16 research outputs found

    Toward an Understanding of the Ambiguous Electron Paramagnetic Resonance Spectra of the Iminoxy Radical from <i>o</i>ā€‘Fluorobenzaldehyde Oxime: Density Functional Theory and <i>ab Initio</i> Studies

    No full text
    Iminoxy radicals (R<sub>1</sub>R<sub>2</sub>Cī—»Nī—øO<sup>ā€¢</sup>) possess an inherent ability to exist as <i>E</i> and <i>Z</i> isomers. Although isotropic hyperfine couplings for the species with R<sub>1</sub> = H allow one to distinguish between <i>E</i> and <i>Z</i>, unequivocal assignment of the parameters observed in the EPR spectra of the radicals without the hydrogen atom at the azomethine carbon to the right isomer is not a simple task. The iminoxyl derived from <i>o</i>-fluoroacetophenone oxime (R<sub>1</sub> = CH<sub>3</sub> and R<sub>2</sub> = <i>o</i>-FC<sub>6</sub>H<sub>5</sub>) appears to be a case in point. Moreover, for its two isomers the rotation of the <i>o</i>-FC<sub>6</sub>H<sub>5</sub> group brings into existence the <i>syn</i> and <i>anti</i> conformers, depending on the mutual orientation of the F atom and Cī—»Nī—øO<sup>ā€¢</sup> group, making a description of hyperfine couplings to structure even more challenging. To accomplish this, a vast array of theoretical methods (DFT, OO-SCS-MP2, QCISD) was used to calculate the isotropic hyperfine couplings. The comparison between experimental and theoretical values revealed that the <i>E</i> isomer is the dominant radical form, for which a fast interconversion between <i>anti</i> and <i>syn</i> conformers is expected. In addition, the origin of the significant <i>A</i><sub>F</sub> increase with solvent polarity was analyzed

    Oxidation of 1ā€‘Methyl-1-phenylhydrazine with Oxidovanadium(V)ā€“Salan Complexes: Insight into the Pathway to the Formation of Hydrazine by Vanadium Nitrogenase

    No full text
    A series of oxidovanadiumĀ­(V) complexes [VOĀ­(L-Īŗ<sup>4</sup>O,N,N,O)Ā­(OR)] (<b>1a</b>, R = Et, L = L<sup>1</sup>; <b>1b</b>, R = Me, L = L<sup>1</sup>; <b>2</b>, R = Me, L = L<sup>2</sup>; <b>3</b>, R = Me, L = L<sup>3</sup>) were synthesized by the Ļƒ-bond metathesis reaction between [VOĀ­(OR)<sub>3</sub>] and the linear diaminebisĀ­(phenol) derivatives H<sub>2</sub>L (salans) containing different para-substituents on the phenoxo group [CMe<sub>3</sub>CH<sub>2</sub>CMe<sub>2</sub>, L<sup>1</sup>; Me, L<sup>2</sup>; Cl, L<sup>3</sup>]. As shown by X-ray crystallography complexes <b>1a</b>, <b>1b</b>, and <b>2</b> exhibit cis-Ī± geometry, do have a stereogenic vanadium center, and exist as a racemic mixture of the Ī” cis-Ī± and Ī› cis-Ī± enantiomers. In solution, as demonstrated by <sup>1</sup>H and <sup>51</sup>V NMR investigations, the structures of complexes <b>1</b>ā€“<b>3</b> are consistent with their solid state. The reactions of <b>1b</b>, <b>2</b>, and <b>3</b> with NH<sub>2</sub>NMePh in <i>n</i>-hexane afforded the oxidovanadiumĀ­(IV) [VOĀ­(L-Īŗ<sup>4</sup>O,N,N,O)] (<b>4</b>, L<sup>1</sup>; <b>5</b>, L<sup>2</sup>; <b>6</b>, L<sup>3</sup>) and 1,4-dimethyl-1,4-diphenyl-2-tetrazene (Me<sub>2</sub>Ph<sub>2</sub>N<sub>4</sub>) (<b>7</b>) as the main products together with a small amount of hydrazidoĀ­(2-) vanadiumĀ­(V) [VĀ­(L<sup>3</sup>-Īŗ<sup>4</sup>O,N,N,O)Ā­(NNMePh)Ā­(OMe)] (<b>8</b>). Proposed reaction course for the oxidation of NH<sub>2</sub>NMePh by <b>1b</b>ā€“<b>3</b> is discussed. Compounds <b>4</b>ā€“<b>8</b> were characterized by chemical and physical techniques including the X-ray crystallography for <b>4</b>, <b>7</b>, and <b>8</b>. The solid-state (powder) electron paramagnetic resonance spectra and magnetic features strongly indicate that complexes <b>4</b>ā€“<b>6</b> are formed as a mixture of a mononuclear (<i>S</i> = 1/2) and a dinuclear (<i>S</i> = 1) species

    Can Carbamates Undergo Radical Oxidation in the Soil Environment? A Case Study on Carbaryl and Carbofuran

    No full text
    Radical oxidation of carbamate insecticides, namely carbaryl and carbofuran, was investigated with spectroscopic (electron paramagnetic resonance [EPR] and UVā€“vis) and theoretical (density functional theory [DFT] and ab initio orbital-optimized spin-component scaled MP2 [OO-SCS-MP2]) methods. The two carbamates were subjected to reaction with <sup>ā€¢</sup>OH, persistent DPPH<sup>ā€¢</sup> and galvinoxyl radical, as well as indigenous radicals of humic acids. The influence of fulvic acids on carbamate oxidation was also tested. The results obtained with EPR and UVā€“vis spectroscopy indicate that carbamates can undergo direct reactions with various radical species, oxidizing themselves into radicals in the process. Hence, they are prone to participate in the prolongation step of the radical chain reactions occurring in the soil environment. Theoretical calculations revealed that from the thermodynamic point of view hydrogen atom transfer is the preferred mechanism in the reactions of the two carbamates with the radicals. The activity of carbofuran was determined experimentally (using pseudo-first-order kinetics) and theoretically to be noticeably higher in comparison with carbaryl and comparable with gallic acid. The findings of this study suggest that the radicals present in soil can play an important role in natural remediation mechanisms of carbamates

    Copper(II) Carboxylate Dimers Prepared from Ligands Designed to Form a Robust Ļ€Ā·Ā·Ā·Ļ€ Stacking Synthon: Supramolecular Structures and Molecular Properties

    No full text
    The reactions of bifunctional carboxylate ligands (1,8-naphthalimido)Ā­propanoate, (<b>L</b><sub><b>C2</b></sub><sup><b>ā€“</b></sup>), (1,8-naphthalimido)Ā­ethanoate, (<b>L</b><sub><b>C1</b></sub><sup><b>ā€“</b></sup>), and (1,8-naphthalimido)Ā­benzoate, (<b>L<sub>C4</sub><sup>ā€“</sup>)</b> with Cu<sub>2</sub>(O<sub>2</sub>CCH<sub>3</sub>)<sub>4</sub>(H<sub>2</sub>O)<sub>2</sub> in methanol or ethanol at room temperature lead to the formation of novel dimeric [Cu<sub>2</sub>(<b>L</b><sub><b>C2</b></sub>)<sub>4</sub>(MeOH)<sub>2</sub>] (<b>1</b>), [Cu<sub>2</sub>(<b>L</b><sub><b>C1</b></sub>)<sub>4</sub>(MeOH)<sub>2</sub>]Ā·2Ā­(CH<sub>2</sub>Cl<sub>2</sub>) (<b>2</b>), [Cu<sub>2</sub>(<b>L</b><sub><b>C4</b></sub>)<sub>4</sub>(EtOH)<sub>2</sub>]Ā·2Ā­(CH<sub>2</sub>Cl<sub>2</sub>) (<b>3</b>) complexes. When the reaction of <b>L</b><sub><b>C1</b></sub><sup><b>ā€“</b></sup> with Cu<sub>2</sub>(O<sub>2</sub>CCH<sub>3</sub>)<sub>4</sub>(H<sub>2</sub>O)<sub>2</sub> was carried out at āˆ’20 Ā°C in the presence of pyridine, [Cu<sub>2</sub>(<b>L</b><sub><b>C1</b></sub>)<sub>4</sub>(py)<sub>4</sub>]Ā·2Ā­(CH<sub>2</sub>Cl<sub>2</sub>) (<b>4</b>) was produced. At the core of complexes <b>1</b>ā€“<b>3</b> lies the square Cu<sub>2</sub>(O<sub>2</sub>CR)<sub>4</sub> ā€œpaddlewheelā€ secondary building unit, where the two copper centers have a nearly square pyramidal geometry with methanol or ethanol occupying the axial coordination sites. Complex <b>4</b> contains a different type of dimeric core generated by two Īŗ<sup>1</sup>-bridging carboxylate ligands. Additionally, two terminal carboxylates and four trans situated pyridine molecules complete the coordination environment of the five-coordinate copperĀ­(II) centers. In all four compounds, robust Ļ€Ā·Ā·Ā·Ļ€ stacking interactions of the naphthalimide rings organize the dimeric units into two-dimensional sheets. These two-dimensional networks are organized into a three-dimensional architecture by two different noncovalent interactions: strong Ļ€Ā·Ā·Ā·Ļ€ stacking of the naphthalimide rings (also the pyridine rings for <b>4</b>) in <b>1</b>, <b>3</b>, and <b>4</b>, and intermolecular hydrogen bonding of the coordinated methanol or ethanol molecules in <b>1</b>ā€“<b>3</b>. Magnetic measurements show that the copper ions in the paddlewheel complexes <b>1</b>ā€“<b>3</b> are strongly antiferromagnetically coupled with ā€“<i>J</i> values ranging from 255 to 325 cm<sup>ā€“1</sup>, whereas the copper ions in <b>4</b> are only weakly antiferromagnetically coupled. Typical values of the zero-field splitting parameter <i>D</i> were found from EPR studies of <b>1</b>ā€“<b>3</b> and the related known complexes [Cu<sub>2</sub>(<b>L</b><sub><b>C2</b></sub>)<sub>4</sub>(py)<sub>2</sub>]<b>Ā·</b>2Ā­(CH<sub>2</sub>Cl<sub>2</sub>)<b>Ā·</b>(CH<sub>3</sub>OH), [Cu<sub>2</sub>(<b>L</b><sub><b>C3</b></sub>)<sub>4</sub>(py)<sub>2</sub>]<b>Ā·</b>2Ā­(CH<sub>2</sub>Cl<sub>2</sub>) and [Cu<sub>2</sub>(<b>L</b><sub><b>C3</b></sub>)<sub>4</sub>(bipy)]<b>Ā·</b>(CH<sub>3</sub>OH)<sub>2</sub><b>Ā·</b>(CH<sub>2</sub>Cl<sub>2</sub>)<sub>3.37</sub> (<b>L</b><sub><b>C3</b></sub><sup><b>ā€“</b></sup> = (1,8-naphthalimido)Ā­butanoate)), while its abnormal magnitude in [Cu<sub>2</sub>(<b>L</b><sub><b>C2</b></sub>)<sub>4</sub>(bipy)] was qualitatively rationalized by structural analysis and DFT calculations

    Dinuclear Complexes Containing Linear Mā€“Fā€“M [M = Mn(II), Fe(II), Co(II), Ni(II), Cu(II), Zn(II), Cd(II)] Bridges: Trends in Structures, Antiferromagnetic Superexchange Interactions, and Spectroscopic Properties

    No full text
    The reaction of MĀ­(BF<sub>4</sub>)<sub>2</sub>Ā·<i>x</i>H<sub>2</sub>O, where M is FeĀ­(II), CoĀ­(II), NiĀ­(II), CuĀ­(II), ZnĀ­(II), and CdĀ­(II), with the new ditopic ligand <i>m</i>-bisĀ­[bisĀ­(3,5-dimethyl-1-pyrazolyl)Ā­methyl]Ā­benzene (<b>L<sub><i>m</i></sub>*</b>) leads to the formation of monofluoride-bridged dinuclear metallacycles of the formula [M<sub>2</sub>(Ī¼-F)Ā­(Ī¼-<b>L<sub><i>m</i></sub>*</b>)<sub>2</sub>]Ā­(BF<sub>4</sub>)<sub>3</sub>. The analogous manganeseĀ­(II) species, [Mn<sub>2</sub>(Ī¼-F)Ā­(Ī¼-<b>L<sub><i>m</i></sub>*</b>)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub>, was isolated starting with MnĀ­(ClO<sub>4</sub>)<sub>2</sub>Ā·6H<sub>2</sub>O using NaBF<sub>4</sub> as the source of the bridging fluoride. In all of these complexes, the geometry around the metal centers is trigonal bipyramidal, and the fluoride bridges are linear. The <sup>1</sup>H, <sup>13</sup>C, and <sup>19</sup>F NMR spectra of the zincĀ­(II) and cadmiumĀ­(II) compounds and the <sup>113</sup>Cd NMR of the cadmiumĀ­(II) compound indicate that the metallacycles retain their structure in acetonitrile and acetone solution. The compounds with M = MnĀ­(II), FeĀ­(II), CoĀ­(II), NiĀ­(II), and CuĀ­(II) are antiferromagnetically coupled, although the magnitude of the coupling increases dramatically with the metal as one moves to the right across the periodic table: MnĀ­(II) (āˆ’6.7 cm<sup>ā€“1</sup>) < FeĀ­(II) (āˆ’16.3 cm<sup>ā€“1</sup>) < CoĀ­(II) (āˆ’24.1 cm<sup>ā€“1</sup>) < NiĀ­(II) (āˆ’39.0 cm<sup>ā€“1</sup>) ā‰Ŗ CuĀ­(II) (āˆ’322 cm<sup>ā€“1</sup>). High-field EPR spectra of the copperĀ­(II) complexes were interpreted using the coupled-spin Hamiltonian with <i>g</i><sub><i>x</i></sub> = 2.150, <i>g</i><sub><i>y</i></sub> = 2.329, <i>g</i><sub><i>z</i></sub> = 2.010, <i>D</i> = 0.173 cm<sup>ā€“1</sup>, and <i>E</i> = 0.089 cm<sup>ā€“1</sup>. Interpretation of the EPR spectra of the ironĀ­(II) and manganeseĀ­(II) complexes required the spin Hamiltonian using the noncoupled spin operators of two metal ions. The values <i>g</i><sub><i>x</i></sub> = 2.26, <i>g</i><sub><i>y</i></sub> = 2.29, <i>g</i><sub><i>z</i></sub> = 1.99, <i>J</i> = āˆ’16.0 cm<sup>ā€“1</sup>, <i>D</i><sub>1</sub> = āˆ’9.89 cm<sup>ā€“1</sup>, and <i>D</i><sub>12</sub> = āˆ’0.065 cm<sup>ā€“1</sup> were obtained for the ironĀ­(II) complex and <i>g</i><sub><i>x</i></sub> = <i>g</i><sub><i>y</i></sub> = <i>g</i><sub><i>z</i></sub> = 2.00, <i>D</i><sub>1</sub> = āˆ’0.3254 cm<sup>ā€“1</sup>, <i>E</i><sub>1</sub> = āˆ’0.0153, <i>J</i> = āˆ’6.7 cm<sup>ā€“1</sup>, and <i>D</i><sub>12</sub> = 0.0302 cm<sup>ā€“1</sup> were found for the manganeseĀ­(II) complex. Density functional theory (DFT) calculations of the exchange integrals and the zero-field splitting on manganeseĀ­(II) and ironĀ­(II) ions were performed using the hybrid B3LYP functional in association with the TZVPP basis set, resulting in reasonable agreement with experiment

    Halide and Hydroxide Linearly Bridged Bimetallic Copper(II) Complexes: Trends in Strong Antiferromagnetic Superexchange Interactions

    No full text
    Centrosymmetric [Cu<sub>2</sub>(Ī¼-X)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub>*)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub> (X = F<sup>ā€“</sup>, Cl<sup>ā€“</sup>, Br<sup>ā€“</sup>, OH<sup>ā€“</sup>, <b>L</b><sub><i><b>m</b></i></sub>* = <i>m</i>-bisĀ­[bisĀ­(3,5-dimethyl-1-pyrazolyl)Ā­methyl]Ā­benzene)], the first example of a series of bimetallic copperĀ­(II) complexes linked by a linearly bridging mononuclear anion, have been prepared and structurally characterized. Very strong antiferromagnetic exchange coupling between the copperĀ­(II) ions increases along the series F<sup>ā€“</sup> < Cl<sup>ā€“</sup> < OH<sup>ā€“</sup> < Br<sup>ā€“</sup>, where āˆ’<i>J</i> = 340, 720, 808, and 945 cm<sup>ā€“1</sup>. DFT calculations explain this trend by an increase in the energy along this series of the antibonding antisymmetric combination of the p orbital of the bridging anion interacting with the copperĀ­(II) d<sub><i>z</i><sup>2</sup></sub> orbital

    Halide and Hydroxide Linearly Bridged Bimetallic Copper(II) Complexes: Trends in Strong Antiferromagnetic Superexchange Interactions

    No full text
    Centrosymmetric [Cu<sub>2</sub>(Ī¼-X)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub>*)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub> (X = F<sup>ā€“</sup>, Cl<sup>ā€“</sup>, Br<sup>ā€“</sup>, OH<sup>ā€“</sup>, <b>L</b><sub><i><b>m</b></i></sub>* = <i>m</i>-bisĀ­[bisĀ­(3,5-dimethyl-1-pyrazolyl)Ā­methyl]Ā­benzene)], the first example of a series of bimetallic copperĀ­(II) complexes linked by a linearly bridging mononuclear anion, have been prepared and structurally characterized. Very strong antiferromagnetic exchange coupling between the copperĀ­(II) ions increases along the series F<sup>ā€“</sup> < Cl<sup>ā€“</sup> < OH<sup>ā€“</sup> < Br<sup>ā€“</sup>, where āˆ’<i>J</i> = 340, 720, 808, and 945 cm<sup>ā€“1</sup>. DFT calculations explain this trend by an increase in the energy along this series of the antibonding antisymmetric combination of the p orbital of the bridging anion interacting with the copperĀ­(II) d<sub><i>z</i><sup>2</sup></sub> orbital

    Syntheses, Structural, Magnetic, and Electron Paramagnetic Resonance Studies of Monobridged Cyanide and Azide Dinuclear Copper(II) Complexes: Antiferromagnetic Superexchange Interactions

    No full text
    The reactions of CuĀ­(ClO<sub>4</sub>)<sub>2</sub> with NaCN and the ditopic ligands <i>m</i>-bisĀ­[bisĀ­(1-pyrazolyl)Ā­methyl]Ā­benzene (<b>L</b><sub><i><b>m</b></i></sub>) or <i>m</i>-bisĀ­[bisĀ­(3,5-dimethyl-1-pyrazolyl)Ā­methyl]Ā­benzene (<b>L</b><sub><i><b>m</b></i></sub>*) yield [Cu<sub>2</sub>(Ī¼-CN)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub>)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub> (<b>1</b>) and [Cu<sub>2</sub>(Ī¼-CN)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub><b>*</b>)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub> (<b>3</b>). In both, the cyanide ligand is linearly bridged (Ī¼-1,2) leading to a separation of the two copperĀ­(II) ions of ca. 5 ƅ. The geometry around copperĀ­(II) in these complexes is distorted trigonal bipyramidal with the cyanide group in an equatorial position. The reaction of [Cu<sub>2</sub>(Ī¼-F)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub>)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub> and (CH<sub>3</sub>)<sub>3</sub>SiN<sub>3</sub> yields [Cu<sub>2</sub>(<i>Ī¼-</i>N<sub>3</sub>)Ā­(<i>Ī¼-</i><b>L</b><sub><i><b>m</b></i></sub>)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub> (<b>2</b>), where the azide adopts end-on (Ī¼-1,1) coordination with a Cuā€“Nā€“Cu angle of 138.0Ā° and a distorted square pyramidal geometry about the copperĀ­(II) ions. Similar chemistry in the more sterically hindered <b>L</b><sub><i><b>m</b></i></sub>* system yielded only the coordination polymer [Cu<sub>2</sub>(<i>Ī¼-</i><b>L</b><sub><i><b>m</b></i></sub>*)Ā­(<i>Ī¼-</i>N<sub>3</sub>)<sub>2</sub>Ā­(N<sub>3</sub>)<sub>2</sub>]. Attempts to prepare a dinuclear complex with a bridging iodide yield the copperĀ­(I) complex [Cu<sub>5</sub>(<i>Ī¼-</i>I<sub>4</sub>)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub>*)<sub>2</sub>]Ā­I<sub>3</sub>. The complexes <b>1</b> and <b>3</b> show strong antiferromagnetic coupling, āˆ’<i>J</i> = 135 and 161 cm<sup>ā€“1</sup>, respectively. Electron paramagnetic resonance (EPR) studies coupled with density functional theory (DFT) calculations show that the exchange interaction is transmitted through the d<sub><i>z</i><sup>2</sup></sub> and the bridging ligand s and p<sub><i>x</i></sub> orbitals. High field EPR studies confirmed the d<sub><i>z</i><sup>2</sup></sub> ground state of the copperĀ­(II) ions. Single-crystal high-field EPR has been able to definitively show that the signs of <i>D</i> and <i>E</i> are positive. The zero-field splitting is dominated by the anisotropic exchange interactions. Complex <b>2</b> has āˆ’<i>J</i> = 223 cm<sup>ā€“1</sup> and DFT calculations indicate a predominantly d<sub><i>x</i><sup>2</sup>ā€“y<sup>2</sup></sub> ground state

    Hydroxide-Bridged Cubane Complexes of Nickel(II) and Cadmium(II): Magnetic, EPR, and Unusual Dynamic Properties

    No full text
    The reactions of MĀ­(ClO<sub>4</sub>)<sub>2</sub>Ā·<i>x</i>H<sub>2</sub>O (M = NiĀ­(II) or CdĀ­(II)) and <i>m</i>-bisĀ­[bisĀ­(1-pyrazolyl)Ā­methyl]Ā­benzene (<b>L</b><sub><b>m</b></sub>) in the presence of triethylamine lead to the formation of hydroxide-bridged cubane compounds of the formula [M<sub>4</sub>(Ī¼<sub>3</sub>-OH)<sub>4</sub>(Ī¼-<b>L</b><sub><b>m</b></sub>)<sub>2</sub>(solvent)<sub>4</sub>]Ā­(ClO<sub>4</sub>)<sub>4</sub>, where solvent = dimethylformamide, water, acetone. In the solid state the metal centers are in an octahedral coordination environment, two sites are occupied by pyrazolyl nitrogens from <b>L</b><sub><b>m</b></sub>, three sites are occupied by bridging hydroxides, and one site contains a weakly coordinated solvent molecule. A series of multinuclear, two-dimensional and variable-temperature NMR experiments showed that the cadmiumĀ­(II) compound in acetonitrile-<i>d</i><sub>3</sub> has <i>C</i><sub>2</sub> symmetry and undergoes an unusual dynamic process at higher temperatures (Ī”<i>G</i><sub>Lm</sub><sup>ā€”</sup> = 15.8 Ā± 0.8 kcal/mol at 25 Ā°C) that equilibrates the pyrazolyl rings, the hydroxide hydrogens, and cadmiumĀ­(II) centers. The proposed mechanism for this process combines two motions in the semirigid <b>L</b><sub><b>m</b></sub> ligand termed the ā€œColumbia Twist and Flip:ā€ twisting of the pyrazolyl rings along the C<sub>pz</sub>ā€“C<sub>methine</sub> bond and 180Ā° ring flip of the phenylene spacer along the C<sub>Ph</sub>ā€“C<sub>methine</sub> bond. This dynamic process was also followed using the spin saturation method, as was the exchange of the hydroxide hydrogens with the trace water present in acetonitrile-<i>d</i><sub>3</sub>. The nickelĀ­(II) analogue, as shown by magnetic susceptibility and electron paramagnetic resonance measurements, has an <i>S</i> = 4 ground state, and the nickelĀ­(II) centers are ferromagnetically coupled with strongly nonaxial zero-field splitting parameters. Depending on the Niā€“Oā€“Ni angles two types of interactions are observed: <i>J</i><sub>1</sub> = 9.1 cm<sup>ā€“1</sup> (97.9 to 99.5Ā°) and <i>J</i><sub>2</sub> = 2.1 cm<sup>ā€“1</sup> (from 100.3 to 101.5Ā°). ā€œBroken symmetryā€ density functional theory calculations performed on a model of the nickelĀ­(II) compound support these observations

    Dinuclear Metallacycles with Single Mā€“O(H)ā€“M Bridges [M = Fe(II), Co(II), Ni(II), Cu(II)]: Effects of Large Bridging Angles on Structure and Antiferromagnetic Superexchange Interactions

    No full text
    The reactions of MĀ­(ClO<sub>4</sub>)<sub>2</sub>Ā·<i>x</i>H<sub>2</sub>O and the ditopic ligands <i>m</i>-bisĀ­[bisĀ­(1-pyrazolyl)Ā­methyl]Ā­benzene (<b>L</b><sub><i><b>m</b></i></sub>) or <i>m</i>-bisĀ­[bisĀ­(3,5-dimethyl-1-pyrazolyl)Ā­methyl]Ā­benzene (<b>L</b><sub><i><b>m</b></i></sub>*) in the presence of triethylamine lead to the formation of monohydroxide-bridged, dinuclear metallacycles of the formula [M<sub>2</sub>(Ī¼-OH)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub>)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub> (M = FeĀ­(II), CoĀ­(II), CuĀ­(II)) or [M<sub>2</sub>(Ī¼-OH)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub>*)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub> (M = CoĀ­(II), NiĀ­(II), CuĀ­(II)). With the exception of the complexes where the ligand is <b>L</b><sub><i><b>m</b></i></sub> and the metal is copperĀ­(II), all of these complexes have distorted trigonal bipyramidal geometry around the metal centers and unusual linear (<b>L</b><sub><i><b>m</b></i></sub>*) or nearly linear (<b>L</b><sub><i><b>m</b></i></sub>) Mā€“Oā€“M angles. For the two solvates of [Cu<sub>2</sub>(Ī¼-OH)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub>)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub>, the Cuā€“Oā€“Cu angles are significantly bent and the geometry about the metal is distorted square pyramidal. All of the copperĀ­(II) complexes have structural distortions expected for the pseudo-Jahnā€“Teller effect. The two cobaltĀ­(II) complexes show moderate antiferromagnetic coupling, āˆ’<i>J</i> = 48ā€“56 cm<sup>ā€“1</sup>, whereas the copperĀ­(II) complexes show very strong antiferromagnetic coupling, āˆ’<i>J</i> = 555ā€“808 cm<sup>ā€“1</sup>. The largest coupling is observed for [Cu<sub>2</sub>(Ī¼-OH)Ā­(Ī¼-<b>L</b><sub><i><b>m</b></i></sub>*)<sub>2</sub>]Ā­(ClO<sub>4</sub>)<sub>3</sub>, the complex with a Cuā€“Oā€“Cu angle of 180Ā°, such that the exchange interaction is transmitted through the d<sub><i>z</i><sup>2</sup></sub> and the oxygen s and p<sub><i>x</i></sub> orbitals. The interaction decreases, but it is still significant, as the Cuā€“Oā€“Cu angle decreases and the character of the metal orbital becomes increasingly d<sub><i>x</i><sup>2</sup>ā€“<i>y</i><sup>2</sup></sub>. These intermediate geometries and magnetic interactions lead to spin Hamiltonian parameters for the copperĀ­(II) complexes in the EPR spectra that have large <i>E</i>/<i>D</i> ratios and one <i>g</i> matrix component very close to 2. Density functional theory calculations were performed using the hybrid B3LYP functional in association with the TZVPP basis set, resulting in reasonable agreement with the experiments
    corecore