II) complex by incorporation of an additional Ni (II). J. Mol. Catal. A Chem. 2016, 424, 54–64. [Google Scholar] [CrossRef]
  21. Mukherjee, P.; Drew, M.G.B.; Gómez-García, C.J.; Ghosh, A. (Ni2),(Ni3), and (Ni2+ Ni3): A Unique Example of Isolated and Cocrystallized Ni2 and Ni3 Complexes. Inorg. Chem. 2009, 48, 4817–4827. [Google Scholar] [CrossRef] [PubMed]
  22. Adhikary, J.; Chakraborty, P.; Das, S.; Chattopadhyay, T.; Bauza, A.; Chattopadhyay, S.K.; Ghosh, B.; Mautner, F.A.; Frontera, A.; Das, D. A Combined Experimental and Theoretical Investigation on the Role of Halide Ligands on the Catecholase-like Activity of Mononuclear Nickel (II) Complexes with a Phenol-Based Tridentate Ligand. Inorg. Chem. 2013, 52, 13442–13452. [Google Scholar] [CrossRef] [PubMed]
  23. Mondal, M.; Giri, S.; Guha, P.M.; Ghosh, A. Dependence of magnetic coupling on ligands at the axial positions of NiII in phenoxido bridged dimers: Experimental observations and DFT studies. Dalton Trans. 2017, 46, 697–708. [Google Scholar] [CrossRef] [PubMed]
  24. Nanda, K.K.; Thompson, L.K.; Bridson, J.N.; Nag, K. Linear dependence of spin exchange coupling constant on bridge angle in phenoxy-bridged dinickel (II) complex. J. Chem. Soc. Chem. Commun. 1994, 11, 1337–1338. [Google Scholar] [CrossRef]
  25. Burkhardt, A.; Spielberg, E.T.; Simon, S.; Görls, H.; Buchholz, A.; Plass, W. Hydrogen Bonds as Structural Directive towards Unusual Polynuclear Complexes: Synthesis, Structure, and Magnetic Properties of Copper (II) and Nickel (II) Complexes with a 2-Aminoglucose Ligand. Chem. Eur. J. 2009, 15, 1261–1271. [Google Scholar] [CrossRef] [PubMed]
  26. Biswas, R.; Giri, S.; Saha, S.K.; Ghosh, A. One Ferromagnetic and Two Antiferromagnetic Dinuclear Nickel (II) Complexes Derived from a Tridentate N, N, O-Donor Schiff Base Ligand: A Density Functional Study of Magnetic Coupling. Eur. J. Inorg. Chem. 2012, 2012, 2916–2927. [Google Scholar] [CrossRef]
  27. Banerjee, S.; Drew, M.G.B.; Lu, C.Z.; Tercero, J.; Diaz, C.; Ghosh, A. Dinuclear Complexes of MII Thiocyanate (M= Ni and Cu) Containing a Tridentate Schiff-Base Ligand: Synthesis, Structural Diversity and Magnetic Properties. Eur. J. Inorg. Chem. 2005, 2005, 2376–2383. [Google Scholar] [CrossRef]
  28. Rodríguez, L.; Labisbal, E.; Sousa-Pedrares, A.; García-Vázquez, J.A.; Romero, J.; Durán, M.L.; Real, J.A.; Sousa, A. Coordination chemistry of amine bis (phenolate) cobalt (II), nickel (II), and copper (II) complexes. Inorg. Chem. 2006, 45, 7903–7914. [Google Scholar] [CrossRef] [PubMed]
  29. Luo, W.; Wang, X.T.; Cheng, G.Z.; Gao, S.; Ji, Z.P. Synthesis, structural characterization, and magnetism of a butterfly-shaped hexanuclear Ni (II) complex. Inorg. Chem. Commun. 2008, 11, 769–771. [Google Scholar] [CrossRef]
  30. Biswas, A.; Drew, M.G.B.; Gómez-García, C.J.; Ghosh, A. Formation of a dinuclear and a trinuclear Ni (II) complex on slight variation of experimental conditions: Structural analysis and magnetic properties. Polyhedron 2017, 121, 80–87. [Google Scholar] [CrossRef]
  31. Ruiz, E.; Cano, J.; Alvarez, S.; Alemany, P. Magnetic coupling in end-on azido-bridged transition metal complexes: A density functional study. J. Am. Chem. Soc. 1998, 120, 11122–11129. [Google Scholar] [CrossRef]
  32. Biswas, R.; Kar, P.; Song, Y.; Ghosh, A. The importance of an additional water bridge in making the exchange coupling of bis (μ-phenoxo) dinickel (II) complexes ferromagnetic. Dalton Trans. 2011, 40, 5324–5331. [Google Scholar] [CrossRef] [PubMed]
  33. Biswas, R.; Diaz, C.; Ghosh, A. Three nickel (II) complexes derived from a tridentate NNO donor Schiff base ligand: Syntheses, crystal structures and magnetic properties. Polyhedron 2013, 56, 172–179. [Google Scholar] [CrossRef]
  34. Mukherjee, P.; Drew, M.G.B.; Gomez-Garcia, C.J.; Ghosh, A. The crucial role of polyatomic anions in molecular architecture: Structural and magnetic versatility of five nickel (II) complexes derived from AN, N, O-donor schiff base ligand. Inorg. Chem. 2009, 48, 5848–5860. [Google Scholar] [CrossRef] [PubMed]
  35. Mukherjee, P.; Drew, M.G.; Tangoulis, V.; Estrader, M.; Diaz, C.; Ghosh, A. Facile strategies for the synthesis and crystallization of linear trinuclear nickel (II)-Schiff base complexes with carboxylate bridges: Tuning of coordination geometry and magnetic properties. Polyhedron 2009, 28, 2989–2996. [Google Scholar] [CrossRef]
  36. Romanović, M.Č.; Čobeljić, B.R.; Pevec, A.; Turel, I.; Spasojević, V.; Tsaturyan, A.A.; Shcherbakov, I.N.; Anđelković, K.K.; Milenković, M.; Radanović, D.; et al. Synthesis, crystal structure, magnetic properties and DFT study of dinuclear Ni(II) complex with the condensation product of 2-quinolinecarboxaldehyde and Girard’s T reagent. Polyhedron 2017, 128, 30–37. [Google Scholar] [CrossRef]
  37. Mahapatra, P.; Ghosh, S.; Giri, S.; Ghosh, A. The unusual intermediate species in the formation of Ni(II) complexes of unsymmetrical Schiff bases by Elder’s method: Structural, electrochemical and magnetic characterizations. Polyhedron 2016, 117, 427–436. [Google Scholar] [CrossRef]
  38. Blanchet-Boiteux, C.; Mouesca, J.M. End-on azido-bridged copper dimers: Spin population analysis and spin polarization effect as exhibited by valence-bond/broken symmetry, density functional methods. J. Am. Chem. Soc. 2000, 122, 861–869. [Google Scholar] [CrossRef]
  39. Biswas, A.; Drew, M.G.; Diaz, C.; Bauzá, A.; Frontera, A.; Ghosh, A. Cis–trans isomerism in diphenoxido bridged dicopper complexes: Role of crystallized water to stabilize the cis isomer, variation in magnetic properties and conversion of both into a trinuclear species. Dalton Trans. 2012, 41, 12200–12212. [Google Scholar] [CrossRef] [PubMed]
  40. Nanda, K.K.; Das, R.; Thompson, L.K.; Venkatsubramanian, K.; Paul, P.; Nag, K. Magneto-structural correlations in macrocyclic dinickel (II) complexes: Tuning of spin exchange by varying stereochemistry and auxiliary ligands. Inorg. Chem. 1994, 33, 1188–1193. [Google Scholar] [CrossRef]
  41. Wang, C.; Fink, K.; Staemmler, V. A quantum chemical ab initio study of the superexchange coupling in binuclear oxygen-bridged Ni (II) complexes. Chem. Phys. 1995, 192, 25–35. [Google Scholar] [CrossRef]
  42. Torić, F.; Pavlović, G.; Pajić, D.; Cindric, M.; Zadro, K. Tetranuclear Ni4 cubane complexes with high χT maxima: Magneto-structural analysis. CrystEngComm 2018, 20, 3917–3927. [Google Scholar] [CrossRef]
  43. Bu, X.H.; Du, M.; Zhang, L.; Liao, D.Z.; Tang, J.K.; Zhang, R.H.; Shionoya, M. Anion-directed assembly: Framework conversion in dimensionality and photoluminescence. J. Chem. Soc. Dalton Trans. 2001, 5, 593–598. [Google Scholar] [CrossRef]
  44. Biswas, A.; Das, L.K.; Drew, M.G.B.; Aromí, G.; Gamez, P.; Ghosh, A. Synthesis, crystal structures, magnetic properties and catecholase activity of double phenoxido-bridged penta-coordinated dinuclear nickel (II) complexes derived from reduced Schiff-base ligands: Mechanistic inference of catecholase activity. Inorg. Chem. 2012, 51, 7993–8001. [Google Scholar] [CrossRef] [PubMed]
  45. Bain, G.A.; Berry, J.F. Diamagnetic corrections and Pascal’s constants. J. Chem. Educ. 2008, 85, 532. [Google Scholar] [CrossRef]
  46. Sheldrick, G.M. Shelxs97, Program for Crystallographic solution and refinement. Acta Cryst. 2008, A64, 112. [Google Scholar] [CrossRef] [PubMed]
  47. Sheldrick, G.M. Shelxl16/6, Program for Crystallographic Refinement. Acta Cryst. 2015, C71, 3. [Google Scholar]
  48. Platon, A.L. Spek. Acta Cryst. 2009, D65, 148. [Google Scholar]
  49. Noodleman, L. Valence bond description of antiferromagnetic coupling in transition metal dimers. J. Chem. Phys. 1981, 74, 5737–5743. [Google Scholar] [CrossRef]
  50. Noodleman, L.; Case, D.A. Density-Functional Theory of Spin Polarization and Spin Coupling in Iron—Sulfur Clusters. Adv. Inorg. Chem. 1992, 38, 423–470. [Google Scholar]
  51. Ruiz, E.; Rodríguez-Fortea, A.; Cano, J.; Alvarez, S.; Alemany, P. About the calculation of exchange coupling constants in polynuclear transition metal complexes. J. Comput. Chem. 2003, 24, 982–989. [Google Scholar] [CrossRef] [PubMed]
  52. Becke, A.D. Becke’s three parameter hybrid method using the LYP correlation functional. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  53. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785. [Google Scholar] [CrossRef]
  54. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098. [Google Scholar] [CrossRef]
  55. Schäfer, A.; Huber, C.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr. J. Chem. Phys. 1994, 100, 5829–5835. [Google Scholar] [CrossRef]
  56. Neese, F. The ORCA program system. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  57. Giri, S.; Biswas, S.; Drew, M.G.B.; Ghosh, A.; Saha, S.K. Structure and magnetic properties of a tetranuclear Cu (II) complex containing the 2-(pyridine-2-yliminomethyl)-phenol ligand. Inorg. Chim.Acta. 2011, 368, 152–156. [Google Scholar] [CrossRef]
  58. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthesis of complexes 13.
Scheme 1. Synthesis of complexes 13.
Magnetochemistry 04 00051 sch001
Figure 1. Structures of 1 (A) and 2 (B) with ellipsoids at 30% probability (a = 1−x,y,3/2−z).
Figure 1. Structures of 1 (A) and 2 (B) with ellipsoids at 30% probability (a = 1−x,y,3/2−z).
Magnetochemistry 04 00051 g001
Figure 2. Structure of 3 with ellipsoids at 30% probability (a = 1−x,y,3/2−z).
Figure 2. Structure of 3 with ellipsoids at 30% probability (a = 1−x,y,3/2−z).
Magnetochemistry 04 00051 g002
Figure 3. Plot of χMT vs. T for complex 1. The circles are the experimental data, and the solid line is generated from the fitted curve.
Figure 3. Plot of χMT vs. T for complex 1. The circles are the experimental data, and the solid line is generated from the fitted curve.
Magnetochemistry 04 00051 g003
Figure 4. Plot of χMT vs. T for complex 2. The circles are the experimental data, and the solid line is generated from the fitted curve.
Figure 4. Plot of χMT vs. T for complex 2. The circles are the experimental data, and the solid line is generated from the fitted curve.
Magnetochemistry 04 00051 g004
Figure 5. Plot of χMT vs. T for complex 3. The circles are the experimental data, and the solid line is generated from the fitted curve.
Figure 5. Plot of χMT vs. T for complex 3. The circles are the experimental data, and the solid line is generated from the fitted curve.
Magnetochemistry 04 00051 g005
Figure 6. Environments of Ni(II) ions in complexes 13 with different bridging modes (phenoxido, chlorido). a Symmetry element 1−x, y, 3/2−z.
Figure 6. Environments of Ni(II) ions in complexes 13 with different bridging modes (phenoxido, chlorido). a Symmetry element 1−x, y, 3/2−z.
Magnetochemistry 04 00051 g006
Table 1. Hydrogen-Bond Parameters in Complexes (in Å and deg).
Table 1. Hydrogen-Bond Parameters in Complexes (in Å and deg).
D-H…Ad(D-H)d(D…A)d(H…A)<(D-H…A)
Complex 1N(19)-H(19)…Cl(2)$10.983.331(3)2.61130
Complex 2N(19)-H(19)…Cl(2)$10.983.332(2)2.56136
Complex 3N(19)-H(19)…Cl(2)$10.983.345(1)2.62131
In 1 and 2 $1 = ½−x, ½−y, 1−z; in 3 $1 = ½− x, 3/2−y,1−z.
Table 2. Magnetic coupling constants (J) and selective structural parameters for complexes 1, 2 and 3.
Table 2. Magnetic coupling constants (J) and selective structural parameters for complexes 1, 2 and 3.
Complex NameNi–O Bond Lengths (Å)Ni–Ni Distance (Å)Ni–O–Ni Angle(°)Coupling Constant (J) Theoretical (BS DFT) (cm−1)Coupling Constant (J) Experimental (cm−1)
12.164(2)
2.049(2)
2.963(1)89.32(8)4.253.97
22.164(2)
2.053(2)
2.967(1)89.39(6)3.564.66
32.034(1)
2.318(1)
3.011(1)87.32(4)5.505.50
Table 3. The spin densities on the selected atoms for complex 1.
Table 3. The spin densities on the selected atoms for complex 1.
Ni1−1.626
Na10.000
Cl10.000
Cl2−0.080
O11−0.026
Ni1a1.626
Cl2a0.080
O11a0.026
a Symmetry element 1−x,y,3/2−z.
Table 4. The spin densities on the selected atoms for complex 2.
Table 4. The spin densities on the selected atoms for complex 2.
Ni1−1.627
Na10.000
Cl10.000
Cl2−0.079
O11−0.024
Ni1a1.627
Cl2a0.079
O11a0.024
a Symmetry element 1−x,y,3/2−z.
Table 5. The spin densities on the selected atoms for complex 3.
Table 5. The spin densities on the selected atoms for complex 3.
Ni1−1.629
Na10.000
Cl10.000
Cl2−0.082
O11−0.033
Ni1a1.629
Cl2a0.082
O11a0.033
a Symmetry element 1−x,y,3/2−z.
Table 6. Crystallographic parameters for complexes 1, 2 and 3.
Table 6. Crystallographic parameters for complexes 1, 2 and 3.
Compound123
Chemical formulaC22H36Cl3N4NaNi2O3C22H34Cl5N4NaNi2O3C28H48Cl3N4Na Ni2O5
Formula weight651.31720.19767.46
Crystal systemMonoclinicMonoclinicMonoclinic
Space groupC2/cC2/cC2/c
a(Å)22.2673(8)17.9326(17)21.7316(10)
b (Å)11.8915(4)17.1847(16)12.1574(5)
c (Å)15.7543(10)11.7063(11)15.8456(12)
β, deg129.106(1)112.796(2)126.9090(10)
V (Å3) 3237.1(3)3325.7(5)3347.4(3)
Z444
ρcalc (g cm−3)1.3361.4381.523
µ (Mo Kα) (mm−1)1.4511.5751.420
F(000)135214801608
Reflections collected17,50740,19332,846
Independent reflections360329563815
Reflections with I > 2σ(I)303128083470
R1a, wR2b0.0489, 0.17450.0316, 0.09880.0269, 0.0706
R1],wR2 [all data]0.0572, 0.18340.0331, 0.10040.0308, 0.0705
GOF c1.0651.1251.036
Residual electron Density, e/Å−31.762, −0.3010.826, −0.4130.518, −0.378
a R1 = ∑||Fo|− |Fc||/∑|Fo|, b wR2 (Fo2) = [∑[w(Fo2−Fc2)2/∑w Fo4]½ and c GOF = [∑[w(Fo2−Fc2)2/(Nobs−Nparams)]½.

Share and Cite

MDPI and ACS Style

Mondal, M.; Chakraborty, M.; Drew, M.G.B.; Ghosh, A. Ni(II) Dimers of NNO Donor Tridentate Reduced Schiff Base Ligands as Alkali Metal Ion Capturing Agents: Syntheses, Crystal Structures and Magnetic Properties. Magnetochemistry 2018, 4, 51. https://doi.org/10.3390/magnetochemistry4040051

AMA Style

Mondal M, Chakraborty M, Drew MGB, Ghosh A. Ni(II) Dimers of NNO Donor Tridentate Reduced Schiff Base Ligands as Alkali Metal Ion Capturing Agents: Syntheses, Crystal Structures and Magnetic Properties. Magnetochemistry. 2018; 4(4):51. https://doi.org/10.3390/magnetochemistry4040051

Chicago/Turabian Style

Mondal, Monotosh, Maharudra Chakraborty, Michael G. B. Drew, and Ashutosh Ghosh. 2018. "Ni(II) Dimers of NNO Donor Tridentate Reduced Schiff Base Ligands as Alkali Metal Ion Capturing Agents: Syntheses, Crystal Structures and Magnetic Properties" Magnetochemistry 4, no. 4: 51. https://doi.org/10.3390/magnetochemistry4040051

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop