Next Article in Journal
Mercury and Organic Pollutants Removal from Aqueous Solutions by Heterogeneous Photocatalysis with ZnO-Based Materials
Next Article in Special Issue
Single-Molecule Magnet Properties in 3d4f Heterobimetallic Iron and Dysprosium Complexes Involving Hydrazone Ligand
Previous Article in Journal
Thermodynamics Investigation and Artificial Neural Network Prediction of Energy, Exergy, and Hydrogen Production from a Solar Thermochemical Plant Using a Polymer Membrane Electrolyzer
Previous Article in Special Issue
Crystal Structure and Magnetic Properties of Trinuclear Transition Metal Complexes (MnII, CoII, NiII and CuII) with Bridging Sulfonate-Functionalized 1,2,4-Triazole Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetostructural Properties of Some Doubly-Bridged Phenoxido Copper(II) Complexes

1
Department of Chemistry, University of Louisiana at Lafayette, P.O. Box 43700, Lafayette, LA 70504, USA
2
Department of Chemistry, Faculty of Science, Alexandria University, Moharam Bey, Alexandria 21511, Egypt
3
Institut für Anorganische Chemische, Technische Universität Graz, Stremayrgasse 9/V, A-8010 Graz, Austria
4
Institut für Physikalische and Theoretische Chemie, Technische Universität Graz, Stremayrgasse 9/II, A-8010 Graz, Austria
5
Department of Chemistry, Interdisciplinary Graduate School of Science and Engineering, Shimane University, 1060 Nishikawatsu, Matsue 690-8504, Japan
6
Department of Applied Chemistry for Environment, School of Biological and Environmental Sciences, Kwansei Gakuin University, 1 Gakuen Uegahara, Sanda 669-1330, Japan
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(6), 2648; https://doi.org/10.3390/molecules28062648
Submission received: 13 February 2023 / Revised: 10 March 2023 / Accepted: 10 March 2023 / Published: 14 March 2023
(This article belongs to the Special Issue Design of Coordination Compounds with Novel Magnetic Properties)

Abstract

:
Three new tripod tetradentate phenolate-amines (H2L1, H2L4 and H2L9), together with seven more already related published ligands, were synthesized, and characterized. With these ligands, two new dinuclear doubly-bridged-phenoxido copper(II) complexes (3, 4), and six more complexes (1, 2, 58), a new trinuclear complex (9) with an alternative doubly-bridged-phenoxido and –methoxido, as well as the 1D polymer (10) were synthesized, and their molecular structures were characterized by spectroscopic methods and X-ray single crystal crystallography. The Cu(II) centers in these complexes exhibit distorted square-pyramidal arrangement in 14, mixed square pyramidal and square planar in 5, 6, and 9, and distorted octahedral (5+1) arrangements in 7 and 8. The temperature dependence magnetic susceptibility study over the temperature range 2–300 K revealed moderate–relatively strong antiferromagnetic coupling (AF) (|J| = 289–145 cm−1) in complexes 16, weak-moderate AF (|J| = 59 cm−1) in the trinuclear complex 9, but weak AF interactions (|J| = 3.6 & 4.6 cm−1) were obtained in 7 and 8. No correlation was found between the exchange coupling J and the geometrical structural parameters of the four-membered Cu2O2 rings.

Graphical Abstract

1. Introduction

Dinuclear copper(II) complexes in which the Cu(II) ions are bridged by a phenolate oxygen atoms (L-O) and an exogenous bridging atom X with the coordination core Cu(µ-L-O)(µ-X) Cu (X = L-O, OH, OR, O2−, OAc, N3and Cl) have been reported for a number of ligands with different skeletons frames [1,2,3,4,5,6,7,8,9,10,11,12]. Some of these complexes have served as cooperative model systems to mimic the active sites in metalloproteins, such as hemocyanin [12,13,14,15,16], metalloenzymes, such as tyrosinase [16,17,18], galactose oxidase (GOase) [19,20,21,22,23,24], superoxide dismutase (SOD) [25,26,27], and as artificial nucleases for promoting the cleavage of DNA [5,6,28,29] as well as anticancer agents [30,31].
The dinuclear copper(II) compounds constructed from the cores Cu(µ-L-O)(µ-X)Cu and Cu(µ-L-O)2Cu are stable and the two metal ions are located in close proximity, which lead to strong implications regarding their reactivity and magnetic properties due to the interactions between the two paramagnetic Cu(II) centers [32,33]. Different strategies have been employed in the design of this class of compounds. The first approach was achieved though the interaction Cu(II) salts with tridentate ligand-based phenolates in their skeletons [34,35,36,37,38,39,40,41], and the second approach was via the design of bicompartmental phenolates bearing two pendent arms of N-donor groups [1,2,3,4,5,6,42,43]. A third approach was also used through the design of tripodal pyridyl tetradentate ligands containing one or two phenolate arms [12,14,15,20,29,44,45,46,47,48,49,50,51].
The structures and magnetic properties of singly-bridged Cu(µ-L-O)Cu and doubly-bridged phenolate Cu(µ-L-O)2Cu compounds were the subject of numerous investigations [1,2,3,4,31,33,34,41,42,44] as well as theoretical studies [31,33,41,51,52,53,54]. The energy gap 2J arising from the spin singlet-triplet (S-T) state interactions between the two local Cu(II) doublets were evaluated. Many structural parameters were reported to affect the magnitude of the magnetic interactions in the four membered ring, Cu2O2 [41,54]. These include the Cu···Cu bond distance and the bridged Cu−O−Cu bond angle [41,54], the geometrical distortion, the dihedral angle between the two copper planes as well as the electronegativity in the bridging phenolate [53]. The presence of electron-withdrawing groups leads to substantial reduction in the antiferromagnetic exchange [53]. The steric effect incorporated into ligand skeleton, which may distort the copper coordination geometry, cannot be ruled out either. The non-covalent bonding interactions (H-bonding, π···π, stacking interactions, and van der Waals’ force) have also been reported to affect the molecular magnetism in these systems [54,55,56,57,58,59,60,61,62]. In general, strong super exchange antiferromagnetic coupling (AF) was observed in the bridged phenolate compounds [25,26,27,33,41,43,51,52,53,54] but, in a few cases very weak ferromagnetic properties (F) were found [4,22,25,54].
Herein, we describe a general effective procedure for the design of novel series of tripodal phenolate compounds containing pyridyl or aliphatic amine arms with N3O and N2O2 chromophores, where the donor O atoms are provided by one or two phenolate groups (Scheme 1) and their corresponding doubly-bridged phenoxido-Cu(II) complexes. The magnetic properties of the complexes were investigated at variable temperatures and the evaluated exchange coupling constants, J, are discussed in relation to their molecular structures.

2. Results and Discussion

2.1. Synthesis of the Ligands and Complexes

The new tetradentate tripodal-phenolate amines (H2L1, H2L4 and H2L9) used in this work, together with related ligands (Scheme 1), were synthesized by a standard general procedure. In a typical experiment, a methanolic mixture containing 2,4-disubstituted phenol and the corresponding amine (2:1), as well as two molar amounts of an aqueous 37% HCHO and triethylamine (Et3N) are used, whereas in the case of 2-methoxy-phenols and pyridyl derivatives (HL7 and HL8 ligands, Scheme 1), equimolar ratios were employed for all reagents. Refluxing the solutions for 3 to 4 days, followed by evaporation of the solution resulted in the formation of the solid products in reasonable yields, but about 30% were observed in the bromo-phenolate H2L3. We should mention that too much evaporation of the resulting solutions may lead to the formation of an oily product and/or a semi-solid, which is hard to recrystallize and leads to impure products. The pure desired products were obtained upon recrystallization from ethyl acetate and activated charcoal (see Experimental section). The synthesis and characterization of H2L2, H2L3, H2L5, H2L6, HL7, HL8, and H2L10 were recently reported by our group [30].
The doubly-bridged phenoxido Cu(II) complexes 16, 9, and the polymeric 1D 10 were obtatined by the reaction of a methanolic solution of Cu(NO3)2·3H2O with the corresponding tripodal-phenolate amine ligand and Et3N in the stochiometric 1:1:2 molar ratio. In the cationic complexes 7 and 8, copper(II) perchlorate was used with the bipyridyl-phenolates HL7 and HL8, and Et3N (1:1:1), respectively. These reactions afforded the desired complexes in moderate to good yields (55–90%). An illustration for the synthesis of the doubly-bridged phenoxido moieties in the complexes derived from N,N-dialkylethylenediamines is shown in Scheme 2. Interestingly, although all double bridged phenoxido compounds display green or olive-green color, brownish-green and purple colored complexes were obtained in case of the trimeric complex 9 and the 1D-polymeric 10. Single crystals suitable for X-ray structural determination were obtained either from dilute methanolic solution and/or recrystallization from CH3CN or acetone. The complexes are slightly soluble in MeOH, but are more soluble in less polar solvents, such as CH3CN and DMSO. The isolated compounds were characterized by elemental microanalyses, spectroscopic techniques, and conductivity measurements as well as by single X-ray crystallography for the copper complexes.

2.2. Characterization of the Ligands

Some very general features exist in the IR spectra of the tetradentate tripodal phenolate amine ligands, such as a very weak broad band or a shoulder over the frequency region 3140–3230 cm−1, which was attributed to the stretching vibration, ν(O-H) of the phenolic groups, in addition to a series of weak to very weak bands observed over the range 2700–3050 cm−1 due to ν(C-H) stretching of the aliphatic and aromatic groups. The moderate intense band detected around the 1590–1600 cm−1 region is assigned to ν(C=N), whereas the moderate to strong series of bands shown in 1200–1590 cm−1 are most likely attributed to ν(C=C, C-O). The ESI-MS of the ligands in MeOH, showed the 100% m/z base peak that corresponds to the protonation of parent ligand (H2Ln + H]+ (n =2–5, 9, 10). The 1H NMR (d6-DMSO) spectra displayed peak positions at 7.1–6.8 (protons-ph); 5.0–4.9 (phenolate-protons), 3.8–3.7 (CH2-py); 3.5–3.3 (CH2-ph); 2.5–2.1 ppm (CH3-ph). The pyridyl protons reveal their signals at δ = 8.7–7.1 in compounds H2L1, HL4 and HL8. The hydroxyl phenolate protons were clearly observed in H2L1 and H2L5 compounds but was not seen in most compounds, most likely due to their low solubility in DMSO.

2.3. Characterization of the Complexes

The IR spectra of the complexes under investigation were almost similar to the spectral pattern observed in their parent ligands, except for the disappearance of the broad band or shoulder in the region 3140–3230 cm−1 of the ν(O-H) of the phenolic groups upon its deprotonation and/or coordination to the Cu2+ ion. The two cationic perchlorate complexes 7 and 8 displayed a very strong band at 1079 and 1076 cm−1, respectively, due to νas(O-Cl) of the perchlorate counter ions. The O-H stretching frequency, ν(O-H), for the water of crystallization in 5 and 6 and were shown as a broad band over the 3320–3400 cm−1 region. The ESI-MS (CH3CN) of the cationic pyridyl complexes 7 and 8 revealed the monomeric species with m/z = [Cu(L7,8)]+ (100%), where the base peak (100%) of the remaining complexes was consistent with the release of the parent protonated ligands with m/z = [H2L2−6,9,10 + H]+.
The UV-vis spectra of the complexes under investigation, measured in DMSO and/or CH3CN at room temperature, displayed a single broad/shoulder band in the 610–750 nm region and a strong absorption band over wavelength region 410–500 nm. The latter intense band can be assigned to the bridged phenoxido charge transfer transition (L-O → CuII LMCT) [12,20,63,64], whereas the former low intense broad band is attributed to d-d transition in five-coordinate Cu(II) complexes, which was occasionally accompanied with a weaker intense broad band around 850–890 nm. The d-d transition feature in solution is consistent with a distorted square pyramidal geometry (SP) around the central Cu(II) ion [32,48,49]. Thus, the distorted SP geometrical assignments observed in DMSO, CH3CN, or acetone solution, were retained in the solid state (see X-ray section). The solution spectra of the complexes in these media did not show any appreciable change over the one-week period, reflecting their high stability.
The position of λmax in the 610–750 nm region can be used as a criterion for the ligand field strength of the tripodal phenolate ligands; λmax decreases in the order: 9max = 748 nm) > 2max = 733 nm) > 5max = 708 nm) > 10max = 687 nm) > 6max = 655 nm) > 48max = 630 nm) > 7max = 625 nm) > 1max = 620 nm). This means that the ligand field strength is decreasing in the reverse order and the strongest fields are for phenolate compounds containing pyridyl arms. However, we should mention that broadening and close location of this band in some complexes did not permit precise prediction of their actual ligand field strength, but results are consistent with previous results [65,66]. In general, ligand field strengths decrease with increasing chelate ring size, the presence of electronegative chlorine or bromine atoms in the phenolate groups, and/or steric effect on the coordinated N-aliphatic amine arms, which tends to reduce the electron density on the coordinated centers and hence its ligand field [65,66].
The non-electronic nature of all complexes, with the exception of 7 and 8, was supported by measuring the molar conductivities, ΛM in CH3CN or DMSO, whenever the solubility permits. The measured ΛM values were ≤ 5 Ω−1·cm2·mol−1, which were fully consistent with their non-ionic properties as predicted by their molecular formulas. The measured ΛM values (DMSO) of the perchlorate complexes 7 and 8 were within the range of 280 ± 2 Ω−1·cm2·mol−1. These values are in full agreement with the 1:2 electrolytic behavior of the doubly-bridged phenoxido compounds ([Cu22-L7,8)2](ClO4)2 (7, 8) [30,67].

2.4. Description of the Structures of Complexes

The crystal structure of title compounds 3 and 4 consist of neutral dinuclear [Cu2(L3)2] and [Cu2(L4)2] units, respectively; the latter co-crystallizes with an acetone solvent molecule per dinuclear unit (Figure 1,b). In both structures, the phenolate oxygen atoms O2 and O4 of two tripod ligand L3 or L4 anions are bridging the two copper metal centers to form four-membered Cu2O2 rings [Cu-O from 1.9465(16) to 2.0349(16) Å, Cu1 Cu2 = 3.047(2) (3) and 2.9949(4) Å (4); Cu1-O-Cu2 from 97.17(7) to 100.4(4)°, O-Cu-O from 74.8(4) to 75.78(7)°] (Table 1). Each Cu1 center is penta-coordinated with further three terminal sites occupied by the amine N1 and N2 atoms as well as the phenolate O1 atom. Each Cu2 center is also penta-coordinated with amine N3, N4, and phenolate O3 atoms. While the CuN2O3 chromophore of Cu1 may be described as less distorted square pyramid (SP) with a τ5 -value of 0.06; the CuN2O3 chromophore of Cu2 is strongly distorted SP with τ5 -value of 0.21 (3) and 0.45 (4) (τ = 0 for ideal SP) and τ5 = 1 for ideal trigonal bipyramid (TBP)) [68]. Their apical positions are occupied by N2 and N4 atoms [Cu-N(apical) from 2.180(10) to 2.449(2) Å]. The Cu-N1/N3(basal) bonds vary from 2.0627(19) to 2.160(10) Å, and the Cu-O1/O3(basal) bonds from 1.8783(17) to 1.905(8) Å. The trans-basal bond angles vary from 138.62(7) to 165.55(7)°.
The non-planar Cu2O2 four-membered rings have hinge distortion expressed by their Cu-O-Cu-O dihedral angles of 24.3° and 26.9° for 3 and 4, respectively. The phenoxido groups deviate from phenyl out-of-plane angles (τ) of 10.5 and 12.2° in 3, and of 3.2 and 12.8° in 4, and phenyl ring torsion angles Cu-O-C-C of 50.9 and 71.0° in 3 and 52.5 and 67.5° in 4.
The crystal structure of 9 features trinuclear complex units (Figure 1c) and partially disordered MeOH solvent molecules. The phenolato O2 and O5 atoms of two tripod ligand L9 anions and O3 and O4 of two MeO anions are bridging the central Cu2 center with the two external Cu1 and Cu3 centers to form two four-membered Cu2O2 rings [Cu-O from 1.920(2) to 2.000(2) Å, Cu1···Cu2 = 2.9652(5), Cu2···Cu3 = 2.9418(5) Å; Cu1···Cu2···Cu3 = 154.63(2)°; Cu-O-Cu from 96.48(9) to 100.19(9)°, O-Cu-O from 75.43(8) to 77.91(8)°] (Table 1). The four oxygen atoms around Cu2 center form a tetragonally distorted square planar CuO4 geometry with a τ4-value of 0.17 (τ4 = 0 for ideal square planar (SQP) and τ4 = 1 for ideal tetrahedral, (Td) geometry) [9]. Coordination number 5 around the Cu1 and Cu3 center is completed by two N and one O donor atoms of two tripod L9 anions. Their CuN2O3 chromophores form SP geometry with τ5(Cu1) of 0.01 and τ5(Cu3) of 0.23. Their apical positions are occupied by N2 and N4 atoms [Cu1-N2 = 2.351(2), Cu3-N4 = 2.324(2) Å]. The Cu-N1/N3(basal) bonds are 2.054(2) and 2.066(2) Å, and the Cu-O1/O3(basal) bonds are 1.929(2) and 1.944(2) Å. The trans-basal bond angles vary from 153.32(9) to 167.39(9)°. The non-planar Cu2O2 four-membered rings have hinge distortion expressed by their Cu-O-Cu-O dihedral angles of 22.9° and 23.5°. The phenoxido groups deviate from phenyl out-of-plane angles, τ of 8.2 and 20.0°, and phenyl ring torsion angles Cu-O-C-C of 44.2 and 41.9°. The trinuclear subunit may formally be created by the insertion of a “Cu(MeO)2” moiety in the center of a bis(phenolato)-bridged Cu2O2 four-membered ring of a dinuclear compound (e.g., of 3).
Perspective views of the dinuclear complexes 1, 2, 5–8 containing the Cu2O2 bis(phenolato)-bridged subunits, as well as a section of the polymeric chain of 10, are presented in Figure 2. The crystal structures have been reported previously, along with their anticancer properties [30].

2.5. Magnetic Properties

According to the crystal structures, the complexes under investigations can be divided into six groups as follows:
Group 1: Centrosymmetric di-µ-phenoxido-bridged dinuclear complexes with a distorted square-pyramidal geometries (1 and 2). Complex [Cu2(L1)2] (1) shows a distorted square-pyramidal arrangement (SP) (τ5 = 0.23) around each copper(II) center with an axial Cu-N distance of 2.310(3) Å, Cu···Cu distance and Cu-O-Cu angle are 3.0915(7) Å and 103.44(9)°, respectively [30]. Thus, Complex 1 is expected to exhibit antiferromagnetic interaction (AF) between the two S = ½ spins through the two phenoxido bridges. Temperature dependence of magnetic susceptibilities, measured under an external magnetic field of 0.5 T, ranging from 2 K to 300 K are represented in Figure 3 in the form of χM and µM vs T plots, where χM is the magnetic susceptibility per Cu2 unit, µM is the magnetic moment per Cu2 unit, and T is the absolute temperature. The magnetic moment of 1.30 BM at 300 K is considerably lower than the spin-only value (2.45 BM) for two non-interacting copper(II) S = ½ ions. The magnetic moment exhibits a continuous decrease with lowering the temperature over the range 2–300 K and reaches 0.28 BM at 2 K. This magnetic behavior reveals significant AF interaction between the copper(II) ions. The magnetic data were analyzed by the Bleaney–Bowers equation based on the Heisenberg model:
χM = (1 − p)(2Ng2µB2/kT)[3 + exp(−2J/kT)]−1 + pNg2µB2/2kT + 2
where g is the g value, J is the exchange coupling constant between the two copper(II) ions, p is the fraction of mononuclear copper(II) impurity, and is the temperature-independent paramagnetism for each copper(II) ion [8]. The best-fitting parameters (g = 2.1 (fixed), 2J = −578 cm−1, p = 0.016, and = 60 × 10−6 cm3 mol−1 (fixed)) are in complete support for strong AF interaction.
The crystal structure of [Cu2(L2)2] (2) also shows the same crystallographic similarity as 1 (τ = 0.24) [68] around each copper(II) center with the axial Cu-N distance of 2.3580(17) Å. The Cu···Cu distance and Cu-O-Cu angle are 3.0064(5) Å and 99.12(6)°, respectively. The µM of 2 is 1.62 BM per dinuclear unit at 300 K, which is also considerably lower than the spin-only value and decreases with lowering of the temperature and reaching to 0.12 BM at 2 K. The best-fitting parameters (g = 2.10, 2J = −403 cm−1, p = 0.0007, and = 60 × 10−6 cm3 mol−1) disclose the expected strong AF interaction. The absolute 2J value is smaller than that of 1, but it reflects the strong AF interaction in 1 compared 2 and is in harmony with the associated wider Cu-O-Cu angle observed in 1.
Group 2: Pseudo-symmetric di-µ-phenoxido-bridged dinuclear complexes with distorted SP geometries (3 and 4). The crystal structure of [Cu2(L3)2] (3) showed that the complex exhibits distorted SP arrangements (τ = 0.06 and 0.37) around two copper(II) centers with axial Cu-N distances of 2.180(10) and 2.235(10) Å. The Cu···Cu distance is 3.047(2) Å and Cu-O-Cu angles are 98.8(4) and 100.4(4)°. Complex 3 crystallizes in non-centrosymmetric with different distortion around each copper(II) center. The magnetic properties of 3 is illustrated in Figure 4. The magnetic moment of 2.07 BM at 300 K is a little higher than those of 1 and 2, and it exhibits a gradual decrease with lowering the temperature, reaching 0.20 BM at 2 K. The magnetic data were analyzed by Equation (1), and the obtained best-fitting parameters (g = 2.14(1), 2J = −291 cm−1, p = 0.009, and = 60 × 10−6 cm3 mol−1 (fixed)) demonstrates the AF interaction between the two copper(II) centers. The absolute 2J value of 3 is smaller than those of 1 and 2, but with a weaker interaction. This may be attributed to the different distortion between the two copper(II) coordination environments, which causes poor overlap between the two magnetic orbitals compared to 1 and 2.
The complex [Cu22-L2)2] (4) exhibits crystallographic similarities comparable to 3. The molecule is a non-centrosymmetric di-µ-phenoxido-bridged complex with distorted SP arrangements (τ5 = 0.07 and 0.45) around the central copper(II) centers with the axial Cu-N distances of 2.449(2) and 2.3532(19) Å. The Cu···Cu distance and Cu-O-Cu angles are 2.9949(4) Å and 97.17(7) and 97.78(7)°, respectively. The magnetic moment µM (1.98 BM at 300 K) shows a gradual decrease with lowering the temperature, reaching 0.20 BM at 2 K. The magnetic data were analyzed by Equation (1) and the obtained best-fitting parameters (g = 2.10 (fixed), 2J = −293 cm−1, p = 0.008, and = 60 × 10−6 cm3 mol−1 (fixed)) showed a comparable AF interaction to that observed in 3. This is in accordance with the relationship between the 2J and Cu-O-Cu values.
Group 3: Unsymmetric di-µ-phenoxido-bridged dinuclear complexes with distorted square-pyramidal (SP) and square-planar (SQP) geometries (5 and 6). The crystal structure of complex [Cu2(L5)2(H2O)]·2H2O (5) showed that the molecule exhibits a distorted SP copper(II) (τ5 = 0.011) with an axial Cu-O bond of 2.538(3) Å, and a distorted SQP copper(II) with the τ4 value of 0.17. Thus, if we consider the long axial distance of 2.538(3) Å, then both of the magnetic orbitals of the two copper(II) centers are expected to exist within the basal NO3 and NO3 square plane, favoring a strong magnetic interaction. The Cu···Cu distance and Cu-O-Cu angles are 2.9331 Å, and 96.70(9) and 97.58(9)°, respectively. The magnetic moment, µM (1.82 BM at 300 K), is lower than the spin-only value (2.45 BM) for two non-interacting copper(II) S = ½ ions. The magnetic moment revealed a gradual decrease with lowering the temperature and reaches 0.31 BM at 2 K (Figure 5). The best-fitting parameters to Equation (1) (g = 2.1 (fixed), 2J = −352 cm−1, p = 0.036, and = 60 × 10−6 cm3 mol−1 (fixed)) are consistent with considerable AF interaction between the two copper(II) ions. The absolute 2J value is comparable to those obtained in 3 and 4.
The crystal structure of [Cu2(L6)2(H2O)]·2H2O (6) shows crystallographic parameters similar to 5, having a distorted SP copper(II) (τ5 = 0.013) with the axial Cu-O bond of 2.512(2) Å and a distorted SP copper(II) with τ4 value of 0.17. The Cu···Cu distance and Cu-O-Cu angles are 2.9410(4) Å and 97.56(7) and 96.93(7)°, respectively. The magnitude of the magnetic moment of 6 is 1.68 BM at 300 K. This value is lower than the spin-only value for copper(II), S = ½ ion, and the magnetic moment exhibits a continuous decrease with lowering temperature and reaches 0.11 BM at 2 K. The best-fitting parameters to Equation (1) (g = 2.10 (fixed), 2J = −500 cm−1, p = 0.0009, and Na = 170 × 10−6 cm3 mol−1) showed a significant AF interaction between the two copper(II) ions. The magnetic behavior of 6 should be considered as similar to that of 5, irrespective of the relatively large –J value of 5, taking into consideration the similarity of the Cu-O-Cu angle, Cu-O-Cu-O dihedral angle, phenyl ring torsion angle (Cu-O-C-C angle), and phenyl out-of-plane shift angle t (Table 2), then one should expect strong AF interaction between the copper(II) centers in the two compounds 5 and 6.
Group 4: Centrosymmetric di(phenoxido)-bridged dinuclear complexes with a distorted (5 + 1) octahedral geometries (7 and 8). The crystal structural analysis of [Cu2(L7)2](ClO4)2 (7) shows a distorted octahedral (5 + 1) geometry with a semi-coordination Cu-OMe of 2.785(4) Å around each copper(II) center. If this semi-coordination is neglected as a substantial bond because of its long distance, then the coordination geometry can be regarded as a distorted square-pyramidal geometry with the τ5 value of 0.41. Therefore, taking into consideration the long distance of the axial Cu-O(phenoxido) bond of 2.213(3) Å, as a result the magnetic orbital of each copper(II) ion may exist around the basal N3O plane mainly and not in the direction of the axial phenoxido-bridging bond, which leads to a poor magnetic interaction. The Cu···Cu distance and Cu-O-Cu angle are 3.130 Å and 97.72 (14)°, respectively. The magnetic data of 7 are illustrated in Figure 6. The µB (2.65 BM at 300 K) is higher than the spin-only value (2.45 BM) for two non-interacting copper(II), S = ½ ions. The magnetic moment exhibits a continuous decrease with lowering temperature and reaches 0.51 BM at 2 K, suggesting an AF interaction. The best-fitting parameters (g = 2.14, 2J = −7.2 cm−1, and = 60 × 10−6 cm3 mol−1 (fixed)) are consistent with a weak AF interaction between the two copper(II) ions.
Complex [Cu2(L8)2](ClO4)2 (8) exhibits very close crystallographic parameters like 7; distorted octahedral (5 + 1) arrangement with a semi-coordination of Cu-OMe of 2.791(3) Å around each copper(II) ion. The coordination geometry of this complex is similar to 7 and can be regarded as indicated above as a distorted SP geometry (τ5 value of 0.40) with a relatively long axial phenoxido-bridging of Cu-O distance of 2.215(2) Å. The Cu···Cu distance and the Cu-O-Cu angle are 3.098(2) Å and 96.63(10)°, respectively. The similarity was also extended to the magnetic properties of 7. The best-fitting parameters (g = 2.1 (fixed), 2J = −9.2 cm−1, and = 60 × 10−6 cm3 mol−1 (fixed)) shows a weak AF interaction between the two copper(II) ions.
In 7 and 8, the Cu2O2 bridging ring with the axial phenoxido-bridges is planar, as can be seen from the Cu-O-Cu-O dihedral angle of 0° (Table 2), and thus the magnetic orbitals of each copper(II) ion should be parallel to each other. This suggests that the phenoxido oxygens are rather axially located relative to the basal dx2-y2 orbital plane. Such axial bridging oxygens cannot effectively participate in mediating the magnetic interaction through the phenoxido-bridges.
Group 5: linear trinuclear complex 9. The crystal structure of [Cu3(L9)2(µ-OCH3)2]·CH3COCH3 (9) shows that the molecule has a µ-phenoxido-µ-methoxido-bridged trinuclear copper(II) complex with a linear arrangement of the three copper(II) ions. The magnetic data (Figure 7) revealed a magnetic moment of 2.50 BM at 300 K, which is lower than the spin-only value (3.00 BM) for three non-interacting copper(II), S = ½ ions. The magnetic moment exhibits a gradual decrease with lowering temperature and reaches 1.46 BM at 2 K. This behavior suggests a remarkable AF interaction between the copper(II) ions. The magnetic analysis was carried out with the susceptibility equation for the linear trinuclear copper(II) ions based on H = −2JCu3(S1·S2 + S2·S3). However, the fitting was not good, and rather better fitting was obtained when the magnetic data were analyzed by the dinuclear model using Equation (1). The results indicated that the magnetic data of 9 contain some impurities arising from the presence of dinuclear copper(II) species. Therefore, the magnetic data were analyzed with the aid of Equation (2) of linear trinuclear and dinuclear copper(II) model.
χM = (1 − pCu2)(Ng2µB2/4kT)[1 + exp(−2JCu3/kT) + 10exp(JCu3/kT)]/[1 + exp(−2JCu3/kT) + 2exp(JCu3/kT)] + pCu2(2Ng2µB2/kT)[3 + exp(−2J/kT)]−1 + (3−pCu2)
where pCu2 is the fraction of dinuclear copper(II) impurity. The best-fitting parameters (g = 2.1 (fixed), JCu3 = −59 cm−1, pCu2 = 0.315, 2J = −592 cm−1, and = 60 × 10−6 cm3 mol−1 (fixed)) revealed the existence of a considerable amount of dinuclear copper(II) species in the sample.
Group 6: polynuclear chain complex 10. The X-ray structure analysis revealed that catena-[Cu(µ-L10)] (10) is a polynuclear chain molecule constructed of tetrahedrally distorted square-planar [Cu(L10)] units (τ4 = 0.23) [9] with Cu···Cu distance of 7.269 Å. The temperature dependence of magnetic susceptibilities χA and magnetic moments µA are represented in Figure 8, where χA and µA are per Cu unit. The µA value of 1.83 BM at 300 K is a little bit higher than the spin-only value (1.73 BM) for copper(II), S = ½ ion. The µA vs. T plot shows a slight gradual increase with a lowering of the temperature over the range 50–300 K and a gradual decrease on lowering the temperature from 50 to 2 K, reaching, 1.70 BM at 2 K. While the increase of the µA value suggests F interaction between two adjacent Cu(II) centers via the L10 anion ligand, the corresponding decrease in µA values indicates an AF intermolecular interaction. The magnetic data were analyzed by the molecular field approximation (Equation (3) [8]), where for this series Equation (4) for the Heisenberg model for ferromagnetically coupled S = ½ ions (derived by Baker et al. [10]), considering the magnetic interaction between the neighboring chain molecules as zJ’ (z = number of interacting neighbors), was used:
χA’ = χA/{1−(2zJ’/Ng2µB2)χA}
χA = (Ng2µB2/4kT)[(1.0 + 5.7979916x + 16.902653x2 + 29.376885x3 + 29.832959x4 + 14.036918x5)/(1.0 + 2.7979916x + 7.0086780x2 + 8.6538644x3 + 4.5743114x4)]2/3 +
where x = J/2kT. The solid line in Figure 8 shows the calculated curve with best-fitting parameters of g = 2.1 (fixed), J = 3.5 cm−1, = 60 × 10−6 cm3 mol−1 (fixed), and zJ’ = −2.9 cm−1.

3. Conclusions

In this work eight dinuclear doubly-bridged-phenoxido copper(II) complexes revealed antiferromagnetic coupling (AF) varied from strong AF interaction (−J = 289–145 cm−1) for complexes 16 to very weak AF interaction (−J = 3.6 & 4.6 cm−1) in 7 and 8. The geometry around the central Cu(II) centers exhibits distorted square-pyramidal arrangement in the first four complexes, mixed square pyramidal and distorted square planar in 5 and 6, and distorted pseudo octahedral (5 + 1) arrangements in 7 and 8. Attempts were made to correlate the exchange coupling (J) to the structural parameters of the non-planar Cu2O2 four-membered rings. These parameters include the Cu···Cu distances (2.941–3.130 Å), Cu-O-Cu bond angles (96.93–103.44°) [69,70], Cu-O-Cu-O dihedral angles (0–29.8°), and the deviation angles of the phenoxido groups from the phenyl out-of-plane angles (11.3–71.0°) as well as the phenyl ring torsion angles Cu-O-C-C (2.7–31.3). Unfortunately, no satisfied correlations were observed between J and the mentioned geometrical parameters [69,70]. This is most likely due to the influence of the steric environment incorporated by the substituents into phenolate groups, and the terminal coordinated dialkylamine, and hence the geometrical arrangement around the central Cu(II) ions as well as the coplanarity of the phenolate groups with the bridging Cu2O2 moiety [53,70,71]. These factors may play a key role in reducing the efficient overlap between the 3d Cu2+ and 2p O-phenoxido magnetic orbitals bearing the unpaired electrons. In general and focusing the discussion only on Cu(II) complexes containing mono- and doubly-bridged phenoxido compounds with no other bridging ligands, our results agree with the magnetic properties determined in many phenoxido-bridged copper complexes, where relatively strong AF couplings were reported [1,54,69,70,72], regardless of the fact that some of these compounds were ferromagnetically coupled [1,54,72,73].
In contrast and unlike the bridged bis(phenoxido) dinuclear copper(II) complexes, where no correlation was observed between J and Cu–O–Cu bond angle, linear relationships were obtained in the corresponding bis(hydroxido) [74] and bis(alkoxido) [75] bridged copper(II) complexes. In addition, a similar linear correlation was reported in bis(phenoxido) macrocyclic complexes in which the unique conjugated π-electron inherited into the macrocycle skeleton constrained the ligand to adopt a planar configuration [71]. On the other hand, a relatively moderate AF coupling (-J = 59 cm−1) was evaluated in the trinuclear 9, where alternative bridged phenoxido and methoxido groups were determined. However, in the 1D polymer 10, a weak ferromagnetic interaction (J = + 3.5 cm−1) was obtained. Interestingly, in the latter complex, the Cu(II) centers were linked via the N,N-dimethylpropyl arms with long Cu···Cu distance (7.269 Å) and a tetrahedral distorted square planar four-coordinate geometry around each Cu(II) center [30].

4. Experimental

4.1. Materials and Physical Measurements

N,N-Dimethyl-, N,N-diethyl-, N,N-isopropyl-ethylenediamine, 3-dimethylamino-propylamine, 2,4-dimethylphenol, 4-chloro-2-methylphenol, 2-bromo-4-methylphenol, 2-methoxy-4-methylphenol, 4-chloro-2-methoxyphenol and 2-tert-butyl-4-methylphenol, picolylamine as well as dipicolylamine were purchased from TCI-America. All other chemicals were commercially available and used without further purification.
Electronic spectra were recorded using an Agilent 8453 HP diode array UV-Vis spectrophotometer. Infrared spectra were recorded on a Cary 630 (ATR-IR) spectrometer. 1H spectra were obtained at room temperature on a Varian 400 NMR spectrometer operating at 400 MHz (1H). 1H NMR chemical shifts (δ) are reported in ppm and were referenced internally to residual solvent resonances (d6-DMSO: δH = 2.49) or TMS. ESI-MS of organic compounds and their Cu(II) were measured in MeOH and CH3CN, respectively, on a LC-MS Varian Saturn 2200 Spectrometer. Conductivity measurements were performed using a Mettler Toledo Seven Easy conductivity meter and calibrated by 1413 μS/cm conductivity standard. The molar conductivity of the complexes was determined from ΛM = (1.0 × 103 κ)/[Cu(II)], where κ = specific conductance and [CuII] is the molar concentration of the complex. Elemental microanalyses were carried out by Atlantic Microlaboratory, Norcross, Georgia U.S.A.
The temperature dependent magnetic susceptibilities were measured over 2–300 K at the constant field of 0.5 T with a Quantum Design MPMS 3 (installed at Shimane University) for 13, 4, 610, and with a Quantum Design MPMS-7 (installed at Institute for Molecular Science (IMS)) for 5. The measured data were corrected for diamagnetic contributions [76].

4.2. Synthesis of the Organic Ligands

The new ligands H2L3, H2L4 and H2L9 illustrated in Scheme 1, and their characterization are given below, whereas the rest of the ligands shown in this scheme (H2L1, H2L2, H2L5, H2L6, and H2L10 as well as HL7 and H2L8) were recently reported [30].

4.2.1. 6,6′-(((2-(Dimethylamino)ethyl)azanediyl)bis(methylene))bis(2-bromo-4-methylphenol) (H2L3)

To a mixture containing 2-bromo-4-methylphenol (3.740 g, 20 mmol), Et3N (2.04 g, 20 mmol) and aqueous 37% HCHO (1.63 g, 20 mmol) dissolved in methanol (50 mL), N,N-dimethylethylenediamine (0.882 g, 10 mmol). The mixture was stirred and refluxed gently for 3 days. Evaporating the resulting solution under reduced pressure resulted in the formation of white crystalline compound, which was then filtered, washed with Et2O, and air dried (yield: 4.35 g, 89.5%). Characterization: calcd for C20H26Br2N2O2 (MM = 486.241 g/mol): C, 49.40; H, 5.39; N, 5.76%. Found: C, 49.44; H, 5.51; N, 5.78% m.p. 189° C. IR bands (ATR, cm−1): 2972 (w), 2950 (vw), 2820 (w), ν(C-H); 1480 (s), 1458 (vs), 1446 (s), 1374 (s), 1291 (s), 1231 (s), 1204 (s) ν(C=C, C-N, C-O); 1165 (s), 1122 (s), 1047 (m), 924 (s), 902 (s), 809 (vs). ESI-MS: m/z = 487.042, calcd [H2L6 + H]+ = 487.042.

4.2.2. 6,6′-(((2-(Diethylamino)ethyl)azanediyl)bis(methylene))bis(2,4-dimethylphenol) (H2L4)

This compound was synthesized using a similar procedure and the same molar ratios as that described above for H2L3, except 2,4-dimethylphenol (2.44 g, 20 mmol) and N,N-dimethylethylenediamine (1.162 g, 10 mmol) were used instead of 2-bromo-4-methylphenol and N,N-dimethylethylenediamine, respectively (yield: 2.57 g, 66.8%). Characterization: m.p. 137–140° C. Anal. calcd for C24H36N2O2 (MM = 384.555 g/mol): C, 74.96; H, 9.44; N, 7.28%. Found: C, 74.81; H, 9.56; N, 7.28%. IR bands (ATR, cm−1): 2978 (vw), 2937 (vw), 2913 (vw), 2904 (m), 2802 (w) ν(C-H); 1484 (vs), 1376 (m), 1224 (vs) ν(C=C, C-O); 1152 (s), 1129 (m), 1096 (m), 1024 (m), 922 (m), 864 (s), 805 (m), 743 (vs). ESI-MS: m/z = 585.286 (100%), calcd [H2L2 + H]+ = 585.285. 1H NMR (d6-DMSO, 400 MHz, δ in ppm): 6.84, 6.66 (2H, 2s, 1:1, ph); 3.55 (2H, s, CH2-ph); 2.59 (2H, m, N-CH2-CH3); 2.40 (4H, m, N-CH2-CH2-N); 2.14, 2.07 (3H, 2s, 1:1, 2xCH3-ph); 0.96 (3H, t, CH3-CH2). 13C NMR: 152 (HO-Cph); 131.0, 128.2, 127,3, 125.39, 121.44 C-ph); 55.8 (N-CH2ph); 49.8, 48.7 (N-CH2-CH2-N); 45.7 (CH2CH3); 20.4 (H3C4ph); 16.1 (H3C2ph); 9.9 (CH2CH3).

4.2.3. 6,6′-(((2-(Diethylamino)ethyl)azanediyl)bis(methylene))bis(4-chloro-2-methylphenol) (H2L9)

A mixture of 4-chloro-2-methylphenol (2.825 g, 20 mmol), Et3N (2.022 g, 20 mmol), aqueous 37% HCHO (1.627 g, 20 mmol) and N,N-diethylethylenediamine (1.162 g, 10 mmol) was dissolved in methanol (60 mL). The mixture was stirred and refluxed gently for 3 days. Evaporating this solution under reduced pressure resulted in the formation of white precipitate, which was then collected by filtration, washed with Et2O, and air dried (yield: 2.35 g, 55.3%). Characterization: m.p. 137–141 °C. Cacd for C22H30Cl2N2O2 (MM = 424.168 g/mol): C, 62.12; H, 7.11; N, 6.59%. Found: C, 62.29; H, 7.21; N, 6.53%. IR bands (ATR, cm−1): 2981 (w), 2915 (vw), 2814 (w) ν(C-H); 1583 (w), 1467 (vs), 1378 (m), 1324 (m), 1273 (s), 1226 (vs) ν(C=C, C-O); 1126 (m), 1093 (s), 1063 (m), 1023 (m), 976 (m), 929 (m), 864 (vs), 804 (m) 739 (vs), 669 (m). ESI-MS: m/z = 425.175 (100%), calcd [H2L4 + H]+ = 425.176. 1H NMR (d6-DMSO, 400 MHz, δ in ppm): 7.05, 7.00 (2H, 2s, 1:1, 2xHph); 3.58 (2H, s, CH2-ph); 3.37 (2H, m, N-CH2-CH3); 2.44 (3H, t, N-CH2-CH2-N); 2.11 (3H, s, CH3-ph); 0.97 (3H, t, CH3-CH2−).

4.3. Synthesis of Copper(II) Complexes

A general method was used to synthesize the copper(II) complexes 3, 4, and 9: to a mixture containing the appropriate tripod phenolate-amine ligand (0.5 mmol) and Et3N (0.102 g, 1.00 mmol) dissolved in MeOH (20–30 mL), Cu(NO3)2·3H2O (0.122 g, 0.5 mmol) was added, and the resulting solution was heated for 5–10 min, filtered while hot through celite and allowed to crystallize at room temperature. The precipitate, which was collected by filtration, was washed with Et2O and dried in air. Crystals suitable for X-ray analysis were obtained from dilute methanolic solutions or by recrystallization from CH3CN, but acetone was used in complex 4.

4.3.1. [Cu22-L3)2] (3)

The reaction of Cu(NO3)2·3H2O with H2L3 and Et3N (1:1:2) in MeOH, as indicated above, resulted in the formation a crude product, which upon recrystallization from CH3CN, afforded large dark green crystalline compound (yield: 81.6%). Anal: calcd for C40H48Br4Cu2N4O4 (MM = 1089.90 g/mol): C, 43.85; H, 4.42; N, 5.11%. Found: C, 44.03; H, 4.47; N, 5.42%. IR bands (ATR, cm−1): 2969 (vw), 2860 (vw), 2831 (vw), 2784 (vw) ν(C-H); 1603 (w), 1464 (vs), 1378 (w), 1278 (m), 1238 (s) ν(C=C, C-O, C-N); 1125 (m), 839 (s), 795 (s), 768 (s), 594 (m), 569 (m), 493 (m), 467 (m), 411 (s). ESI-MS (CH3CN): m/z = 487.042. (100%), calcd [H2L3 + H]+ = 487.042. UV-VIS: λmax in nm (εmax, M−1cm−1): in DMSO: 452 (2330), 763 (216, b). ΛM (DMSO) = 3.0 Ω−1·cm2·mol−1.

4.3.2. [Cu22-L4)2]·CH3COCH3 (4)

This complex was isolated as big chunks of light olive-green solid (yield: 98 mg, 82.9%.). Recrystallization from acetone afforded X-ray quality crystals. Anal: calcd for C51H74Cu2N4O5 (MM = 948.425 g/mol): C, 64.46; H, 7.85; N, 5.90%. Found: C, 64.62; H, 7.88; N, 6.03%. IR bands (ATR, cm−1): 2966 (w), 2903 (w), 2840 (w) ν(C-H); 1609 (w) ν(C=N); 1479 (s), 1377 (m), 1319 (s), 1243 (s) ν(C=C, C-O); 1158 (m), 1091 (m), 1017 (m), 972 (m), 855 (vs), 794 (s), 751 (s), 729 (m). ESI-MS (CH3CN): 385.285 (100%), calcd [H2L4 + H]+ = 385.286. UV-VIS in acetone: λmax in nm (saturated solution): 409, 630 (b).

4.3.3. [Cu32-L9)2(µ-OCH3)2]·CH3OH (9)

The complex was isolated as shiny brownish-green crystalline compound atter recrystallization from acetone (yield: 69.3%). Anal: calcd for C47H66Cl4Cu3N4O6 (MM = 1131.500 g/mol): C, 49.89; H, 4.88; N, 4.95%. Found: C, 50.44; H, 5.16; N, 5.37%. IR bands (ATR, cm−1): 2976 (w), 2922 (vw), 2845 (w) ν(C-H); 1585 (w) ν(C=N); 1462 (vs), 1285 (s), 1243 (vs) ν(C=C, C-O, C-N); 1010 (m), 935 (w), 864 (s), 766 (vs), 747 (s), 661 (m), 449 (s). ESI-MS: m/z = 425.176 (100%), calcd [H2L9 + H]+ = 425.176. UV-VIS: λmax in nm (εmax, M−1cm−1): in DMSO: 447 (2650), 748 (251, b). ΛM (DMSO) = 3.5 Ω−1·cm2·mol−1.

4.4. X-ray Crystal Structure Analysis

The X-ray single-crystal data of the title compounds 3, 4, and 9 were collected on a Bruker-AXS APEX II CCD diffractometer at 100 (2) K. The crystallographic data, conditions retained for the intensity data collection and some features of the structure refinements are listed in Table 3. Data collections were performed with Mo-Kα radiation (λ= 0.71073 Å); data processing, Lorentz-polarization, and absorption corrections were performed using APEX and the SADABS computer programs [77,78,79]. The structures were solved by direct methods and refined by full-matrix least-squares methods on F2, using the SHELX [80,81,82] program library. Additional programs used: Mercury and PLATON [83]. Packing plots are given in the Supplementary Materials (Figures S1–S3).

Supplementary Materials

The following are available online at www.mdpi.com/xxx/s1.CCDC 2241168-2241170 contains the supplementary crystallographic data for 3, 4, and 9, respectively. These data can be obtained free of charge via http//www.ccdc.cam.ac.uk/conts/retrieving.html, or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: (+44) 1223-336-033; or e-mail: deposit@ccdc cam.ac.uk. Further supplementary data for Packing plots (Figures S1–S3) and Figure S4 for temperature dependence of magnetic susceptibility and magnetic moment µeff of 2. Table S1: Crystallographic data and processing parameters of 3, 4 and 9.

Author Contributions

Conceptualization, Methodology, Validation, Investigation, Resources, Data curation, Writing—original draft, Writing—review & editing, Visualization, Supervision, Project administration, Funding acquisition: S.S.M., F.A.M., M.M. and M.H.; Methodology, Validation, Formal analysis, Investigation. Writing—review & editing: F.R.L. and N.N.M.H.S.; Methodology, Software, Validation, Formal analysis, Investigation, Data curation: R.C.F. and A.T.; Methodology, Formal analysis, Investigation, Data curation, Investigation, Resources, Data curation: M.T.D. and K.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the available article and Supplementary Materials.

Acknowledgments

This research was financially supported by the Department of Chemistry—University of Louisiana at Lafayette. Our thanks also to the reviewers for their constructive comments and suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds 110 are not available.

References

  1. Massoud, S.S.; Spell, M.; Ledet, C.; Junk, T.; Herchel, R.; Fischer, R.C.; Travnicek, Z.; Mautner, F.A. Magnetic and structural properties of dinuclear singly bridged-phenoxido metal(II) complexes. Dalton Trans. 2015, 44, 2110–2121. [Google Scholar] [CrossRef]
  2. Massoud, S.S.; Junk, T.; Louka, F.R.; Herchel, R.; Travnicek, Z.; Fischer, R.C.; Mautner, F.A. Synthesis, structure and magnetic characterization of dinuclear copper(II) complexes bridged by bicompartmental phenolate. RSC Adv. 2015, 5, 87139–87150. [Google Scholar] [CrossRef]
  3. Massoud, S.S.; Junk, J.; Herchel, R.; Travnicek, Z.; Mikuriya, M.; Fischer, R.C.; Mautner, F.A. Structural characterization of ferromagnetic bridged-acetato and -dichlorido copper(II) complexes based on bicompartmental 4-t-butylphenol. Inorg. Chem. Commun. 2015, 60, 1–3. [Google Scholar] [CrossRef]
  4. Massoud, S.S.; Ledet, C.C.; Junk, T.; Bosch, S.; Comba, P.; Herchel, R.; Hošek, J.; Trávníček, Z.; Fischer, R.C.; Mautner, F.A. Dinuclear metal(II)-acetato complexes based on bicompartmental chlorophenol: Synthesis, structure, magnetic properties, DNA interaction and phosphodiester hydrolysis. Dalton Trans. 2016, 45, 12933–12950. [Google Scholar] [CrossRef] [PubMed]
  5. Mikuriya, M.; Sato, S.; Yoshioka, D. µ-Phenolato-µ-acetato-bridged dinuclear copper(II) complex with dinucleating Schiff-base ligand having three phenolate groups. X-Ray Struct. Anal. Online 2018, 34, 45–47. [Google Scholar] [CrossRef]
  6. Mikuriya, M.; Sato, S.; Yoshioka, D. µ-Phenolato-µ-benzoato-bridged dinuclear copper(II) cluster with a ferromagnetic coupling. X-Ray Struct. Anal. Online 2018, 34, 51–53. [Google Scholar] [CrossRef]
  7. Kakuta, Y.; Masuda, N.; Kurushima, M.; Hashimoto, T.; Yoshioka, D.; Sakiyama, H.; Hiraoka, Y.; Handa, M.; Mikuriya, M. Synthesis, crystal structures, spectral, electrochemical, and magnetic properties of di-µ-phenoxido-bridged dinuclear copper(II) complexes with N-salicylidene-2-hydroxybenzylamine derivatives: Axial coordination effect of dimethyl sulfoxide molecule. Chem. Pap. 2014, 68, 923–931. [Google Scholar] [CrossRef]
  8. Mikuriya, M.; Yamakawa, C.; Tanabe, K.; Nukita, R.; Amabe, Y.; Yoshioka, D.; Mitsuhashi, R.; Tatehata, R.; Tanaka, H.; Handa, M. Copper(II) carboxylates with 2,3,4-trimethoxybenzoate and 2,4,6-trimethoxybenzoate: Dinuclear Cu(II) cluster and µ-aqua-bridged Cu(II) chain molecule. Magnetochemistry 2021, 7, 35. [Google Scholar] [CrossRef]
  9. Yang, L.; Powell, D.R.; House, R.P. Structural variation in copper(I) complexes with pyridylmethylamide ligands: Structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 9, 955–964. [Google Scholar] [CrossRef]
  10. Mautner, F.A.; Koikawa, M.; Mikuriya, M.; Harrelson, E.V.; Massoud, S.S. Copper(II)-azido complexes constructed from polypyridyl amine ligands. Polyhedron 2013, 59, 17–22. [Google Scholar] [CrossRef]
  11. Mukherjee, J.; Mukherje, R. Catecholase activity of dinuclear copper(II) complexes with variable endogenous and exogenous bridge. Inorg. Chim. Acta 2002, 337, 429–438. [Google Scholar] [CrossRef]
  12. Karlin, K.D.; Farooq, A.; Hayes, J.C.; Cohen, B.I.; Rowe, T.M.; Sinn, E.; Zubieta, J. Models for met-hemocyanin derivatives: Structural and spectroscopic comparisons of analogous phenolate and X (X = OH, OMe, N3,C1, OAc, OBz) doubly bridged dinuclear copper(II) Complexes. Inorg. Chem. 1987, 26, 1271–1280. [Google Scholar] [CrossRef]
  13. Karlin, K.D.; Tyeklár, Z. (Eds.) Bioinorganic Chemistry of Copper; Chapman and Hall: New York, NY, USA, 1993. [Google Scholar]
  14. Lubben, M.; Hage, R.; Meetsma, A.; Bÿma, K.; Fering, B.L. Modeling Dinuclear Copper Sites of Biological Relevance: Synthesis, Molecular Structure, Magnetic Properties, and 1H NMR Spectroscopy of a Nonsymmetric Dinuclear Copper(II) Complex. Microcalorimetric Determination of Stepwise Complexation of Copper(II) by a Nonsymmetric Dinucleating Ligand. Inorg. Chem. 1995, 34, 2217–2224. [Google Scholar] [CrossRef]
  15. Kitajima, N.; Fujisawa, K.; Fujimoto, C.; Morooka, Y.; Hashimoto, S.; Kitagawa, T.; Toriumi, K.; Tatsumi, K.; Nakamura, A. A new model for dioxygen binding in hemocyanin. Synthesis, characterization, and molecular structure of µ-η22 peroxo dinuclear copper(II) complexes, [Cu(HB(3,5-R2pz)3)]2(O2) (R = i-Pr and Ph). J. Am. Chem. Soc. 1992, 114, 1277–1291. [Google Scholar] [CrossRef]
  16. Fujieda, N.; Umakoshi, K.; Ochi, Y.; Nishikawa, Y.; Yanagisawa, S.; Kubo, M.; Kurisu, G.; Itoh, S. Copper–Oxygen Dynamics in the Tyrosinase Mechanism. Angew. Chem. Int. Ed. 2020, 132, 13487–13492. [Google Scholar] [CrossRef]
  17. Ma, L.; Lu, L.; Zhu, M.; Wang, Q.; Gao, F.; Yuan, C.; Wu, Y.; Xing, S.; Fu, X.; Mei, Y.; et al. Dinuclear copper complexes of organic claw: Potent inhibition of protein tyrosine phosphatases. J. Inorg. Biochem. 2011, 105, 1138–1147. [Google Scholar] [CrossRef] [PubMed]
  18. Neves, A.; Rossia, L.M.; Horn, A., Jr.; Vencato, I.; Bortoluzzi, A.J.; Zucco, C.; Mangrich, A.S. Synthesis, structure and properties of the first dinuclear copper(II) complex as a structural model for the phenolic intermediate in tyrosinase–cresolase activity. Inorg. Chem. Commun. 1999, 2, 334–337. [Google Scholar] [CrossRef]
  19. Whittaker, J.W. Metal Ions in Biological Systems; Sigel, H., Sigel, A., Eds.; Marcel Dekker: New York, NY, USA, 1994; Volume 30, pp. 315–360. [Google Scholar]
  20. Thomas, F.; Gellon, G.; Gautier-Luneau, I.; Saint-Aman, E.; Pierre, J.-L. A Structural and Functional Model of Galactose Oxidase: Control of the One-Electron Oxidized Active Form through Two Differentiated Phenolic Arms in a Tripodal Ligand. Angew. Chem. Int. Ed. 2002, 41, 3047–3050. [Google Scholar] [CrossRef]
  21. Taki, M.; Kumei, H.; Nagatomo, S.; Kitagawa, T.; Itoh, S.; Fukuzumi, S. Active site models for galactose oxidase containing two different phenol groups. Inorg. Chim. Acta 2000, 300–302, 622–632. [Google Scholar] [CrossRef]
  22. Sokolowski, A.; Leutbecher, H.; Weyhermüller, T.; Schnepf, R.; Bothe, E.; Bill, E.; Hildebrandt, P.; Wieghardt, K. Phenoxyl-copper(II) complexes: Models for the active site of galactose oxidase. J. Biol. Inorg. Chem. 1997, 2, 444–453. [Google Scholar] [CrossRef]
  23. Ochs, C.; Hahn, F.E.; Fröhlich, R. Coordination Chemistry of Unsymmetrical Tripodal Ligands with NNO2 Donor Set. Eur. J. Inorg. Chem. 2001, 2427–2436. [Google Scholar] [CrossRef]
  24. Ito, N.; Phillips, S.E.V.; Yadav, K.D.S.; Knowles, P.F. Crystal Structure of a Free Radical Enzyme, Galactose Oxidase. J. Mol. Biol. 1994, 238, 794–814. [Google Scholar] [CrossRef]
  25. Mukherjee, D.; Nag, P.; Shteinman, A.A.; Vennapusa, S.R.; Mandal, U.; Mitra, M. Catechol oxidation promoted by bridging phenoxo moieties in a bis(μ-phenoxo)-bridged dicopper(II) complex. RSC Adv. 2021, 11, 22951–22959. [Google Scholar] [CrossRef]
  26. Patriarca, M.; Daier, V.; Camí, G.; Rivière, E.; Hureau, C.; Signorella, S. Preparation, characterization and activity of CuZn and Cu2 superoxide dismutase mimics encapsulated in mesoporous silica. J. Inorg. Biochem. 2020, 207, 111050. [Google Scholar] [CrossRef] [PubMed]
  27. Biswas, A.; Das, L.K.; Drew, M.G.B.; Diaz, C.; Ghosh, A. Insertion of a Hydroxido Bridge into a Diphenoxido Dinuclear Copper(II) Complex: Drastic Change of the Magnetic Property from Strong Antiferromagnetic to Ferromagnetic and Enhancement in the Catecholase Activity. Inorg. Chem. 2012, 51, 10111–10121. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, Z.; Zhang, J.; Zhang, S. Oxidative DNA cleavage promoted by two phenolate-bridged binuclear copper(II) complexes. New J. Chem. 2015, 39, 1814–1821. [Google Scholar] [CrossRef]
  29. Berthet, N.; Martel-Frachet, V.; Michel, F.; Philouze, C.; Hamman, S.; Ronot, X.; Thomas, F. Nuclease and anti-proliferative activities of copper(II) complexes of N3O tripodal ligands involving a sterically hindered phenolate. Dalton Trans. 2013, 42, 8468–8483. [Google Scholar] [CrossRef]
  30. Massoud, S.S.; Louka, F.R.; Salem, N.M.H.; Fischer, R.C.; Torvisco, A.; Mautner, F.A.; Vančo, J.; Belza, J.; Dvořák, Z.; Trávníček, Z. Dinuclear Doubly Bridged Phenoxido Copper(II) Complexes as Efficient Anticancer Agents. Eur. J. Med. Chem. 2023, 246, 114992. [Google Scholar] [CrossRef]
  31. Jahromi, Z.M.; Asadi, Z.; Eigner, V.; Dusek, M.; Rastegari, B. A new phenoxo-bridged dicopper Schiff base complex: Synthesis, crystal structure DNA/BSA interaction, cytotoxicity assay and catecholase activity. Polyhedron 2022, 221, 115891. [Google Scholar] [CrossRef]
  32. Ferraresso, L.G.; de Arruda, E.G.R.; de Moraes, T.P.L.; Fazzi, R.B.; Ferreira, A.M.D.C.; Abbehausen, C. Copper(II) and zinc(II) dinuclear enzymes model compounds: The nature of the metal ion in the biological function. J. Mol. Struct. 2017, 1150, 316–328. [Google Scholar] [CrossRef]
  33. Novoa, N.; Justaud, F.; Hamon, P.; Roisnel, T.; Cador, O.; Guennic, B.L.; Manzur, C.; Carrillo, D.; Hamon, J.-R. Doubly phenoxide-bridged binuclear copper(II) complexes with ono tridentate Schiff base ligand: Synthesis, structural, magnetic and theoretical studies. Polyhedron 2015, 86, 81–88. [Google Scholar] [CrossRef]
  34. Anbu, S.; Kandaswamy, M. Electrochemical, magnetic, catalytic, DNA binding and cleavage studies of new mono and binuclear copper(II) complexes. Polyhedron 2011, 30, 123–131. [Google Scholar] [CrossRef]
  35. Assey, G.E.; Yisgedu, T.; Gultneh, Y.; Butcher, R.J.; Tesema, Y. Di-μ-perchlorato-bis-{μ-2-[(2-pyrid-yl)methyl-amino-meth-yl] phenolato)dicopper(II) acetonitrile disolvate. Acta Crystallogr. Sect. E Struct. Rep. Online 2009, 65, m1007–m1008. [Google Scholar] [CrossRef]
  36. Assey, G.E.; Tesema, Y.; Yisgedu, T.; Gultneh, Y.; Butcher, R.J. Bis[μ-2-(2-pyridylmethyl-amino-meth-yl) phenolato]-κN,N′,O:O;κO:N,N′,O-bis-[(thio-cyanato-κN)copper(II)]. Acta Crystallogr. Sect. E Struct. Rep. Online 2009, 65, m1121–m1122. [Google Scholar] [CrossRef]
  37. Ma, J.-C.; Yang, J.; Ma, J.-F. Bis[μ-2,4-dibromo-6-(2-pyridylmethylaminomethyl) phenolato]bis[nitratocopper(II)]. Acta Crystallogr. Sect. E Struct. Rep. Online 2007, 63, m2284. [Google Scholar] [CrossRef]
  38. Choi, K.-Y. Synthesis and Crystal Structure of Phenolato-Bridged Dinuclear Copper(II) Complex with (2-Hydroxybenzyl)(2-pyridylmethyl)amine. J. Chem. Cryst. 2010, 40, 1016–1020. [Google Scholar] [CrossRef]
  39. Tandon, S.S.; Bunge, S.D.; Patel, N.; Wang, E.C.; Thompson, L.K. Self-Assembly of Antiferromagnetically-Coupled Copper(II) Supramolecular Architectures with Diverse Structural Complexities. Molecules 2020, 25, 5549. [Google Scholar] [CrossRef]
  40. Adams, H.; Bailey, N.A.; Campbell, I.K.; Fenton, D.E.; He, Q.-Y. Formation of axial phenolate–metal bonds in square-pyramidal complexes. J. Chem. Soc. Dalton Trans. 1996, 2233–2237. [Google Scholar] [CrossRef]
  41. Mallah, T.; Boillot, M.-L.; Kahn, O.; Gouteron, J.; Jeannin, S.; Jeannin, Y. Crystal Structures and Magnetic Properties of p-Phenolato Copper(I1) Binuclear Complexes with Hydroxo, Azido, and Cyanato-O Exogenous Bridges. Inorg. Chem. 1986, 25, 3058–3065. [Google Scholar] [CrossRef]
  42. You, X.; Wei, Z. Two multidentate ligands utilizing triazolyl, pyridinyl and phenolate groups as donors for constructing dinuclear copper(II) and iron(III) complexes: Syntheses, structures, and electrochemistry. Inorg. Chim. Acta 2014, 423, 332–339. [Google Scholar] [CrossRef]
  43. Rajendiran, T.M.; Kannappan, R.; Mahalakshmy, R.; Rajeswari, J.; Venkatesan, R.; Rao, P. New unsymmetrical μ-phenoxo- bridged binuclear copper(II) complexes. Transit. Met. Chem. 2003, 28, 447–454. [Google Scholar] [CrossRef]
  44. Manzur, J.; Mora, H.; Vega, A.; Spodine, E.; Venegas-Yazigi, D.; Garland, M.T.; El Fallah, M.S.; Escuer, A. Copper(II) complexes with new polypodal ligands presenting axial−equatorial phenoxo bridges {2-[(bis(2-pyridylmethyl)amino)methyl]-4-methylphenol,2-[(bis(2-pyridylmethyl)amino)methyl]-4-methyl-6-(methylthio) phenol}: Examples of ferromagnetically coupled bi- and trinuclear copper(II) complexes. Inorg. Chem. 2007, 46, 6924–6932. [Google Scholar] [CrossRef]
  45. Kani, Y.; Ohba, S.; Ito, S.; Nishida, Y. Redetermination of bis{[(2-hydroxyphenylmethyl)bis(2-pyridylmethyl)aminato]copper(II)} diperchlorate. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 2000, 56, e201. [Google Scholar] [CrossRef] [PubMed]
  46. Michel, F.; Torelli, S.; Thomas, F.; Duboc, C.; Philouze, C.; Belle, C.; Hamman, S.; Aman, E.S.; Pierre, J.-L. An Unprecedented Bridging Phenoxyl Radical in Dicopper(II) Complexes: Evidence for an S = 3/2 Spin State. Angew. Chem. Int. Ed. 2005, 44, 438–441. [Google Scholar] [CrossRef]
  47. Philibert, A.; Thomas, F.; Philouze, C.; Hamman, S.; Saint-Aman, E.; Pierre, J.-L. Galactose Oxidase Models: Tuning the Properties of CuII–Phenoxyl Radicals. Chem. Eur. J. 2003, 9, 3803–3812. [Google Scholar] [CrossRef]
  48. Wendt, F.; Rolff, M.; Thimm, W.; Näther, C.; Tuczek, F. A Small-molecule Model System of Galactose Oxidase: Geometry, Reactivity, and Electronic Structure. Z. Anorg. Allg. Chem. 2013, 639, 2502–2509. [Google Scholar] [CrossRef]
  49. Taki, M.; Hattori, H.; Osako, T.; Nagatomo, S.; Shiro, M.; Kitagawa, T.; Itoh, S. Model complexes of the active site of galactose oxidase. Effects of the metal ion binding sites. Inorg. Chim. Acta 2004, 357, 3369–3381. [Google Scholar] [CrossRef]
  50. Yajima, T.; Shimazaki, Y.; Ishigami, N.; Odani, A.; Yamauchi, O. Conformational preference of the side Conformational preference of the side chain aromatic ring in Cu(II) and Pd(II) complexes of 2N1O-donor ligands. Inorg. Chim. Acta 2002, 337, 193–202. [Google Scholar] [CrossRef]
  51. Vaidyanathan, M.; Viswanathan, R.; Palaniandavar, M.; Balasubramanian, T.; Prabhaharan, P.; Muthiah, T.P. Copper(II) Complexes with Unusual Axial Phenolate Coordination as Structural Models for the Active Site in Galactose Oxidase: X-ray Crystal Structures and Spectral and Redox Properties of [Cu(bpnp)X] Complexes. Inorg. Chem. 1998, 37, 6418–6427. [Google Scholar] [CrossRef]
  52. Banerjee, I.; Dolai, M.; Jana, A.D.; Das, K.K.; Ali, M. σ-Aromaticity in dinuclear copper(ii) complexes: Novel interaction between perchlorate anion and σ-aromatic [Cu2X2] (X = N or O) core. CrystEngComm 2012, 14, 4972–4975. [Google Scholar] [CrossRef]
  53. Thompson, L.K.; Mandal, S.K.; Tandon, S.S.; Bridson, J.N.; Park, M.K. Magnetostructural Correlations in Bis(μ2-phenoxide)-Bridged Macrocyclic Dinuclear Copper(II) Complexes. Influence of Electron-Withdrawing -Withdrawing Substituents on Exchange Coupling. Inorg. Chem. 1996, 35, 3117–3125. [Google Scholar] [CrossRef]
  54. Yadav, A.; Lama, P.; Bieńko, A.; Bieńko, D.; Siddiqui, K.A. H-bonded supramolecular synthon induced magnetic superexchange phenomenon results weak ferromagnetic and strong antiferromagnetic interactions in two new copper-orotate coordination network. Polyhedron 2018, 141, 247–261. [Google Scholar] [CrossRef]
  55. Tang, J.; Costa, J.S.; Golobič, A.; Kozlečvar, B.; Robertazzi, A.; Vargiu, A.V.; Gamez, P.; Reedijk, J. Magnetic Coupling between Copper(II) Ions Mediated by Hydrogen-Bonded (Neutral) Water Molecules. Inorg. Chem. 2009, 48, 5473–5479. [Google Scholar] [CrossRef] [PubMed]
  56. Biswas, C.; Drew, M.G.B.; Asthana, S.; Desplanches, C.; Ghosh, A. Mono-aqua-bridged dinuclear complexes of Cu(II) containing NNO donor Schiff base ligand: Hydrogen-bond-mediated exchange coupling. J. Mol. Struct. 2010, 965, 39–44. [Google Scholar] [CrossRef]
  57. Talukder, P.; Sen, S.; Mitra, S.; Dahlenberg, L.; Desplanches, C.; Sutter, J.-P. Evidence for Hydrogen-Bond-Mediated Exchange Coupling in an Aqua-Bridged CuII Dimer: Synthesis, Magnetic Study and Correlation with Density Functional Calculations. Eur. J. Inorg. Chem. 2006, 2006, 329–333. [Google Scholar] [CrossRef]
  58. Okazawa, A.; Ishida, T. Super–superexchange coupling through a hydrogen bond in a linear copper(II) complex, [Cu(LH)(L)]·BF4·2H2O (LH = N-tert-butyl-N-2-pyridylhydroxylamine). Chem. Phys. Lett. 2009, 480, 198–202. [Google Scholar] [CrossRef]
  59. Valigura, D.; Moncol, J.; Korabik, M.; Pucekova, Z.; Lis, T.; Mrozinski, J.; Melnik, M. New Dimeric Copper(II) Complex [Cu(5-MeOsal)2(μ-nia)(H2O)]2 with Magnetic Exchange Interactions through H-Bonds. Eur. J. Inorg. Chem. 2006, 3813–3817. [Google Scholar] [CrossRef]
  60. Desplanches, C.; Ruiz, E.; Rodriguez-Fortea, A.; Alvarez, S. Exchange Coupling of Transition-Metal Ions through Hydrogen Bonding: A Theoretical Investigation. J. Am. Chem. Soc. 2002, 124, 5197–5205. [Google Scholar] [CrossRef]
  61. Plass, W.; Pohlmann, A.; Rautengarten, J. Magnetic Interactions as Supramolecular Function: Structure and Magnetic Properties of Hydrogen-Bridged Dinuclear Copper(II) Complexes. Angew. Chem. Int. Ed. 2001, 40, 4207–4210. [Google Scholar] [CrossRef]
  62. Moreno, J.M.; Ruiz, J.; Dominguez-Vera, J.M.; Colacio, E. Spectroscopic and magnetic properties, and crystal structure of a dimer copper(II) complex via hydrogen bonding with a weak ferromagnetic interaction. Inorg. Chim. Acta 1993, 208, 111–115. [Google Scholar] [CrossRef]
  63. Colacio, E.; Costes, J.P.; Kivekas, R.; Laurent, J.P.; Ruiz, J.; Sundberg, M. A new hydrogen-bonded dinuclear complex involving copper(II) ions in a pseudotetrahedral N3O environment: Molecular and crystal structure and magnetic and spectroscopic properties. Inorg. Chem. 1991, 30, 1475–1479. [Google Scholar] [CrossRef]
  64. Oberhausen, K.J.; Richardson, J.F.; Buchanan, R.M.; McCusker, J.K.; Hendrickson, D.N.; Latour, J.M. Synthesis and characterization of dinuclear copper(II) complexes of the dinucleating ligand 2,6-bis[(bis((1-methylimidazol-2-yl)methyl)amino)methyl]-4-methylphenol. Inorg. Chem. 1991, 30, 1357–1365. [Google Scholar] [CrossRef]
  65. Mautner, F.A.; Fischer, R.C.; Rashmawi, L.G.; Louka, F.R.; Massoud, S.S. Structural characterization of metal(II) thiocyanato complexes derived from bis(2-(H-pyrazol-1-yl)ethyl)amine. Polyhedron 2017, 124, 237–242. [Google Scholar] [CrossRef]
  66. Massoud, S.S.; Louka, R.F.; David, R.N.; Dartez, M.J.; Nguyn, Q.L.; Labry, N.J.; Fischer, R.C.; Mautner, F.A. Five-coordinate metal(II) complexes based pyrazolyl ligands. Polyhedron 2015, 90, 258–265. [Google Scholar] [CrossRef]
  67. Geary, W.J. The use of conductivity measurements in organic solvents for the characterisation of coordination compounds. Coord. Chem. Rev. 1971, 7, 81–122. [Google Scholar] [CrossRef]
  68. Addison, A.W.; Rao, T.N.; Reedijk, J.; Rijin, J.V.; Verschoor, G.C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc. Dalton Trans. 1984, 1349–1356. [Google Scholar] [CrossRef]
  69. Archana, V.; Imamura, Y.; Sakiyama, H.; Hada, M. Correlating Magnetic Exchange in Dinuclear Bis(phenolate)-Bridged Com-plexes: A Computational Perspective. Bull. Chem. Soc. Jpn. (BCSJ) 2016, 89, 657–665. [Google Scholar] [CrossRef]
  70. Venegas-Yazigi, D.; Aravena, D.; Spodine, E.; Ruiz, E.; Alvarez, S. Structural and electronic effects on the exchange interactions in dinuclear bis(phenoxo)-bridged copper(II) complexes. Coord. Chem. Rev. 2010, 254, 2086–2095. [Google Scholar] [CrossRef]
  71. Mondal, D.; Majee, M.C.; Bhattacharya, K.; Long, J.; Larionova, J.; Khusniyarov, M.M.; Chaudhury, M. Crossover from Antiferromagnetic to Ferromagnetic Exchange Coupling in a New Family of Bis-(μ-phenoxido)dicopper(II) Complexes: A Comprehensive Magneto-Structural Correlation by Experimental and Theoretical Study. ACS Omega. 2019, 4, 10558–10570. [Google Scholar] [CrossRef]
  72. Mondal, D.; Majee, M.C.; Bhattacharya, K.; Long, J.; Larionova, J.; Khusniyarov, M.M.; Chaudhury, M. Synthesis and characterization of binuclear [ONOX]-type amine-bis(phenolate) copper(II) complexes. Inorg. Chim. Acta 2011, 375, 158–165. [Google Scholar]
  73. Chaudhuri, P.; Wagner, R.; Weyhermüller, T. Ferromagnetic vs. antiferromagnetic coupling in bis(μ-phenoxo)dicopper(II) complexes. Tuning of the nature of exchange coupling by remote ligand substituents. Inorg. Chem. 2007, 46, 5134–5136. [Google Scholar] [CrossRef] [PubMed]
  74. Van Crawford, H.; Richardson, H.W.; Wasson, J.R.; Hodgson, D.J.; Hatfield, W.E. Relation between the singlet-triplet splitting and the copper-oxygen-copper bridge angle in hydroxo-bridged copper dimers. Inorg. Chem. 1976, 15, 2107–2110. [Google Scholar] [CrossRef]
  75. Merz, L.; Hasse, W. Exchange interaction in tetrameric oxygen-bridged copper clusters of the cubane type. Dalton Trans. 1980, 875–879. [Google Scholar] [CrossRef]
  76. Kahn, O. Molecular Magnetism; VCH: Cambridge, UK, 1993; Chapter 1. [Google Scholar]
  77. Bruker. SAINT, version 7.23; Bruker: Billerica, MA, USA, 2005. [Google Scholar]
  78. Bruker. APEX 2, version 2.0-2; Bruker: Billerica, MA, USA, 2006. [Google Scholar]
  79. Sheldrick, G.M. SADABS, version 2; University of Goettingen: Goettingen, Germany, 2001. [Google Scholar]
  80. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  81. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–5. [Google Scholar] [CrossRef] [PubMed]
  82. Macrae, C.F.; Edington, P.R.; McCabe, P.; Pidcock, E.; Shields, G.P.; Taylor, R.; Towler, T.; van de Streek, J.J. Mercury: Visualization and analysis of crystal structures. J. Appl. Cryst. 2006, 39, 453–457. [Google Scholar] [CrossRef]
  83. Speck, A.L. PLATON, a Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2001. [Google Scholar]
Scheme 1. Phenolate ligands used in this study and their abbreviations.
Scheme 1. Phenolate ligands used in this study and their abbreviations.
Molecules 28 02648 sch001
Scheme 2. Synthesis of the tripodal phenolate amine ligands and the doubly-bridged phenoxido-Cu(II) coordination core of complexes derived from N,N-dialkylethylenediamine backbone.
Scheme 2. Synthesis of the tripodal phenolate amine ligands and the doubly-bridged phenoxido-Cu(II) coordination core of complexes derived from N,N-dialkylethylenediamine backbone.
Molecules 28 02648 sch002
Figure 1. Perspective views of (a) 3, (b) 4, and (c) trinuclear subunit of 9.
Figure 1. Perspective views of (a) 3, (b) 4, and (c) trinuclear subunit of 9.
Molecules 28 02648 g001
Figure 2. Perspective views of 1, 2, 58, and 10 represented as (ag), respectively.
Figure 2. Perspective views of 1, 2, 58, and 10 represented as (ag), respectively.
Molecules 28 02648 g002
Figure 3. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu2(L1)2] (1).
Figure 3. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu2(L1)2] (1).
Molecules 28 02648 g003
Figure 4. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu2(L3)2] (3).
Figure 4. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu2(L3)2] (3).
Molecules 28 02648 g004
Figure 5. Temperature dependence of magnetic susceptibility χM () and magnetic moment µeff () of [Cu2(L5)2(H2O)]·2H2O (5).
Figure 5. Temperature dependence of magnetic susceptibility χM () and magnetic moment µeff () of [Cu2(L5)2(H2O)]·2H2O (5).
Molecules 28 02648 g005
Figure 6. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu2(L7)2](ClO4)2 (7).
Figure 6. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu2(L7)2](ClO4)2 (7).
Molecules 28 02648 g006
Figure 7. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu3(L9)2(µ-OCH3)2] (9).
Figure 7. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu3(L9)2(µ-OCH3)2] (9).
Molecules 28 02648 g007
Figure 8. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu(L10)] (10).
Figure 8. Temperature dependence of magnetic susceptibility χM (Δ) and magnetic moment µeff () of [Cu(L10)] (10).
Molecules 28 02648 g008
Table 1. Selected bond distances (Å) and angles (°) of 3, 4, and 9.
Table 1. Selected bond distances (Å) and angles (°) of 3, 4, and 9.
3
Cu1-N12.089(10)Cu2-N32.160(10)
Cu1-N22.325(10)Cu2-N42.180(10)
Cu1-O11.892(8)Cu2-O31.905(8)
Cu1-O21.963(9)Cu2-O22.002(8)
Cu1-O42.013(8)Cu2-O41.999(8)
Cu1-O2-Cu2100.4(4)O2-Cu1-O475.4(3)
Cu1-O4-Cu298.8(4)O2-Cu2-O474.8(4)
O1-Cu1-O2157.0(4)O3-Cu2-O4164.3(4)
O4-Cu1-N1160.4(4)O2-Cu2-N4142.0(4)
4
Cu1-N12.0627(19)Cu2-N32.090(2)
Cu1-N22.449(2)Cu2-N42.3532(19)
Cu1-O11.8783(17)Cu2-O31.8805(17)
Cu1-O21.9580(17)Cu2-O22.0349(16)
Cu1-O42.0218(16)Cu2-O41.9465(16)
Cu1-O2-Cu297.17(7)O2-Cu1-O475.78(7)
Cu1-O4-Cu297.98(7)O2-Cu2-O475.73(7)
O1-Cu1-O2160.59(7)O3-Cu2-O4165.55(7)
O4-Cu1-N1156.88(7)O2-Cu2-N4138.62(7)
9
Cu1-N12.054(2)Cu3-N32.066(2)
Cu1-N22.351(2)Cu3-N42.324(2)
Cu1-O11.929(2)Cu3-O41.941(2)
Cu1-O22.000(2)Cu3-O51.985(2)
Cu1-O31.957(2)Cu3-O61.944(2)
Cu2-O21.9488(19)Cu2-O31.999(2)
Cu2-O41.920(2)Cu2-O51.959(2)
Cu1-O2-Cu297.33(9)O2-Cu1-O375.43(8)
Cu1-O3-Cu2100.19(9)O2-Cu2-O377.73(8)
Cu2-O4-Cu399.27(9)O4-Cu2-O577.91(8)
Cu2-O5-Cu396.48(9)O4-Cu3-O576.79(8)
O2-Cu2-O4102.80(8)O3-Cu2-O5103.85(9)
O1-Cu1-O2161.11(8)O5-Cu3-O6153.32(9)
O3-Cu1-N1161.70(8)O4-Cu3-N3167.39(9)
Table 2. Magnetostructural correlation of Cu2O2-phenoxido complexes.
Table 2. Magnetostructural correlation of Cu2O2-phenoxido complexes.
ComplexJ (cm−1)C.N. (Geom., τ5
or τ4)
Cu…Cu
(Å)
Cu-O-Cu’
(°)
Cu-O-Cu-O
(°)
Cu-O-C-C
(°)
τ (°)
1−2895 (SP, 0.23)3.092103.4415.952.17.8
2−2025 (SP, 0.24)3.00699.1224.957.33.6
3−1455 (SP, 0.06/0.37)3.04798.8, 100.424.850.9, 71.010.5, 12.2
4−1465(SP, 0.07/0.45)2.99597.17, 97.7826.952.5, 67.53.2, 12.8
5−1764(SQP, 0.17)
5(SP, 0.011)
2.933196.70, 97.5829.450.5, 49.92.7, 5.5
6−2504(SQP, 0.17)
5(SP, 0.013)
2.94197.56, 96.9329.849.9, 51.76.5, 4.6
7−3.65(SP, 0.41) + 13.13097.72011.331.3
8−4.65(SP, 0.40) + 13.09896.63011.330.9
9−594(SQP, 0.17)
5(SP: 0.01, 0.23)
2.9652 2.941896.48, 97.33, 99.27, 100.1922.9, 23.544.2, 41.98.2, 20.0
10+3.54(SQP, 0.23)7.269
Table 3. Crystallographic data and processing parameters of 3, 4, and 9.
Table 3. Crystallographic data and processing parameters of 3, 4, and 9.
Compound349
Empirical formulaC40H48Br4Cu2N4O4C51H74Cu2N4O5C47H62Cl4Cu3N4O7
Formula mass1095.53950.221131.45
SystemMonoclinicTriclinicMonoclinic
Space groupP21/cP-1P21/n
a (Å)15.433(3)12.6538(12)10.7647(5)
b (Å)14.198(3)13.0192(12)29.4374(14)
c (Å)20.973(4)15.4646(15)16.5922(8)
α (°)9095.347(5)90
β (°)100.460(7)101.633(5)90.673(2)
γ (°)90103.880(4)90
V (Å3)4519.2(16)2395.5(4)5275.5(4)
Z424
θ max (°)26.00030.55824.999
Data collected82442133738159804
Unique refl. 8840 14144 9251
Parameters 495573 600
Goodness-of-fit on F21.0691.0571.109
R1/wR2 (all data)0.1102/0.28450.0528/0.13360.0357/0.0923
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Massoud, S.S.; Louka, F.R.; Dial, M.T.; Salem, N.N.M.H.; Fischer, R.C.; Torvisco, A.; Mautner, F.A.; Nakashima, K.; Handa, M.; Mikuriya, M. Magnetostructural Properties of Some Doubly-Bridged Phenoxido Copper(II) Complexes. Molecules 2023, 28, 2648. https://doi.org/10.3390/molecules28062648

AMA Style

Massoud SS, Louka FR, Dial MT, Salem NNMH, Fischer RC, Torvisco A, Mautner FA, Nakashima K, Handa M, Mikuriya M. Magnetostructural Properties of Some Doubly-Bridged Phenoxido Copper(II) Complexes. Molecules. 2023; 28(6):2648. https://doi.org/10.3390/molecules28062648

Chicago/Turabian Style

Massoud, Salah S., Febee R. Louka, Madison T. Dial, Nahed N. M. H. Salem, Roland C. Fischer, Ana Torvisco, Franz A. Mautner, Kai Nakashima, Makoto Handa, and Masahiro Mikuriya. 2023. "Magnetostructural Properties of Some Doubly-Bridged Phenoxido Copper(II) Complexes" Molecules 28, no. 6: 2648. https://doi.org/10.3390/molecules28062648

Article Metrics

Back to TopTop