11 research outputs found

    Formation of Specific Configurations at Stereogenic Nitrogen Centers upon Their Coordination to Zinc and Mercury

    No full text
    The coordination of (<i>R</i>,<i>R</i>)-tetramethylcyclohexane-1,2-diamine derivatives with stereogenic nitrogen centers to zinc and mercury halides is investigated. It is shown that the resulting complexes display one specific configuration at the stereogenic nitrogen centers. This fact is unusual due to the fast inversion of nitrogen centers but highly desirable as the stereoinformation of the ligands is brought closer to the metal centers of the potential catalysts. A combination of NMR studies and quantum chemical calculations gives insight into the selective formation of one specific configuration at the stereogenic nitrogen centers of the zinc complexes

    Formation of Specific Configurations at Stereogenic Nitrogen Centers upon Their Coordination to Zinc and Mercury

    No full text
    The coordination of (<i>R</i>,<i>R</i>)-tetramethylcyclohexane-1,2-diamine derivatives with stereogenic nitrogen centers to zinc and mercury halides is investigated. It is shown that the resulting complexes display one specific configuration at the stereogenic nitrogen centers. This fact is unusual due to the fast inversion of nitrogen centers but highly desirable as the stereoinformation of the ligands is brought closer to the metal centers of the potential catalysts. A combination of NMR studies and quantum chemical calculations gives insight into the selective formation of one specific configuration at the stereogenic nitrogen centers of the zinc complexes

    Experimental and Theoretical Studies on the Mechanism of the C–S Bond Activation of Pd<sup>II</sup> Thiolate/Thioether Complexes

    No full text
    Two equivalents of <b>L</b> (<b>L</b> = 4-methylthio-2-thioxo-1,3-dithiole-5-thiolate or Medmit) react with <i>cis</i>-Pd­(PR<sub>3</sub>)<sub>2</sub>Cl<sub>2</sub> (R = Ph and Et) to afford Pd­(PR<sub>3</sub>)­(η<sup>1</sup>-<b>L</b>)­(η<sup>2</sup>-<b>L</b>) (R = Et: <b>1</b> ; R = Ph: <b>2</b>) complexes, which have been characterized by X-ray crystallography. These compounds are dynamic in solution due to an exchange of the thiomethyl groups on palladium. Variable-temperature <sup>1</sup>H NMR spectroscopy reveals a low coalescence temperature (173 K). Treatment of Pd­(<i>diphos</i>)­Cl<sub>2</sub> (<i>diphos</i> = dppe or dppm) with 2 equiv of <b>L</b> affords thiolato complexes Pd­(dppe)­(η<sup>1</sup>-<b>L</b>)<sub>2</sub> (<b>3</b>) and Pd­(dppm)­(η<sup>1</sup>-<b>L</b>)<sub>2</sub> (<b>4</b>). Whereas <b>3</b> is rigid in solution with firm η<sup>2</sup>-coordination of dppe and η<sup>1</sup>-coordination of the thiolate, two linkage isomers Pd­(η<sup>2</sup>-dppm)­(η<sup>1</sup>-<b>L</b>)<sub>2</sub> and Pd­(η<sup>1</sup>-dppm)­(η<sup>1</sup>-<b>L</b>)­(η<sup>2</sup>-<b>L</b>) coexist in a solution of <b>4</b>. <b>L</b> coordinated on Pd<sup>II</sup> undergoes a S-demethylation reaction leading to dithiolene complexes and Me<b>L</b>. This transformation requires high temperature, and its efficiency depends on the nature of the phosphines as well as the nature of the metal (Pd vs Pt). DFT calculations reveal that the most likely mechanism depends on the lability of phosphines. Starting from M­(PR<sub>3</sub>)<sub>2</sub>(η<sup>1</sup>-<b>L</b>)<sub>2</sub> (M= Pd and Pt; R = Ph and Et), the favored sequence implies decoordination of one triethyl phosphine (M­(PEt<sub>3</sub>)­(η<sup>1</sup>-<b>L</b>)­(η<sup>2</sup>-<b>L</b>)<sub>2</sub> as intermediate) or two triphenylphosphines (Pd­(η<sup>2</sup>-<b>L</b>)<sub>2</sub> as intermediate) followed by oxidative addition and reductive elimination (OA/RE) reactions. In the case of PEt<sub>3</sub>, this OA/RE sequence can also compete with an intramolecular nucleophilic addition (<b><b>A<sub>N</sub></b></b>), which can be described as an attack of a thiolate sulfur atom on a CH<sub>3</sub><sup>+</sup> carbocation. An intramolecular <b>S<sub>N</sub>2</b> process was found to be the most feasible, starting from M­(dppe)­(η<sup>1</sup>-<b>L</b>)<sub>2</sub> (M= Pd and Pt), with the nucleophile approaching the thiomethyl group at an angle of 180° with respect to the C<sub>CH<sub>3</sub></sub>–S bond. The influence of the coligand has also been studied experimentally. Structurally characterized disulfide <b>L</b>–<b>L</b> dimer has been isolated upon reaction of 2 equiv of <b>L</b> with MCl<sub>2</sub> (M = Pd and Pt)

    Experimental and Theoretical Studies on the Mechanism of the C–S Bond Activation of Pd<sup>II</sup> Thiolate/Thioether Complexes

    No full text
    Two equivalents of <b>L</b> (<b>L</b> = 4-methylthio-2-thioxo-1,3-dithiole-5-thiolate or Medmit) react with <i>cis</i>-Pd­(PR<sub>3</sub>)<sub>2</sub>Cl<sub>2</sub> (R = Ph and Et) to afford Pd­(PR<sub>3</sub>)­(η<sup>1</sup>-<b>L</b>)­(η<sup>2</sup>-<b>L</b>) (R = Et: <b>1</b> ; R = Ph: <b>2</b>) complexes, which have been characterized by X-ray crystallography. These compounds are dynamic in solution due to an exchange of the thiomethyl groups on palladium. Variable-temperature <sup>1</sup>H NMR spectroscopy reveals a low coalescence temperature (173 K). Treatment of Pd­(<i>diphos</i>)­Cl<sub>2</sub> (<i>diphos</i> = dppe or dppm) with 2 equiv of <b>L</b> affords thiolato complexes Pd­(dppe)­(η<sup>1</sup>-<b>L</b>)<sub>2</sub> (<b>3</b>) and Pd­(dppm)­(η<sup>1</sup>-<b>L</b>)<sub>2</sub> (<b>4</b>). Whereas <b>3</b> is rigid in solution with firm η<sup>2</sup>-coordination of dppe and η<sup>1</sup>-coordination of the thiolate, two linkage isomers Pd­(η<sup>2</sup>-dppm)­(η<sup>1</sup>-<b>L</b>)<sub>2</sub> and Pd­(η<sup>1</sup>-dppm)­(η<sup>1</sup>-<b>L</b>)­(η<sup>2</sup>-<b>L</b>) coexist in a solution of <b>4</b>. <b>L</b> coordinated on Pd<sup>II</sup> undergoes a S-demethylation reaction leading to dithiolene complexes and Me<b>L</b>. This transformation requires high temperature, and its efficiency depends on the nature of the phosphines as well as the nature of the metal (Pd vs Pt). DFT calculations reveal that the most likely mechanism depends on the lability of phosphines. Starting from M­(PR<sub>3</sub>)<sub>2</sub>(η<sup>1</sup>-<b>L</b>)<sub>2</sub> (M= Pd and Pt; R = Ph and Et), the favored sequence implies decoordination of one triethyl phosphine (M­(PEt<sub>3</sub>)­(η<sup>1</sup>-<b>L</b>)­(η<sup>2</sup>-<b>L</b>)<sub>2</sub> as intermediate) or two triphenylphosphines (Pd­(η<sup>2</sup>-<b>L</b>)<sub>2</sub> as intermediate) followed by oxidative addition and reductive elimination (OA/RE) reactions. In the case of PEt<sub>3</sub>, this OA/RE sequence can also compete with an intramolecular nucleophilic addition (<b><b>A<sub>N</sub></b></b>), which can be described as an attack of a thiolate sulfur atom on a CH<sub>3</sub><sup>+</sup> carbocation. An intramolecular <b>S<sub>N</sub>2</b> process was found to be the most feasible, starting from M­(dppe)­(η<sup>1</sup>-<b>L</b>)<sub>2</sub> (M= Pd and Pt), with the nucleophile approaching the thiomethyl group at an angle of 180° with respect to the C<sub>CH<sub>3</sub></sub>–S bond. The influence of the coligand has also been studied experimentally. Structurally characterized disulfide <b>L</b>–<b>L</b> dimer has been isolated upon reaction of 2 equiv of <b>L</b> with MCl<sub>2</sub> (M = Pd and Pt)

    Reactivity of Silyl-Substituted Iron–Platinum Hydride Complexes toward Unsaturated Molecules: 4. Insertion of Fluorinated Aromatic Alkynes into the Platinum–Hydride Bond. Synthesis and Reactivity of Heterobimetallic Dimetallacylopentenone, Dimetallacyclobutene, μ‑Vinylidene, and μ<sub>2</sub>‑σ-Alkenyl Complexes

    No full text
    Insertion of <i>p</i>-FC<sub>6</sub>H<sub>4</sub>CCH, <i>p</i>-CF<sub>3</sub>C<sub>6</sub>H<sub>4</sub>CCH, and <i>m</i>-CF<sub>3</sub>C<sub>6</sub>H<sub>4</sub>CCH into the Pt–H bond of [(OC)<sub>3</sub>Fe­{Si­(OMe)<sub>3</sub>}­(μ-dppm)­Pt­(H)­(PPh<sub>3</sub>)] (<b>1a</b>) yields first the σ-alkenyl complexes [(OC)<sub>3</sub>Fe­{μ-Si­(OMe)<sub>2</sub>(OMe)}­(μ-dppm)­Pt­(ArCCH<sub>2</sub>)] (<b>2a</b>, Ar = C<sub>6</sub>H<sub>4</sub>F-<i>p</i>; <b>2d</b>, C<sub>6</sub>H<sub>4</sub>CF<sub>3</sub>-<i>p</i>; <b>2e</b>, C<sub>6</sub>H<sub>4</sub>CF<sub>3</sub>-<i>m</i>), which react in a second step with the liberated PPh<sub>3</sub> ligand to afford the structurally characterized μ-vinylidene complexes [(OC)<sub>3</sub>Fe­(μ-dppm)­{μ-CC­(H)­C<sub>6</sub>H<sub>4</sub>F-<i>p</i>}­Pt­(PPh<sub>3</sub>)] (<b>3a</b>) and [(OC)<sub>3</sub>Fe­(μ-dppm)­{μ-CC­(H)­C<sub>6</sub>H<sub>4</sub>R}­Pt­(PPh<sub>3</sub>)] (<b>3d</b>, R = <i>p</i>-CF<sub>3</sub>; <b>3e</b>, R = <i>m</i>-CF<sub>3</sub>). In contrast, treatment of <b>1a</b> with <i>o</i>-FC<sub>6</sub>H<sub>4</sub>CCH produces first [(OC)<sub>3</sub>Fe­{μ-Si­(OMe)<sub>2</sub>(OMe)}­(μ-dppm)­Pt­(<i>o</i>-FC<sub>6</sub>H<sub>4</sub>CCH<sub>2</sub>)] (<b>2b</b>), which evolves to the dimetallacyclopentenone complex [(OC)<sub>2</sub>Fe­(μ-dppm)­{μ-C­(O)­C­(H)C­(C<sub>6</sub>H<sub>4</sub>F-<i>o</i>)}­Pt­(PPh<sub>3</sub>)] (<b>4b′</b>). The latter slowly rearranges to the structurally characterized thermodynamic isomer [(OC)<sub>2</sub>Fe­(μ-dppm)­{μ-C­(O)­C­(C<sub>6</sub>H<sub>4</sub>F-<i>o</i>)C­(H)}­Pt­(PPh<sub>3</sub>)] (<b>4b</b>). Treatment of <b>1a</b> with 2,4-F<sub>2</sub>C<sub>6</sub>H<sub>3</sub>CCH produces via transient alkenyl complex <b>2c</b> an isomeric mixture of [(OC)<sub>3</sub>Fe­(μ-dppm)­{μ-CC­(H)­C<sub>6</sub>H<sub>3</sub>F<sub>2</sub>-2,4}­Pt­(PPh<sub>3</sub>)] (<b>3c</b>), [(OC)<sub>2</sub>Fe­(μ-dppm)­{μ-C­(O)­C­(H)C­(C<sub>6</sub>H<sub>3</sub>F<sub>2</sub>-2,4)}­Pt­(PPh<sub>3</sub>)] (<b>4c′</b>), and [(OC)<sub>2</sub>Fe­(μ-dppm)­{μ-C­(O)­C­(C<sub>6</sub>H<sub>3</sub>F<sub>2</sub>-2,4)C­(H)}­Pt­(PPh<sub>3</sub>)] (<b>4c</b>). Alternatively, <b>4b</b>,<b>c</b> and [(OC)<sub>2</sub>Fe­(μ-dppm)­{μ-C­(O)­C­(Ar)C­(H)}­Pt­(PPh<sub>3</sub>)] (<b>4a</b>, Ar = C<sub>6</sub>H<sub>4</sub>F-<i>p</i>, <b>4d</b>, Ar = C<sub>6</sub>H<sub>4</sub>CF<sub>3</sub>-<i>p</i>) were obtained by reaction of [(OC)<sub>3</sub>Fe­(μ-dppm)­(μ-CO)­Pt­(PPh<sub>3</sub>)] with the respective terminal alkyne. Upon reaction of <b>1a</b>, [(OC)<sub>3</sub>Fe­{Si­(OMe)<sub>3</sub>}­(μ-dppa)­Pt­(H)­(PPh<sub>3</sub>)] (<b>1b</b>; dppa = bis­(diphenylphosphino)­amine), and [(OC)<sub>3</sub>Fe­{Si­(OMe)<sub>3</sub>}­(μ-dppm)­Pt­(H)­(PMePh<sub>2</sub>)] (<b>1c</b>) with <i>o</i>-F<sub>3</sub>CC<sub>6</sub>H<sub>4</sub>CCH, the dimetallacyclobutenes [(OC)<sub>3</sub>Fe­(μ-PPh<sub>2</sub>XPPh<sub>2</sub>)­{μ-C­(C<sub>6</sub>H<sub>4</sub>CF<sub>3</sub>-<i>o</i>)­CC­(H)}­Pt­(PPh<sub>2</sub>R)] (<b>5a</b>, X = CH<sub>2</sub>, R = Ph; <b>5b</b>, X = NH, R = Ph; <b>5c</b>, X = CH<sub>2</sub>, R = Me) are formed as the sole products. Complexes <b>5</b> result also from the reaction of [(OC)<sub>3</sub>Fe­(μ-Ph<sub>2</sub>PXPPh<sub>2</sub>)­(μ-CO)­Pt­(PPh<sub>3</sub>)] (X = CH<sub>2</sub>, NH) with <i>o</i>-trifluorophenylacetylene. NMR studies at variable temperatures reveal that dimetallacyclobutenes <b>5</b> are in equilibrium with dimetallacyclopentenones [(OC)<sub>2</sub>Fe­(μ-Ph<sub>2</sub>PXPPh<sub>2</sub>)­{μ-C­(O)­C­(C<sub>6</sub>H<sub>4</sub>CF<sub>3</sub>-<i>o</i>)C­(H)}­Pt­(PPh<sub>2</sub>R)] (<b>4e</b>, X = CH<sub>2</sub>, R = Ph; <b>4f</b>, X = NH, R = Ph; <b>4g</b>, X = CH<sub>2</sub>, R = Me). Addition of HBF<sub>4</sub> to <b>5</b> leads to formation of the Fe-σ:μ<sub>2</sub>-alkenyl salts [(OC)<sub>3</sub>Fe­(μ-Ph<sub>2</sub>PXPPh<sub>2</sub>)­{μ-C­(C<sub>6</sub>H<sub>4</sub>CF<sub>3</sub>-<i>o</i>)CH<sub>2</sub>}­Pt­(PPh<sub>3</sub>)]­[BF<sub>4</sub>] (<b>6a</b>, X = CH<sub>2</sub>;<b> 6b</b>, X = NH). Protonation of <b>4</b> gives the isomeric Pt-σ:μ<sub>2</sub>-alkenyl salts [(OC)<sub>3</sub>Fe­(μ-dppm)­{μ-CH<sub>2</sub>C­(Ar)}­Pt­(PPh<sub>3</sub>)]­[BF<sub>4</sub>] (<b>7</b>) together with small amounts of the Fe-σ:μ<sub>2</sub>-alkenyl salts [(OC)<sub>3</sub>Fe­(μ-dppm)­{μ-C­(Ar)CH<sub>2</sub>}­Pt­(PPh<sub>3</sub>)]­[BF<sub>4</sub>] (Ar = <i>p</i>-C<sub>6</sub>H<sub>4</sub>CF<sub>3</sub>, <i>p</i>-C<sub>6</sub>H<sub>4</sub>F, 2,4-C<sub>6</sub>H<sub>3</sub>F<sub>2</sub>). Protonation of the vinylidene complexes [(OC)<sub>3</sub>Fe­(μ-dppm)­{μ-CC­(H)­Ar}­Pt­(PPh<sub>3</sub>)] (<b>3</b>; Ar = <i>p</i>-C<sub>6</sub>H<sub>4</sub>CF<sub>3</sub>, Ph, <i>p</i>-C<sub>6</sub>H<sub>4</sub>CH<sub>3</sub>) with HBF<sub>4</sub> occurs exclusively at the α-position of the vinylidene unit to produce a mixture of the isomeric σ-alkenyl salts <i>cis</i>-[(OC)<sub>3</sub>Fe­(μ-dppm)­{μ-C­(H)C­(H)­Ar}­Pt­(PPh<sub>3</sub>)]­[BF<sub>4</sub>] (<b>8-<i>cis</i></b>) and <i>trans</i>-[(OC)<sub>3</sub>Fe­(μ-dppm)­{μ-C­(H)C­(H)­Ar}­Pt­(PPh<sub>3</sub>)]­[BF<sub>4</sub>] (<b>8-</b><i><b>trans</b></i>)

    1,4-Bis(arylthio)but-2-enes as Assembling Ligands for (Cu<sub>2</sub>X<sub>2</sub>)<sub><i>n</i></sub> (X = I, Br; <i>n</i> = 1, 2) Coordination Polymers: Aryl Substitution, Olefin Configuration, and Halide Effects on the Dimensionality, Cluster Size, and Luminescence Properties

    No full text
    CuI reacts with <i>E</i>-PhS­(CH<sub>2</sub>CHCHCH<sub>2</sub>)­SPh, <b>L1</b>, to afford the coordination polymer (CP) [Cu<sub>2</sub>I<sub>2</sub>­{μ-<i>E</i>-PhS­(CH<sub>2</sub>CHCHCH<sub>2</sub>)­SPh}<sub>2</sub>]<sub><i>n</i></sub> (<b>1a</b>). The unprecedented square-grid network of <b>1</b> is built upon alternating two-dimensional (2D) layers with an ABAB sequence and contains rhomboid Cu<sub>2</sub>(μ<sub>2</sub>-I)<sub>2</sub> clusters as secondary building units (SBUs). Notably, layer A, interconnected by bridging <b>L1</b> ligands, contains exclusively dinuclear units with short Cu···Cu separations [2.6485(7) Å; 115 K]. In contrast, layer B exhibits Cu···Cu distances of 2.8133(8) Å. The same network is observed when CuBr reacts with <b>L1</b>. In the 2D network of [Cu<sub>2</sub>Br<sub>2</sub>­{μ-<i>E</i>-PhS­(CH<sub>2</sub>CHCHCH<sub>2</sub>)­SPh}<sub>2</sub>]<sub><i>n</i></sub> (<b>1b</b>), isotype to <b>1a</b>, one square-grid-type layer contains Cu<sub>2</sub>(μ<sub>2</sub>-Br)<sub>2</sub> SBUs with short Cu···Cu contacts [2.7422(6) Å at 115K], whereas the next layer incorporates exclusively Cu<sub>2</sub>(μ<sub>2</sub>-Br)<sub>2</sub> SBUs with a significantly longer Cu···Cu separation [2.9008(10) Å]. The evolution of the crystallographic parameters of <b>1a</b> and <b>1b</b> was monitored between 115 and 275 K. Conversely, the isomeric <i>Z</i>-PhS­(CH<sub>2</sub>CHCHCH<sub>2</sub>)­SPh ligand <b>L2</b> reacts with CuI to form the 2D CP [Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub>(μ-<i>­Z</i>-PhS­(CH<sub>2</sub>CHCHCH<sub>2</sub>)­SPh}<sub>2</sub>]<sub><i>n</i></sub> (<b>2a</b>) with closed-cubane SBUs. A dinuclear zero-dimensional complex [Cu<sub>2</sub>Br<sub>2</sub>­{μ-<i>Z</i>-PhS­(CH<sub>2</sub>CHCHCH<sub>2</sub>)­SPh}<sub>2</sub>] (<b>2b</b>) is formed when CuBr is reacted with <b>L2</b>. Upon reaction of <i>E</i>-TolS­(CH<sub>2</sub>CHCHCH<sub>2</sub>)­STol, <b>L3</b>, with CuI, the 2D CP [{Cu­(μ<sub>3</sub>-I)}<sub>2</sub>­(μ-<b>L3</b>)]<sub><i>n</i></sub> containing parallel-arranged infinite inorganic staircase ribbons, is generated. When CuX reacts with <i>Z</i>-TolS­(CH<sub>2</sub>CHCHCH<sub>2</sub>)­STol, <b>L4</b>, the isostructural 2D CPs [Cu<sub>2</sub>X<sub>2</sub>­{μ-<i>Z</i>-TolS­(CH<sub>2</sub>CHCHCH<sub>2</sub>)­STol}<sub>2</sub>] <b>(4a</b> X = I; <b>4b</b> X = Br) are formed. In contrast to the CPs <b>1a,b</b>, the layers based on rhombic grids of <b>4a,b</b> incorporate Cu<sub>2</sub>(μ<sub>2</sub>-X)<sub>2</sub> SBUs featuring uniformly identical Cu···Cu distances within each layer. The TGA traces showed that all these materials are stable up to ∼200 °C. Moreover, the photophysical properties have been studied, including absorption, emission, excitation spectra, and emission lifetimes at 298 and 77 K. The spectra were interpreted using density functional theory (DFT) and time-dependent DFT calculations

    Reactivity of CuI and CuBr toward Dialkyl Sulfides RSR: From Discrete Molecular Cu<sub>4</sub>I<sub>4</sub>S<sub>4</sub> and Cu<sub>8</sub>I<sub>8</sub>S<sub>6</sub> Clusters to Luminescent Copper(I) Coordination Polymers

    No full text
    The 1D coordination polymer (CP) [(Me<sub>2</sub>S)<sub>3</sub>{Cu<sub>2</sub>(μ-I)<sub>2</sub>}]<sub><i>n</i></sub> (<b>1</b>) is formed when CuI reacts with SMe<sub>2</sub> in <i>n</i>-heptane, whereas in acetonitrile (MeCN), the reaction forms exclusively the 2D CP [(Me<sub>2</sub>S)<sub>3</sub>{Cu<sub>4</sub>(μ-I)<sub>4</sub>}]<sub><i>n</i></sub> (<b>2</b>) containing “flower-basket” Cu<sub>4</sub>I<sub>4</sub> units. The reaction product of CuI with MeSEt is also solvent-dependent, where the 1D polymer [(MeSEt)<sub>2</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>2</sub>(μ<sub>2</sub>-I)<sub>2</sub>}­(MeCN)<sub>2</sub>]<sub><i>n</i></sub> (<b>3</b>) containing “stepped-cubane” Cu<sub>4</sub>I<sub>4</sub> units is isolated in MeCN. In contrast, the reaction in <i>n</i>-heptane affords the 1D CP [(MeSEt)<sub>3</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub>}]<sub><i>n</i></sub> (<b>4</b>) containing “closed-cubane” Cu<sub>4</sub>I<sub>4</sub> clusters. The reaction of MeSPr with CuI provides the structurally related 1D CP [(MeSPr)<sub>3</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub>}]<sub><i>n</i></sub> (<b>5</b>), for which the X-ray structure has been determined at 115, 155, 195, 235, and 275 K, addressing the evolution of the metric parameters. Similarly to <b>4</b> and the previously reported CP [(Et<sub>2</sub>S)<sub>3</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub>}]<sub><i>n</i></sub> (<i>Inorg. Chem.</i> <b>2010</b>, <i>49</i>, 5834), the 1D chain is built upon closed cubanes Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub> as secondary building units (SBUs) interconnected via μ-MeSPr ligands. The 0D tetranuclear clusters [(L)<sub>4</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub>}] [L = EtSPr (<b>6</b>), Pr<sub>2</sub>S (<b>7</b>)] respectively result from the reaction of CuI with EtSPr and <i>n</i>-Pr<sub>2</sub>S. With <i>i</i>-Pr<sub>2</sub>S, the octanuclear cluster [(<i>i</i>-Pr<sub>2</sub>S)<sub>6</sub>{Cu<sub>8</sub>(μ<sub>3</sub>-I)<sub>3</sub>}­(μ<sub>4</sub>-I)<sub>2</sub>}] (<b>8</b>) is formed. An X-ray study has also been performed at five different temperatures for the 2D polymer [(Cu<sub>3</sub>Br<sub>3</sub>)­(MeSEt)<sub>3</sub>]<sub><i>n</i></sub> (<b>9</b>) formed from the reaction between CuBr and MeSEt in heptane. The unprecedented framework of <b>9</b> consists of layers with alternating Cu­(μ<sub>2</sub>-Br)<sub>2</sub>Cu rhomboids, which are connected through two μ-MeSEt ligands to tetranuclear open-cubane Cu<sub>4</sub>Br<sub>4</sub> SBUs. MeSPr forms with CuBr in heptane the 1D CP [(Cu<sub>3</sub>Br<sub>3</sub>)­(MeSPr)<sub>3</sub>]<sub><i>n</i></sub> (<b>10</b>), which is converted to a 2D metal–organic framework [(Cu<sub>5</sub>Br<sub>5</sub>)­(μ<sub>2</sub>-MeSPr)<sub>3</sub>]<sub><i>n</i></sub> (<b>11</b>) incorporating pentanuclear [(Cu<sub>5</sub>(μ<sub>4</sub>-Br)­(μ<sub>2</sub>-Br)] SBUs when recrystallized in MeCN. The thermal stability and photophysical properties of these materials are also reported

    Reactivity of CuI and CuBr toward Dialkyl Sulfides RSR: From Discrete Molecular Cu<sub>4</sub>I<sub>4</sub>S<sub>4</sub> and Cu<sub>8</sub>I<sub>8</sub>S<sub>6</sub> Clusters to Luminescent Copper(I) Coordination Polymers

    No full text
    The 1D coordination polymer (CP) [(Me<sub>2</sub>S)<sub>3</sub>{Cu<sub>2</sub>(μ-I)<sub>2</sub>}]<sub><i>n</i></sub> (<b>1</b>) is formed when CuI reacts with SMe<sub>2</sub> in <i>n</i>-heptane, whereas in acetonitrile (MeCN), the reaction forms exclusively the 2D CP [(Me<sub>2</sub>S)<sub>3</sub>{Cu<sub>4</sub>(μ-I)<sub>4</sub>}]<sub><i>n</i></sub> (<b>2</b>) containing “flower-basket” Cu<sub>4</sub>I<sub>4</sub> units. The reaction product of CuI with MeSEt is also solvent-dependent, where the 1D polymer [(MeSEt)<sub>2</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>2</sub>(μ<sub>2</sub>-I)<sub>2</sub>}­(MeCN)<sub>2</sub>]<sub><i>n</i></sub> (<b>3</b>) containing “stepped-cubane” Cu<sub>4</sub>I<sub>4</sub> units is isolated in MeCN. In contrast, the reaction in <i>n</i>-heptane affords the 1D CP [(MeSEt)<sub>3</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub>}]<sub><i>n</i></sub> (<b>4</b>) containing “closed-cubane” Cu<sub>4</sub>I<sub>4</sub> clusters. The reaction of MeSPr with CuI provides the structurally related 1D CP [(MeSPr)<sub>3</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub>}]<sub><i>n</i></sub> (<b>5</b>), for which the X-ray structure has been determined at 115, 155, 195, 235, and 275 K, addressing the evolution of the metric parameters. Similarly to <b>4</b> and the previously reported CP [(Et<sub>2</sub>S)<sub>3</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub>}]<sub><i>n</i></sub> (<i>Inorg. Chem.</i> <b>2010</b>, <i>49</i>, 5834), the 1D chain is built upon closed cubanes Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub> as secondary building units (SBUs) interconnected via μ-MeSPr ligands. The 0D tetranuclear clusters [(L)<sub>4</sub>{Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub>}] [L = EtSPr (<b>6</b>), Pr<sub>2</sub>S (<b>7</b>)] respectively result from the reaction of CuI with EtSPr and <i>n</i>-Pr<sub>2</sub>S. With <i>i</i>-Pr<sub>2</sub>S, the octanuclear cluster [(<i>i</i>-Pr<sub>2</sub>S)<sub>6</sub>{Cu<sub>8</sub>(μ<sub>3</sub>-I)<sub>3</sub>}­(μ<sub>4</sub>-I)<sub>2</sub>}] (<b>8</b>) is formed. An X-ray study has also been performed at five different temperatures for the 2D polymer [(Cu<sub>3</sub>Br<sub>3</sub>)­(MeSEt)<sub>3</sub>]<sub><i>n</i></sub> (<b>9</b>) formed from the reaction between CuBr and MeSEt in heptane. The unprecedented framework of <b>9</b> consists of layers with alternating Cu­(μ<sub>2</sub>-Br)<sub>2</sub>Cu rhomboids, which are connected through two μ-MeSEt ligands to tetranuclear open-cubane Cu<sub>4</sub>Br<sub>4</sub> SBUs. MeSPr forms with CuBr in heptane the 1D CP [(Cu<sub>3</sub>Br<sub>3</sub>)­(MeSPr)<sub>3</sub>]<sub><i>n</i></sub> (<b>10</b>), which is converted to a 2D metal–organic framework [(Cu<sub>5</sub>Br<sub>5</sub>)­(μ<sub>2</sub>-MeSPr)<sub>3</sub>]<sub><i>n</i></sub> (<b>11</b>) incorporating pentanuclear [(Cu<sub>5</sub>(μ<sub>4</sub>-Br)­(μ<sub>2</sub>-Br)] SBUs when recrystallized in MeCN. The thermal stability and photophysical properties of these materials are also reported

    Construction of (CuX)<sub>2<i>n</i></sub> Cluster-Containing (X = Br, I; <i>n</i> = 1, 2) Coordination Polymers Assembled by Dithioethers ArS(CH<sub>2</sub>)<sub><i>m</i></sub>SAr (Ar = Ph, <i>p</i>‑Tol; <i>m</i> = 3, 5): Effect of the Spacer Length, Aryl Group, and Metal-to-Ligand Ratio on the Dimensionality, Cluster Nuclearity, and the Luminescence Properties of the Metal–Organic Frameworks

    No full text
    Reaction of CuI with bis­(phenylthio)­propane in a 1:1 ratio yields the two-dimensional coordination polymer [{Cu­(μ<sub>2</sub>-I)<sub>2</sub>Cu}­{μ-PhS­(CH<sub>2</sub>)<sub>3</sub>SPh}<sub>2</sub>]<sub><i>n</i></sub> (<b>1</b>). The 2D-sheet structure of <b>1</b> is built up by dimeric Cu<sub>2</sub>I<sub>2</sub> units, which are connected via four bridging 1,3-bis­(phenylthio)­propane ligands. In contrast, treatment of 2 equiv of CuI with 1,3-bis­(phenylthio)­propane in MeCN solution affords in a self-assembly reaction the strongly luminescent metal–organic 2D-coordination polymer [Cu<sub>4</sub>I<sub>4</sub>{μ-PhS­(CH<sub>2</sub>)<sub>3</sub>Ph}<sub>2</sub>]<sub><i>n</i></sub> (<b>2</b>), in which cubane-like Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub> cluster units are linked by the dithioether ligands. The crystallographically characterized one-dimensional (1D) compound [{Cu­(μ<sub>2</sub>-Br)<sub>2</sub>Cu}­{μ-PhS­(CH<sub>2</sub>)<sub>3</sub>SPh}<sub>2</sub>]<sub><i>n</i></sub> (<b>3</b>) is obtained using CuBr. The outcome of the reaction of PhS­(CH<sub>2</sub>)<sub>5</sub>SPh with CuI also depends of the metal-to-ligand ratio employed. Mixing CuI and the dithioether in a 2:1 ratio results in formation of [Cu<sub>4</sub>I<sub>4</sub>{μ-PhS­(CH<sub>2</sub>)<sub>5</sub>Ph}<sub>2</sub>]<sub><i>n</i></sub> (<b>4</b>) in which cubane-like Cu<sub>4</sub>(μ<sub>3</sub>-I)<sub>4</sub> clusters are linked by the bridging dithioether ligand giving rise to a 1D necklace structure. A ribbon-like 1D-polymer with composition [{Cu­(μ<sub>2</sub>-I)<sub>2</sub>Cu}­{μ-PhS­(CH<sub>2</sub>)<sub>5</sub>SPh}<sub>2</sub>]<sub><i>n</i></sub> (<b>5</b>), incorporating rhomboid Cu<sub>2</sub>I<sub>2</sub> units, is produced upon treatment of CuI with 1,5-bis­(phenylthio)­pentane in a 1:1 ratio. Reaction of CuBr with PhS­(CH<sub>2</sub>)<sub>5</sub>SPh produces the isomorphous 1D-compound [{Cu­(μ<sub>2</sub>-Br)<sub>2</sub>Cu}­{μ-PhS­(CH<sub>2</sub>)<sub>5</sub>SPh}<sub>2</sub>]<sub><i>n</i></sub> (<b>6</b>). Strongly luminescent [Cu<sub>4</sub>I<sub>4</sub>{μ-<i>p</i>-TolS­(CH<sub>2</sub>)<sub>5</sub>STol-<i>p</i>}<sub>2</sub>]<sub><i>n</i></sub> (<b>7</b>) is obtained after mixing 1,5-bis­(<i>p</i>-tolylthio)­pentane with CuI in a 1:2 ratio, and the 2D-polymer [{Cu­(μ<sub>2</sub>-I)<sub>2</sub>Cu}<sub>2</sub>{μ-<i>p</i>-TolS­(CH<sub>2</sub>)<sub>5</sub>STol-<i>p</i>}<sub>2</sub>]<sub><i>n</i></sub> (<b>8</b>) results from reaction in a 1:1 metal-to-ligand ratio. Under the same reaction conditions, 1D-polymeric [{Cu­(μ<sub>2</sub>-Br)<sub>2</sub>Cu}­{μ-<i>p</i>-TolS­(CH<sub>2</sub>)<sub>5</sub>STol-p}<sub>2</sub>]<sub><i>n</i></sub> (<b>9</b>) is formed using CuBr. This study reveals that the structure of the self-assembly process between CuX and ArS­(CH<sub>2</sub>)<sub><i>m</i></sub>SAr ligands is hard to predict. The solid-state luminescence spectra at 298 and 77 K of <b>2</b> and <b>4</b> exhibit very strong emissions around 535 and 560 nm, respectively, whereas those for <b>1</b> and <b>5</b> display weaker ones at about 450 nm. The emission lifetimes are longer for the longer wavelength emissions (>1.0 μs arising from the cubane species) and shorter for the shorter wavelength ones (<1.4 μs arising from the rhomboid units). The Br-containing species are found to be weakly fluorescent

    Regio- and Stereoselective Synthesis of Spiropyrrolizidines and Piperazines through Azomethine Ylide Cycloaddition Reaction

    No full text
    A series of original spiropyrrolizidine derivatives has been prepared by a one-pot three-component [3 + 2] cycloaddition reaction of (<i>E</i>)-3-arylidene-1-phenyl-pyrrolidine-2,5-diones, l-proline, and the cyclic ketones 1<i>H</i>-indole-2,3-dione (isatin), indenoquinoxaline-11-one and acenaphthenequinone. We disclose an unprecedented isomerization of some spiroadducts leading to a new family of spirooxindolepyrrolizidines. Furthermore, these cycloadducts underwent retro-1,3-dipolar cycloaddition yielding unexpected regioisomers. Upon treatment of the dipolarophiles with <i>in situ</i> generated azomethine ylides from l-proline or acenaphthenequinone, formation of spiroadducts and unusual polycyclic fused piperazines through a stepwise [3 + 3] cycloaddition pathway is observed. The stereochemistry of these N-heterocycles has been confirmed by several X-ray diffraction studies. Some of these compounds exhibit extensive hydrogen bonding in the crystalline state. To enlighten the observed regio- and stereoselectivity of the [3 + 2] cycloaddition, calculations using the DFT approach at the B3LYP/6-31G­(d,p) level were carried out. It was found that this reaction is under kinetic control
    corecore