29 research outputs found

    Organoarsonate Functionalization of Heteropolyoxotungstates

    No full text
    Functionalization of the {P<sub>8</sub>W<sub>48</sub>} polyoxotungstate (POT) archetype with aromatic organoarsonates results in the first homometallic {P<sub>8</sub>W<sub>48</sub>} derivatives, with the general formula [(RAs<sup>V</sup>O)<sub>4</sub>P<sup>V</sup><sub>8</sub>W<sup>VI</sup><sub>48</sub>O<sub>184</sub>]<sup>32−</sup> [R = C<sub>6</sub>H<sub>5</sub> (<b>1</b>) or <i>p</i>-(H<sub>2</sub>N)­C<sub>6</sub>H<sub>4</sub> (<b>2</b>)]. Short As−O bonds here induce unusual bending of the otherwise rigid {P<sub>8</sub>W<sub>48</sub>} macrocycle, breaking its <i>D</i><sub>4<i>h</i></sub> symmetry. The obtained species also represent the first lacunary POTs functionalized with organoarsonates and can potentially act as polyoxometalate precursors themselves. We elaborate solution stability in different aqueous media using <sup>1</sup>H and <sup>31</sup>P NMR spectroscopy and possible pathways for subsequent transformations in aqueous solutions of the functionalized polyanions. Recrystallization of the K<sup>+</sup>/Li<sup>+</sup>/dimethylammonium salt of <b>2</b> from 4 M LiCl solution yielded a further functionalized POT, [(H<sub>3</sub>NC<sub>6</sub>H<sub>4</sub>AsO)<sub>3</sub>P<sub>8</sub>W<sub>48</sub>O<sub>184</sub>H<sub><i>x</i></sub>{WO<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>}<sub>0.4</sub>]<sup>(30.2−<i>x</i>)−</sup> (<b>3</b>), revealing dissociation of the organoarsonate fragments in slightly acidic aqueous solutions followed by their rearrangement within the inner POT cavity

    Expansion of Antimonato Polyoxovanadates with Transition Metal Complexes: (Co(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>)<sub>2</sub>[{Co(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]·5H<sub>2</sub>O and (Ni(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>)<sub>2</sub>[{Ni(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]·8H<sub>2</sub>O

    No full text
    Two new polyoxovanadates (Co­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>)<sub>2</sub>[{Co­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}­V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]·5H<sub>2</sub>O (<b>1</b>) and (Ni­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>)<sub>2</sub>[{Ni­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}­V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]·8H<sub>2</sub>O (<b>2</b>) (N<sub>3</sub>C<sub>5</sub>H<sub>15</sub> = <i>N</i>-(2-aminoethyl)-1,3-propanediamine) were synthesized under solvothermal conditions and structurally characterized. In both structures the [V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]<sup>6–</sup> shell displays the main structural motif, which is strongly related to the {V<sub>18</sub>O<sub>42</sub>} archetype cluster. Both compounds crystallize in the triclinic space group <i>P</i>1̅ with <i>a</i> = 14.3438(4), <i>b</i> = 16.6471(6), <i>c</i> = 18.9186(6) Å, α = 87.291(3)°, β = 83.340(3)°, γ = 78.890(3)°, and <i>V</i> = 4401.4(2) Å<sup>3</sup> (<b>1</b>) and <i>a</i> = 14.5697(13), <i>b</i> = 15.8523(16), <i>c</i> = 20.2411(18) Å, α = 86.702(11)°, β = 84.957(11)°, γ = 76.941(11)°, and <i>V</i> = 4533.0(7) Å<sup>3</sup> (<b>2</b>). In the structure of <b>1</b> the [V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]<sup>6–</sup> cluster anion is bound to a [Co­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>]<sup>2+</sup> complex via a terminal oxygen atom. In the Co<sup>2+</sup>-centered complex, one of the amine ligands coordinates in tridentate mode and the second one in bidentate mode to form a strongly distorted CoN<sub>5</sub>O octahedron. Similarly, in compound <b>2</b> an analogous NiN<sub>5</sub>O complex is joined to the [V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]<sup>6–</sup> anion via the same attachment mode. A remarkable difference between the two compounds is the orientation of the noncoordinated propylamine group leading to intermolecular Sb···O contacts in <b>1</b> and to Sb···N interactions in <b>2</b>. In the solid-state lattices of <b>1</b> and <b>2</b>, two additional [M­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>]<sup>2+</sup> complexes act as countercations and are located between the [{M­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}­V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]<sup>4–</sup> anions. Between the anions and cations strong N–H···O hydrogen bonds are observed. In both compounds the clusters are stacked along the <i>b</i> axis in an ABAB fashion with cations and water molecules occupying the space between the clusters. Magnetic characterization demonstrates that the Ni<sup>2+</sup> and Co<sup>2+</sup> cations do not significantly couple with the <i>S</i> = 1/2 vanadyl groups. The susceptibility data can be successfully reproduced assuming a distorted ligand field for the Co<sup>2+</sup> ions (<b>1</b>) and an <i>O</i><sub><i>h</i></sub>-symmetric Ni<sup>2+</sup> ligand field (<b>2</b>)

    Chiral Hexanuclear Ferric Wheels

    No full text
    The homochiral iron­(III) wheels [Fe<sub>6</sub>{(<i>S</i>)-pedea}<sub>6</sub>Cl<sub>6</sub>] and [Fe<sub>6</sub>{(<i>R</i>)-pedea)}<sub>6</sub>Cl<sub>6</sub>] [(<i>R</i>)- and (<i>S</i>)-<b>2</b>; pedea = phenylethylaminodiethoxide] exhibit high optical activities and antiferromagnetic exchange. These homochiral products react with each other, producing the centrosymmetric, crystallographically characterized [Fe<sub>6</sub>{(<i>S</i>)-pedea}<sub>3</sub>{(<i>R</i>)-pedea}<sub>3</sub>Cl<sub>6</sub>] diastereomer [(<i>RSRSRS</i>)-<b>2</b>]. <sup>1</sup>H NMR and UV–vis studies indicate that exchange processes are slow in both homo- and heterochiral systems but that, upon combination, the reaction between (<i>R</i>)- and (<i>S</i>)-<b>2</b> occurs quickly

    Expansion of Antimonato Polyoxovanadates with Transition Metal Complexes: (Co(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>)<sub>2</sub>[{Co(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]·5H<sub>2</sub>O and (Ni(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>)<sub>2</sub>[{Ni(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]·8H<sub>2</sub>O

    No full text
    Two new polyoxovanadates (Co­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>)<sub>2</sub>[{Co­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}­V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]·5H<sub>2</sub>O (<b>1</b>) and (Ni­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>)<sub>2</sub>[{Ni­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}­V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]·8H<sub>2</sub>O (<b>2</b>) (N<sub>3</sub>C<sub>5</sub>H<sub>15</sub> = <i>N</i>-(2-aminoethyl)-1,3-propanediamine) were synthesized under solvothermal conditions and structurally characterized. In both structures the [V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]<sup>6–</sup> shell displays the main structural motif, which is strongly related to the {V<sub>18</sub>O<sub>42</sub>} archetype cluster. Both compounds crystallize in the triclinic space group <i>P</i>1̅ with <i>a</i> = 14.3438(4), <i>b</i> = 16.6471(6), <i>c</i> = 18.9186(6) Å, α = 87.291(3)°, β = 83.340(3)°, γ = 78.890(3)°, and <i>V</i> = 4401.4(2) Å<sup>3</sup> (<b>1</b>) and <i>a</i> = 14.5697(13), <i>b</i> = 15.8523(16), <i>c</i> = 20.2411(18) Å, α = 86.702(11)°, β = 84.957(11)°, γ = 76.941(11)°, and <i>V</i> = 4533.0(7) Å<sup>3</sup> (<b>2</b>). In the structure of <b>1</b> the [V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]<sup>6–</sup> cluster anion is bound to a [Co­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>]<sup>2+</sup> complex via a terminal oxygen atom. In the Co<sup>2+</sup>-centered complex, one of the amine ligands coordinates in tridentate mode and the second one in bidentate mode to form a strongly distorted CoN<sub>5</sub>O octahedron. Similarly, in compound <b>2</b> an analogous NiN<sub>5</sub>O complex is joined to the [V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]<sup>6–</sup> anion via the same attachment mode. A remarkable difference between the two compounds is the orientation of the noncoordinated propylamine group leading to intermolecular Sb···O contacts in <b>1</b> and to Sb···N interactions in <b>2</b>. In the solid-state lattices of <b>1</b> and <b>2</b>, two additional [M­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>]<sup>2+</sup> complexes act as countercations and are located between the [{M­(N<sub>3</sub>C<sub>5</sub>H<sub>15</sub>)<sub>2</sub>}­V<sub>15</sub>Sb<sub>6</sub>O<sub>42</sub>(H<sub>2</sub>O)]<sup>4–</sup> anions. Between the anions and cations strong N–H···O hydrogen bonds are observed. In both compounds the clusters are stacked along the <i>b</i> axis in an ABAB fashion with cations and water molecules occupying the space between the clusters. Magnetic characterization demonstrates that the Ni<sup>2+</sup> and Co<sup>2+</sup> cations do not significantly couple with the <i>S</i> = 1/2 vanadyl groups. The susceptibility data can be successfully reproduced assuming a distorted ligand field for the Co<sup>2+</sup> ions (<b>1</b>) and an <i>O</i><sub><i>h</i></sub>-symmetric Ni<sup>2+</sup> ligand field (<b>2</b>)

    Tetrapalladium-Containing Polyoxotungstate [Pd<sup>II</sup><sub>4</sub>(α‑P<sub>2</sub>W<sub>15</sub>O<sub>56</sub>)<sub>2</sub>]<sup>16–</sup>: A Comparative Study

    No full text
    The novel tetrapalladium­(II)-containing polyoxometalate [Pd<sup>II</sup><sub>4</sub>(<i>α-</i>P<sub>2</sub>W<sub>15</sub>O<sub>56</sub>)<sub>2</sub>]<sup>16–</sup> has been prepared in aqueous medium and characterized as its hydrated sodium salt Na<sub>16</sub>[Pd<sub>4</sub>(α-P<sub>2</sub>W<sub>15</sub>O<sub>56</sub>)<sub>2</sub>]·71H<sub>2</sub>O by single-crystal XRD, elemental analysis, IR, Raman, multinuclear NMR, and UV–vis spectroscopy. The complex exists in anti and syn conformations, which form in a 2:1 ratio, and possesses unique structural characteristics in comparison with known {M<sub>4</sub>(P<sub>2</sub>W<sub>15</sub>)<sub>2</sub>} species. <sup>31</sup>P and <sup>183</sup>W NMR spectroscopy are consistent with the long-term stability of the both isomers in aqueous solutions

    Tuning the Condensation Degree of {Fe<sup>III</sup><sub><i>n</i></sub>} Oxo Clusters via Ligand Metathesis, Temperature, and Solvents

    No full text
    Trinuclear μ<sub>3</sub>-oxo-centered iron­(III) isobutyrate clusters readily react with polyalcohol organic ligands under one-pot synthesis conditions. Depending on the ligand, solvent, and temperature, a range of hexa-, dodeca-, and doicosanuclear iron­(III) oxo-hydroxo condensation products, isolated as (mdeaH<sub>3</sub>)<sub>2</sub>[Fe<sub>6</sub>O­(thme)<sub>4</sub>Cl<sub>6</sub>]·0.5­(MeCN)·0.5­(H<sub>2</sub>O) (<b>1</b>), [Fe<sub>12</sub>O<sub>4</sub>(OH)<sub>2</sub>(teda)<sub>4</sub>(N<sub>3</sub>)<sub>4</sub>(MeO)<sub>4</sub>]­N<sub>3</sub>(NO<sub>3</sub>)<sub>0.5</sub>(MeO)<sub>0.5</sub>·2.5­(H<sub>2</sub>O) (<b>2</b>), [Fe<sub>12</sub>O<sub>6</sub>(teda)<sub>4</sub>Cl<sub>8</sub>]·6­(CHCl<sub>3</sub>) (<b>3</b>), [Fe<sub>22</sub>O<sub>16</sub>(OH)<sub>2</sub>(O<sub>2</sub>CCHMe<sub>2</sub>)<sub>18</sub>(bdea)<sub>6</sub>(EtO)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]·2­(EtOH)·5­(MeCN)·6­(H<sub>2</sub>O) (<b>4</b>), and [Fe<sub>22</sub>O<sub>14</sub>(OH)<sub>4</sub>(O<sub>2</sub>CCHMe<sub>2</sub>)<sub>18</sub>(mdea)<sub>6</sub>(EtO)<sub>2</sub>(H<sub>2</sub>O)<sub>2</sub>]­(NO<sub>3</sub>)<sub>2</sub>·EtOH·H<sub>2</sub>O (<b>5</b>), where tedaH<sub>4</sub> = <i>N</i>,<i>N</i>,<i>N</i>′,<i>N</i>′-tetrakis­(2-hydroxyethyl)­ethylenediamine; thmeH<sub>3</sub> = 1,1,1-tris­(hydroxymethyl)­ethane; mdeaH<sub>2</sub> = <i>N</i>-methyldiethanolamine; and bdeaH<sub>2</sub> = <i>N</i>-butyldiethanolamine. Complete carboxylate metathesis in the {Fe<sub>3</sub>} precursor complexes by thme<sup>3–</sup> or teda<sup>4–</sup> and the agglomeration of the formed species under solvothermal conditions afforded carboxylate-free {Fe<sub>6</sub>} product (<b>1</b>) in MeCN/CH<sub>2</sub>Cl<sub>2</sub> or {Fe<sub>12</sub>} complexes (<b>2</b> and <b>3</b>) in MeOH/EtOH and CHCl<sub>3</sub>/thf, respectively (thf = tetrahydrofuran). Single-crystal X-ray diffraction analyses revealed that <b>1</b> contains a [Fe<sub>6</sub>O­(thme)<sub>4</sub>Cl<sub>6</sub>]<sup>2–</sup> cluster anion with a Lindqvist-type {Fe<sub>6</sub>(μ<sub>6</sub>-O)} core motif, charge-compensated by two protonated mdeaH<sub>3</sub><sup>+</sup> cations. <b>2</b> comprises a [Fe<sub>12</sub>O<sub>4</sub>(OH)<sub>2</sub>(teda)<sub>4</sub>­(N<sub>3</sub>)<sub>4</sub>(MeO)<sub>4</sub>]<sup>2+</sup> cation with a {Fe<sub>12</sub>O<sub>4</sub>(OH)<sub>2</sub>}<sup>26+</sup> core, whereas <b>3</b> contains a charge-neutral [Fe<sub>12</sub>O<sub>6</sub>(teda)<sub>4</sub>(Cl)<sub>8</sub>] complex with an {Fe<sub>12</sub>O<sub>6</sub>}<sup>24+</sup> core. Finally, employing flexible bdeaH<sub>2</sub> or mdeaH<sub>2</sub> ligands under soft reaction conditions afforded giant {Fe<sub>22</sub>} oxo-hydroxo complexes (<b>4</b> and <b>5</b>) with a central {Fe<sub>6</sub>} layer sandwiched between two outer {Fe<sub>8</sub>} groups. Magnetic studies of <b>1</b>–<b>5</b> revealed strong antiferromagnetic coupling between the Fe<sup>III</sup> spin centers in all clusters

    Synthesis, Structure, and Magnetic Properties of a New Family of Tetra-nuclear {Mn<sub>2</sub><sup>III</sup>Ln<sub>2</sub>}(Ln = Dy, Gd, Tb, Ho) Clusters With an Arch-Type Topology: Single-Molecule Magnetism Behavior in the Dysprosium and Terbium Analogues

    No full text
    Sequential reaction of Mn­(II) and lanthanide­(III) salts with a new multidentate ligand, 2,2′-(2-hydroxy-3-methoxy-5-methylbenzylazanediyl)­diethanol (<b>LH</b><sub><b>3</b></sub>), containing two flexible ethanolic arms, one phenolic oxygen, and a methoxy group afforded heterometallic tetranuclear complexes [Mn<sub>2</sub>Dy<sub>2</sub>(LH)<sub>4</sub>(μ-OAc)<sub>2</sub>]­(NO<sub>3</sub>)<sub>2</sub>·2CH<sub>3</sub>OH·3H<sub>2</sub>O (<b>1</b>), [Mn<sub>2</sub>Gd<sub>2</sub>(LH)<sub>4</sub>(μ-OAc)<sub>2</sub>]­(NO<sub>3</sub>)<sub>2</sub>·2CH<sub>3</sub>OH·3H<sub>2</sub>O (<b>2</b>), [Mn<sub>2</sub>Tb<sub>2</sub>(LH)<sub>4</sub>(μ-OAc)<sub>2</sub>]­(NO<sub>3</sub>)<sub>2</sub>·2H<sub>2</sub>O·2CH<sub>3</sub>OH·Et<sub>2</sub>O (<b>3</b>), and [Mn<sub>2</sub>Ho<sub>2</sub>(LH)<sub>4</sub>(μ-OAc)<sub>2</sub>]­Cl<sub>2</sub>·5CH<sub>3</sub>OH (<b>4</b>). All of these dicationic complexes possess an arch-like structural topology containing a central Mn<sup>III</sup>–Ln–Ln–Mn<sup>III</sup> core. The two central lanthanide ions are connected via two phenolate oxygen atoms. The remaining ligand manifold assists in linking the central lanthanide ions with the peripheral Mn­(III) ions. Four doubly deprotonated LH<sup>2–</sup> chelating ligands are involved in stabilizing the tetranuclear assembly. A magnetochemical analysis reveals that single-ion effects dominate the observed susceptibility data for all compounds, with comparably weak Ln···Ln and very weak Ln···Mn­(III) couplings. The axial, approximately square-antiprismatic coordination environment of the Ln<sup>3+</sup> ions in <b>1</b>–<b>4</b> causes pronounced zero-field splitting for Tb<sup>3+</sup>, Dy<sup>3+</sup>, and Ho<sup>3+</sup>. For <b>1</b> and <b>3</b>, the onset of a slowing down of the magnetic relaxation was observed at temperatures below approximately 5 K (<b>1</b>) and 13 K (<b>3</b>) in frequency-dependent alternating current (AC) susceptibility measurements, yielding effective relaxation energy barriers of Δ<i>E</i> = 16.8 cm<sup>–1</sup> (<b>1</b>) and 33.8 cm<sup>–1</sup> (<b>3</b>)

    Synthesis, Structure, and Magnetic Properties of a New Family of Tetra-nuclear {Mn<sub>2</sub><sup>III</sup>Ln<sub>2</sub>}(Ln = Dy, Gd, Tb, Ho) Clusters With an Arch-Type Topology: Single-Molecule Magnetism Behavior in the Dysprosium and Terbium Analogues

    No full text
    Sequential reaction of Mn­(II) and lanthanide­(III) salts with a new multidentate ligand, 2,2′-(2-hydroxy-3-methoxy-5-methylbenzylazanediyl)­diethanol (<b>LH</b><sub><b>3</b></sub>), containing two flexible ethanolic arms, one phenolic oxygen, and a methoxy group afforded heterometallic tetranuclear complexes [Mn<sub>2</sub>Dy<sub>2</sub>(LH)<sub>4</sub>(μ-OAc)<sub>2</sub>]­(NO<sub>3</sub>)<sub>2</sub>·2CH<sub>3</sub>OH·3H<sub>2</sub>O (<b>1</b>), [Mn<sub>2</sub>Gd<sub>2</sub>(LH)<sub>4</sub>(μ-OAc)<sub>2</sub>]­(NO<sub>3</sub>)<sub>2</sub>·2CH<sub>3</sub>OH·3H<sub>2</sub>O (<b>2</b>), [Mn<sub>2</sub>Tb<sub>2</sub>(LH)<sub>4</sub>(μ-OAc)<sub>2</sub>]­(NO<sub>3</sub>)<sub>2</sub>·2H<sub>2</sub>O·2CH<sub>3</sub>OH·Et<sub>2</sub>O (<b>3</b>), and [Mn<sub>2</sub>Ho<sub>2</sub>(LH)<sub>4</sub>(μ-OAc)<sub>2</sub>]­Cl<sub>2</sub>·5CH<sub>3</sub>OH (<b>4</b>). All of these dicationic complexes possess an arch-like structural topology containing a central Mn<sup>III</sup>–Ln–Ln–Mn<sup>III</sup> core. The two central lanthanide ions are connected via two phenolate oxygen atoms. The remaining ligand manifold assists in linking the central lanthanide ions with the peripheral Mn­(III) ions. Four doubly deprotonated LH<sup>2–</sup> chelating ligands are involved in stabilizing the tetranuclear assembly. A magnetochemical analysis reveals that single-ion effects dominate the observed susceptibility data for all compounds, with comparably weak Ln···Ln and very weak Ln···Mn­(III) couplings. The axial, approximately square-antiprismatic coordination environment of the Ln<sup>3+</sup> ions in <b>1</b>–<b>4</b> causes pronounced zero-field splitting for Tb<sup>3+</sup>, Dy<sup>3+</sup>, and Ho<sup>3+</sup>. For <b>1</b> and <b>3</b>, the onset of a slowing down of the magnetic relaxation was observed at temperatures below approximately 5 K (<b>1</b>) and 13 K (<b>3</b>) in frequency-dependent alternating current (AC) susceptibility measurements, yielding effective relaxation energy barriers of Δ<i>E</i> = 16.8 cm<sup>–1</sup> (<b>1</b>) and 33.8 cm<sup>–1</sup> (<b>3</b>)

    Ultralarge 3d/4f Coordination Wheels: From Carboxylate/Amino Alcohol-Supported {Fe<sub>4</sub>Ln<sub>2</sub>} to {Fe<sub>18</sub>Ln<sub>6</sub>} Rings

    No full text
    A family of wheel-shaped charge-neutral heterometallic {Fe<sup>III</sup><sub>4</sub>Ln<sup>III</sup><sub>2</sub>}- and {Fe<sup>III</sup><sub>18</sub>M<sup>III</sup><sub>6</sub>}-type coordination clusters demonstrates the intricate interplay of solvent effects and structure-directing roles of semiflexible bridging ligands. The {Fe<sub>4</sub>Ln<sub>2</sub>}-type compounds [Fe<sub>4</sub>Ln<sub>2</sub>(O<sub>2</sub>CCMe<sub>3</sub>)<sub>6</sub>­(N<sub>3</sub>)<sub>4</sub>(Htea)<sub>4</sub>]·2­(EtOH), Ln = Dy (<b>1a</b>), Er (<b>1b</b>), Ho (<b>1c</b>); [Fe<sub>4</sub>Tb<sub>2</sub>(O<sub>2</sub>CCMe<sub>3</sub>)<sub>6</sub>­(N<sub>3</sub>)<sub>4</sub>(Htea)<sub>4</sub>] (<b>1d</b>); [Fe<sub>4</sub>Ln<sub>2</sub>(O<sub>2</sub>CCMe<sub>3</sub>)<sub>6</sub>­(N<sub>3</sub>)<sub>4</sub>(Htea)<sub>4</sub>]·2­(CH<sub>2</sub>Cl<sub>2</sub>), Ln = Dy (<b>2a</b>), Er (<b>2b</b>); [Fe<sub>4</sub>Ln<sub>2</sub>(O<sub>2</sub>CCMe<sub>3</sub>)<sub>4</sub>­(N<sub>3</sub>)<sub>6</sub>(Htea)<sub>4</sub>]·2­(EtOH)·2­(CH<sub>2</sub>Cl<sub>2</sub>), Ln = Dy (<b>3a</b>), Er (<b>3b</b>) and the {Fe<sub>18</sub>M<sub>6</sub>}-type compounds [Fe<sub>18</sub>M<sub>6</sub>(O<sub>2</sub>CCHMe<sub>2</sub>)<sub>12</sub>­(Htea)<sub>18</sub>(tea)<sub>6</sub>(N<sub>3</sub>)<sub>6</sub>]·<i>n</i>(solvent), M = Dy (<b>4</b>, <b>4a</b>), Gd (<b>5</b>), Tb (<b>6</b>), Ho (<b>7</b>), Sm (<b>8</b>), Eu (<b>9</b>), and Y (<b>10</b>) form in ca. 20–40% yields in direct reaction of trinuclear Fe<sup>III</sup> pivalate or isobutyrate clusters, lanthanide/yttrium nitrates, and bridging triethanolamine (H<sub>3</sub>tea) and azide ligands in different solvents: EtOH for the smaller {Fe<sub>4</sub>Ln<sub>2</sub>} wheels and MeOH/MeCN or MeOH/EtOH for the larger {Fe<sub>18</sub>M<sub>6</sub>} wheels. Single-crystal X-ray diffraction analyses revealed that <b>1</b>–<b>3</b> consist of planar centrosymmetric hexanuclear clusters built from Fe<sup>III</sup> and Ln<sup>III</sup> ions linked by an array of bridging carboxylate, azide, and aminopolyalcoholato-based ligands into a cyclic structure with a cavity, and with distinct sets of crystal solvents (2 EtOH per formula unit in <b>1a</b>–<b>c</b>, 2 CH<sub>2</sub>Cl<sub>2</sub> in <b>2</b>, and 2 EtOH and 2 CH<sub>2</sub>Cl<sub>2</sub> in <b>3</b>). In <b>4</b>–<b>10</b>, the largest 3d/4f wheels currently known, nearly linear Fe<sub>3</sub> fragments are joined via mononuclear Ln/Y units by a set of isobutyrates and amino alcohol ligands into virtually planar rings. The magnetic properties of <b>1</b>–<b>10</b> reveal slow magnetization relaxation for {Fe<sub>4</sub>Tb<sub>2</sub>} (<b>1d</b>) and slow relaxation for {Fe<sub>4</sub>Ho<sub>2</sub>} (<b>1c</b>), {Fe<sub>18</sub>Dy<sub>6</sub>} (<b>4</b>), and {Fe<sub>18</sub>Tb<sub>6</sub>} (<b>6</b>)

    Assembly of Cerium(III) 2,2′-Bipyridine-5,5′-dicarboxylate-based Metal–Organic Frameworks by Solvent Tuning

    No full text
    Small changes to the reaction conditions differentiate between two metal–organic frameworks (MOFs), {[Ce<sub>2</sub>(H<sub>2</sub>O)­(bpdc)<sub>3</sub>(dmf)<sub>2</sub>]·2­(dmf)}<sub><i>n</i></sub> (<b>1</b>) and {[Ce<sub>4</sub>(H<sub>2</sub>O)<sub>5</sub>(bpdc)<sub>6</sub>(dmf)]·<i>x</i>(dmf)}<sub><i>n</i></sub> (<b>2</b>), that were solvothermally synthesized from cerium­(III) nitrate hexahydrate and 2,2′-bipyridine-5,5′-dicarboxylic acid (H<sub>2</sub>bpdc) in dimethylformamide (dmf). The two compounds illustrate how the flexibility of the coordination geometry of Ce<sup>III</sup> translates into MOFs, the formation of which readily adapts to different solvent environments
    corecore