20 research outputs found

    Metal-Templated Hydrogen Bond Donors as ā€œOrganocatalystsā€ for Carbonā€“Carbon Bond Forming Reactions: Syntheses, Structures, and Reactivities of 2ā€‘Guanidinobenzimidazole Cyclopentadienyl Ruthenium Complexes

    No full text
    The reaction of 2-guanidinobenzimidazole (GBI) and (Ī·<sup>5</sup>-C<sub>5</sub>H<sub>5</sub>)Ā­RuĀ­(PPh<sub>3</sub>)<sub>2</sub>(Cl) in refluxing toluene gives the chelate [(Ī·<sup>5</sup>-C<sub>5</sub>H<sub>5</sub>)Ā­RuĀ­(PPh<sub>3</sub>)Ā­(GBI)]<sup>+</sup>Cl<sup>ā€“</sup> (<b>1</b><sup>+</sup>Cl<sup>ā€“</sup>; 96%). Subsequent anion metatheses yield the BF<sub>4</sub><sup>ā€“</sup>, PF<sub>6</sub><sup>ā€“</sup>, and BAr<sub>f</sub><sup>ā€“</sup> (BĀ­(3,5-C<sub>6</sub>H<sub>3</sub>(CF<sub>3</sub>)<sub>2</sub>)<sub>4</sub><sup>ā€“</sup>) salts (77ā€“85%). Reactions with CO give the carbonyl complexes [(Ī·<sup>5</sup>-C<sub>5</sub>H<sub>5</sub>)Ā­RuĀ­(CO)Ā­(GBI)]<sup>+</sup>X<sup>ā€“</sup> (<b>2</b><sup>+</sup>X<sup>ā€“</sup>; X<sup>ā€“</sup> = Cl<sup>ā€“</sup>, BF<sub>4</sub><sup>ā€“</sup>, PF<sub>6</sub><sup>ā€“</sup>, BAr<sub>f</sub><sup>ā€“</sup>; 87ā€“92%). The last three salts can also be obtained by anion metatheses of <b>2</b><sup>+</sup>Cl<sup>ā€“</sup> (77ā€“87%), as can one with the chiral enantiopure anion PĀ­(<i>o</i>-C<sub>6</sub>Cl<sub>4</sub>O<sub>2</sub>)<sub>3</sub><sup>ā€“</sup> ((Ī”)-TRISPHAT<sup>ā€“</sup>; 81%). The reaction of [(Ī·<sup>5</sup>-C<sub>5</sub>H<sub>5</sub>)Ā­RuĀ­(CO)Ā­(NCCH<sub>3</sub>)<sub>2</sub>]<sup>+</sup>PF<sub>6</sub><sup>ā€“</sup> and GBI also gives <b>2</b><sup>+</sup>PF<sub>6</sub><sup>ā€“</sup> (81%). The pentamethylcyclopentadienyl analogues [(Ī·<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)Ā­RuĀ­(CO)Ā­(GBI)]<sup>+</sup>X<sup>ā€“</sup> (<b>3</b><sup>+</sup>X<sup>ā€“</sup>; X<sup>ā€“</sup> = Cl<sup>ā€“</sup>, BF<sub>4</sub><sup>ā€“</sup>, PF<sub>6</sub><sup>ā€“</sup>, BAr<sub>f</sub><sup>ā€“</sup>; 61ā€“84%) are prepared from (Ī·<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)Ā­RuĀ­(PPh<sub>3</sub>)<sub>2</sub>(Cl), GBI, and CO followed (for the last three) by anion metatheses. An indenyl complex [(Ī·<sup>5</sup>-C<sub>9</sub>H<sub>7</sub>)Ā­RuĀ­(PPh<sub>3</sub>)Ā­(GBI)]<sup>+</sup>Cl<sup>ā€“</sup> (96%) is prepared from (Ī·<sup>5</sup>-C<sub>9</sub>H<sub>7</sub>)Ā­RuĀ­(PPh<sub>3</sub>)<sub>2</sub>(Cl) and GBI. All complexes are characterized by NMR (<sup>1</sup>H, <sup>13</sup>C, <sup>31</sup>P, <sup>19</sup>F, <sup>11</sup>B), with 2D spectra aiding assignments. Crystal structures of <b>1</b><sup>+</sup>PF<sub>6</sub><sup>ā€“</sup>Ā·CH<sub>2</sub>Cl<sub>2</sub> and <b>1</b><sup>+</sup>BAr<sub>f</sub><sup>ā€“</sup>Ā·CH<sub>2</sub>Cl<sub>2</sub> are determined; the anion is hydrogen bonded to the cation in the former. Complexes <b>1</b>ā€“<b>3</b><sup>+</sup>X<sup>ā€“</sup> are evaluated as catalysts (10 mol %, RT) for condensations of indoles and <i>trans</i>-Ī²-nitrostyrene. The chloride salts are ineffective (0ā€“5% yields, 48ā€“60 h), but the BAr<sub>f</sub><sup>ā€“</sup> salts exhibit excellent reactivities (97ā€“46% yields, 1ā€“48 h), with the BF<sub>4</sub><sup>ā€“</sup> and PF<sub>6</sub><sup>ā€“</sup> salts intermediate. Evidence for hydrogen bonding of the nitro group to the GBI ligand is presented. GBI shows no catalytic activity; a BAr<sub>f</sub><sup>ā€“</sup> salt of methylated GBI is active, but much less so than <b>2</b>ā€“<b>3</b><sup>+</sup>BAr<sub>f</sub><sup>ā€“</sup>

    Octahedral Gyroscope-like Molecules Consisting of Rhenium Rotators within Cage-like Dibridgehead Diphosphine Stators: Syntheses, Substitution Reactions, Structures, and Dynamic Properties

    No full text
    Reactions of ReĀ­(CO)<sub>5</sub>(X) (X = Cl, Br) or [Re<sub>2</sub>(CO)<sub>4</sub>(NO)<sub>2</sub>(Ī¼-Cl)<sub>2</sub>(Cl)<sub>2</sub>] and the phosphines PĀ­((CH<sub>2</sub>)<sub><i>m</i></sub>CHī—»CH<sub>2</sub>)<sub>3</sub> (<i>m</i> = 6, <b>a</b>; 7, <b>b</b>; 8, <b>c</b>) give <i>mer,trans</i>-ReĀ­(CO)<sub>3</sub>(X)Ā­(PĀ­((CH<sub>2</sub>)<sub><i>m</i></sub>CHī—»CH<sub>2</sub>)<sub>3</sub>)<sub>2</sub> (53ā€“95%) or <i>cis,trans</i>-ReĀ­(CO) (NO) (Cl)<sub>2</sub>(PĀ­((CH<sub>2</sub>)<sub>6</sub>CHī—»CH<sub>2</sub>)<sub>3</sub>)<sub>2</sub> (57%), respectively. Additions of Grubbsā€™ catalyst (5ā€“10 mol %, 0.0010ā€“0.0012 M) and subsequent hydrogenations (PtO<sub>2</sub>, ā‰¤5 bar) yield the gyroscope-like complexes <i>mer,trans</i>-ReĀ­(CO)<sub>3</sub>(X)Ā­(PĀ­((CH<sub>2</sub>)<sub><i>n</i></sub>)<sub>3</sub>P) (<i>n</i> = 2<i>m</i> + 2; X = Cl, <b>7a</b>,<b>c</b>; Br, <b>8a</b>,<b>c</b>; 18ā€“61%) or <i>cis,trans</i>-ReĀ­(CO)(NO)(Cl)<sub>2</sub>(PĀ­((CH<sub>2</sub>)<sub>14</sub>)<sub>3</sub>P) (14%), respectively, and/or the isomers <i>mer,trans</i>-ReĀ­(CO)<sub>3</sub>(X)Ā­(PĀ­(CH<sub>2</sub>)<sub><i>n</i>āˆ’1</sub>CH<sub>2</sub>)Ā­((CH<sub>2</sub>)<sub><i>n</i></sub>)Ā­(PĀ­(CH<sub>2</sub>)<sub><i>n</i>āˆ’1</sub>CH<sub>2</sub>) (X = Cl, <b>7ā€²a</b>ā€“<b>c</b>; Br, <b>8ā€²b</b>; 6ā€“27%). The latter are derived from a combination of interligand and intraligand metatheses. Reactions of <b>7a</b> or <b>8a</b> with NaI, Ph<sub>2</sub>Zn, or MeLi give <i>mer,trans</i>-ReĀ­(CO)<sub>3</sub>(X)Ā­(PĀ­((CH<sub>2</sub>)<sub>14</sub>)<sub>3</sub>P) (X = I, <b>11a</b>; Ph, <b>12a</b>; Me, <b>13a</b>; 34ā€“87%). The <sup>13</sup>C NMR spectra of <b>7a</b>ā€“<b>c</b>, <b>8a</b>ā€“<b>c</b>, <b>11a</b>, and <b>13a</b> show rotation of the ReĀ­(CO)<sub>3</sub>(X) moieties to be fast on the NMR time scale at room temperature (and at āˆ’90 Ā°C for <b>8a</b>). In contrast, the phenyl group in <b>12a</b> acts as a brake, and two sets of <sup>13</sup>C NMR signals (2:1) are observed for the methylene chains. The crystal structures of <b>7a</b>, <b>8a</b>, <b>12a</b>, and <b>13a</b> are analyzed with respect to ReĀ­(CO)<sub>3</sub>(X) rotation in solution and the solid state

    Synthesis and Properties of Arylvinylidene-Bridged Triphenylamines

    No full text
    A series of arylvinylidene-bridged triphenylamines were efficiently synthesized through the thionation/Bartonā€“Kellogg olefination sequence from their corresponding carbonyl precursors. The electrochemical investigations identified these highly distorted scaffolds as fairly strong electron donors capable of several reversible oxidation steps with the first oxidation occurring at a potential comparable to that of ferrocene for the <i>n</i>-hexyl-substituted diphenylvinylidene-bridged compound

    Coordination of Terpyridine to Li<sup>+</sup> in Two Different Ionic Liquids

    No full text
    On the basis of <sup>7</sup>Li NMR experiments, the complex-formation reaction between Li<sup>+</sup> and the tridentate N-donor ligand terpyridine was studied in the ionic liquids [emim]Ā­[NTf<sub>2</sub>] and [emim]Ā­[ClO<sub>4</sub>] as solvents. For both ionic liquids, the NMR data implicate the formation of [Li(terpy)<sub>2</sub>]<sup>+</sup>. Density functional theory calculations show that partial coordination of terpyridine involving the coordination of a solvent anion can be excluded. In contrast to the studies in solution, X-ray diffraction measurements led to completely different results. In the case of [emim]Ā­[NTf<sub>2</sub>], the polymeric lithium species [LiĀ­(terpy)Ā­(NTf<sub>2</sub>)]<i><sub>n</sub></i> was found to control the stacking of this complex, whereas crystals grown from [emim]Ā­[ClO<sub>4</sub>] exhibit the discrete dimeric species [LiĀ­(terpy)Ā­(ClO<sub>4</sub>)]<sub>2</sub>. However, both structures indicate that each lithium ion is formally coordinated by one terpy molecule and one solvent anion in the solid state, suggesting that charge neutralization and Ļ€ stacking mainly control the crystallization process

    Gyroscope-Like Platinum and Palladium Complexes with Trans-Spanning Bis(pyridine) Ligands

    No full text
    Pyridines with one or two substituents terminating in vinyl groups are prepared. Intramolecular ring-closing metatheses of <i>trans</i>-MCl<sub>2</sub> adducts and hydrogenations supply the title compounds. Williamson ether syntheses using the alcohols HOĀ­(CH<sub>2</sub>)<sub><i>n</i></sub>CHī—»CH<sub>2</sub> (<i>n</i> = 1 (<b>a</b>), 2 (<b>b</b>), 3 (<b>c</b>), 4 (<b>d</b>), 5 (<b>e</b>), 6 (<b>f</b>), 8 (<b>h</b>), 9 (<b>i</b>)) and appropriate halides give the pyridines 2-NC<sub>5</sub>H<sub>4</sub>(CH<sub>2</sub>OĀ­(CH<sub>2</sub>)<sub><i>n</i></sub>CHī—»CH<sub>2</sub>) (<b>1a</b>,<b>b</b>), 3-NC<sub>5</sub>H<sub>4</sub>(CH<sub>2</sub>OĀ­(CH<sub>2</sub>)<sub><i>n</i></sub>CHī—»CH<sub>2</sub>) (<b>2a</b>ā€“<b>e</b>,<b>h</b>,<b>i</b>), and 2,6-NC<sub>5</sub>H<sub>3</sub>(CH<sub>2</sub>OĀ­(CH<sub>2</sub>)<sub><i>n</i></sub>CHī—»CH<sub>2</sub>)<sub>2</sub> (<b>4a</b>ā€“<b>d</b>) in 92ā€“45% yields. Reactions of 3,5-NC<sub>5</sub>H<sub>3</sub>(COCl)<sub>2</sub> and HOĀ­(CH<sub>2</sub>)<sub><i>n</i></sub>CHī—»CH<sub>2</sub> afford the diesters 3,5-NC<sub>5</sub>H<sub>3</sub>(COOĀ­(CH<sub>2</sub>)<sub><i>n</i></sub>CHī—»CH<sub>2</sub>)<sub>2</sub> (<b>5a</b>ā€“<b>f</b>,<b>h</b>, 90ā€“41%). The reaction of 3,5-NC<sub>5</sub>H<sub>3</sub>(4-C<sub>6</sub>H<sub>4</sub>OH)<sub>2</sub>, BrĀ­(CH<sub>2</sub>)<sub>5</sub>CHī—»CH<sub>2</sub>, and Cs<sub>2</sub>CO<sub>3</sub> yields 3,5-NC<sub>5</sub>H<sub>3</sub>(4-C<sub>6</sub>H<sub>4</sub>OĀ­(CH<sub>2</sub>)<sub>5</sub>CHī—»CH<sub>2</sub>)<sub>2</sub> (<b>8</b>; 32%). Reactions of PtCl<sub>2</sub> with <b>1a</b>,<b>b</b>, <b>2a</b>ā€“<b>e</b>,<b>h</b>,<b>i</b>, <b>4a</b>,<b>b</b> (but not <b>4c</b>,<b>d</b>), <b>5a</b>,<b>c</b>ā€“<b>f</b>,<b>h</b>, and <b>8</b> afford the corresponding bisĀ­(pyridine) complexes <i>trans</i>-<b>10a</b>,<b>b</b> (40ā€“12%), <i>trans</i>-<b>12a</b>ā€“<b>e</b>,<b>h</b>,<b>i</b> (84ā€“46%), <i>trans</i>-<b>17a</b>,<b>b</b> (88ā€“22%), <i>trans</i>-<b>19a</b>,<b>c</b>ā€“<b>f</b>,<b>h</b> (94ā€“63%), and <i>trans</i>-<b>22</b> (96%). Selected adducts are treated with Grubbsā€™ catalyst and then H<sub>2</sub> (Pd/C) to give <i>trans</i>-PtCl<sub>2</sub>[2,2ā€²-(NC<sub>5</sub>H<sub>4</sub>(CH<sub>2</sub>OĀ­(CH<sub>2</sub>)<sub>2<i>n</i>+2</sub>OCH<sub>2</sub>)Ā­H<sub>4</sub>C<sub>5</sub>N)] (<i>trans</i>-<b>11a</b>,<b>b</b>; 79ā€“63%), <i>trans</i>-PtCl<sub>2</sub>[3,3ā€²-(NC<sub>5</sub>H<sub>4</sub>(CH<sub>2</sub>OĀ­(CH<sub>2</sub>)<sub>2<i>n</i>+2</sub>OCH<sub>2</sub>)Ā­H<sub>4</sub>C<sub>5</sub>N)] (<i>trans</i>-<b>13</b>,<b>d</b>,<b>h</b>,<b>i</b>; 93ā€“80%), <i>trans</i>-PtCl<sub>2</sub>[2,6,2ā€²,6ā€²-(NC<sub>5</sub>H<sub>3</sub>(CH<sub>2</sub>OĀ­(CH<sub>2</sub>)<sub>2<i>n</i>+2</sub>OCH<sub>2</sub>)<sub>2</sub>H<sub>3</sub>C<sub>5</sub>N)] (<i>trans</i>-<b>18a</b>,<b>b</b>; 22ā€“10%), <i>trans</i>-PtCl<sub>2</sub>[3,5,3ā€²,5ā€²-(NC<sub>5</sub>H<sub>3</sub>(COOĀ­(CH<sub>2</sub>)<sub>2<i>n</i>+2</sub>OCO)<sub>2</sub>H<sub>3</sub>C<sub>5</sub>N)] (<i>trans-</i><b>20d</b>ā€“<b>f</b>,<b>h</b>; 45ā€“14%), and <i>trans</i>-PtCl<sub>2</sub>[3,5,3ā€²,5ā€²-(NC<sub>5</sub>H<sub>3</sub>(4-C<sub>6</sub>H<sub>4</sub>OĀ­(CH<sub>2</sub>)<sub>12</sub>O-4-C<sub>6</sub>H<sub>4</sub>)<sub>2</sub>H<sub>3</sub>C<sub>5</sub>N)] (40%). A previously reported ring-closing metathesis of <i>trans</i>-PdCl<sub>2</sub>[2,6-NC<sub>5</sub>H<sub>3</sub>(CH<sub>2</sub>CH<sub>2</sub>CHī—»CH<sub>2</sub>)<sub>2</sub>]<sub>2</sub> is confirmed, and the new hydrogenation product <i>trans</i>-PdCl<sub>2</sub>[2,6,2ā€²,6ā€²-(NC<sub>5</sub>H<sub>3</sub>((CH<sub>2</sub>)<sub>6</sub>)<sub>2</sub>H<sub>3</sub>C<sub>5</sub>N)] (<i>trans-</i><b>16</b>; 62%) is isolated. Additions of CH<sub>3</sub>MgBr to <b>12b</b>,<b>h</b> and <b>13d</b>,<b>h</b> afford the corresponding PtClCH<sub>3</sub> species (94ā€“41%), but analogous reactions fail with 2-substituted pyridine adducts. The reaction of <i>trans</i>-<b>19c</b> with PhCī—¼CH and CuI/<i>i</i>-Pr<sub>2</sub>NH gives the corresponding PtClĀ­(Cī—¼CPh) adduct (18%). The crystal structures of <i>trans</i>-<b>17a</b>, <i>trans</i>-<b>11b</b>, <i>trans</i>-<b>13d</b>, <i>trans</i>-<b>13h</b>Ā·CH<sub>2</sub>Cl<sub>2</sub>, <i>trans</i>-<b>16</b>, <i>trans</i>-<b>18a</b>,<b>b</b>, and <i>trans</i>-<b>20e</b>Ā·2CHCl<sub>3</sub><sub></sub> are determined. Steric effects in the preceding data, especially involving 2-substituents and the MCl<sub>2</sub> or MClĀ­(X) rotators, are analyzed in detail

    Coordination of Terpyridine to Li<sup>+</sup> in Two Different Ionic Liquids

    No full text
    On the basis of <sup>7</sup>Li NMR experiments, the complex-formation reaction between Li<sup>+</sup> and the tridentate N-donor ligand terpyridine was studied in the ionic liquids [emim]Ā­[NTf<sub>2</sub>] and [emim]Ā­[ClO<sub>4</sub>] as solvents. For both ionic liquids, the NMR data implicate the formation of [Li(terpy)<sub>2</sub>]<sup>+</sup>. Density functional theory calculations show that partial coordination of terpyridine involving the coordination of a solvent anion can be excluded. In contrast to the studies in solution, X-ray diffraction measurements led to completely different results. In the case of [emim]Ā­[NTf<sub>2</sub>], the polymeric lithium species [LiĀ­(terpy)Ā­(NTf<sub>2</sub>)]<i><sub>n</sub></i> was found to control the stacking of this complex, whereas crystals grown from [emim]Ā­[ClO<sub>4</sub>] exhibit the discrete dimeric species [LiĀ­(terpy)Ā­(ClO<sub>4</sub>)]<sub>2</sub>. However, both structures indicate that each lithium ion is formally coordinated by one terpy molecule and one solvent anion in the solid state, suggesting that charge neutralization and Ļ€ stacking mainly control the crystallization process

    Partially Shielded Fe(CO)<sub>3</sub> Rotors: Syntheses, Structures, and Dynamic Properties of Complexes with Doubly <i>trans</i> Spanning Diphosphines, <i>trans</i>-Fe(CO)<sub>3</sub>(PhP((CH<sub>2</sub>)<sub><i>n</i></sub>)<sub>2</sub>PPh)

    No full text
    Reactions of FeĀ­(CO)<sub>3</sub>(Ī·<sup>4</sup>-benzylideneacetone) and PhPĀ­((CH<sub>2</sub>)<sub><i>m</i></sub>CHī—»CH<sub>2</sub>)<sub>2</sub> (<i>m</i> = <b>a</b>, 4; <b>b</b>, 5; <b>c</b>, 6) give <i>trans</i>-FeĀ­(CO)<sub>3</sub>(PhPĀ­((CH<sub>2</sub>)<sub><i>m</i></sub>CHī—»CH<sub>2</sub>)<sub>2</sub>)<sub>2</sub> (<b>2a</b>ā€“<b>c</b>, 28ā€“70%), which are treated with Grubbsā€™ catalyst (15 mol %; refluxing CH<sub>2</sub>Cl<sub>2</sub>). NMR analyses of the crude <i>inter</i>ligand metathesis products <i>trans</i>-FeĀ­(CO)<sub>3</sub>(PhPĀ­((CH<sub>2</sub>)<sub><i>m</i></sub>CHī—»CHĀ­(CH<sub>2</sub>)<sub><i>m</i></sub>)<sub>2</sub>PPh) (<b>3a</b>ā€“<b>c</b>, 30ā€“31%) suggest <i>Z</i>/<i>E</i> Cī—»C mixtures and/or byproducts from <i>intra</i>ligand metathesis or oligomers. Subsequent hydrogenations (5 bar/cat. RhĀ­(Cl)Ā­(PPh<sub>3</sub>)<sub>3</sub> or PtO<sub>2</sub>) afford <i>trans</i>-FeĀ­(CO)<sub>3</sub>(PhPĀ­((CH<sub>2</sub>)<sub><i>n</i></sub>)<sub>2</sub>PPh) (<b>4a</b>ā€“<b>c</b>, 69ā€“77%; <i>n</i> = 2<i>m</i> + 2, <i>synperiplanar</i> phenyl groups), which density functional theory calculations show to be more stable than isomers derived from other metathesis modes. Crystallizations give (<i>E</i>,<i>E</i>)-<b>3a</b> and <b>4b</b>, the X-ray structures of which are determined and analyzed. Variable-temperature <sup>13</sup>CĀ­{<sup>1</sup>H} NMR experiments show that rotation of the FeĀ­(CO)<sub>3</sub> moiety in <b>4b</b> is rapid on the NMR time scale (RT to 0 Ā°C; Ī”<i>G</i><sup>ā§§</sup><sub>273Ā K</sub> ā‰¤ 12.8 kcal/mol), but that in <b>4a</b> is not (RT to 105 Ā°C; Ī”<i>G</i><sup>ā§§</sup><sub>378Ā K</sub> ā‰„ 17.9 kcal/mol). These data indicate rotational barriers lower than those in analogues in which three methylene chains connect the phosphorus atoms, <i>trans</i>-FeĀ­(CO)<sub>3</sub>(PĀ­((CH<sub>2</sub>)<sub><i>n</i></sub>)<sub>3</sub>P)

    Polyyne Rotaxanes: Stabilization by Encapsulation

    No full text
    Active metal template Glaser coupling has been used to synthesize a series of rotaxanes consisting of a polyyne, with up to 24 contiguous <i>sp-</i>hybridized carbon atoms, threaded through a variety of macrocycles. Cadiotā€“Chodkiewicz cross-coupling affords higher yields of rotaxanes than homocoupling. This methodology has been used to prepare [3]Ā­rotaxanes with two polyyne chains locked through the same macrocycle. The crystal structure of one of these [3]Ā­rotaxanes shows that there is extremely close contact between the central carbon atoms of the threaded hexayne chains (CĀ·Ā·Ā·C distance 3.29 ƅ vs 3.4 ƅ for the sum of van der Waals radii) and that the bond-length-alternation is perturbed in the vicinity of this contact. However, despite the close interaction between the hexayne chains, the [3]Ā­rotaxane is remarkably stable under ambient conditions, probably because the two polyynes adopt a crossed geometry. In the solid state, the angle between the two polyyne chains is 74Ā°, and this crossed geometry appears to be dictated by the bulk of the ā€œsupertritylā€ end groups. Several rotaxanes have been synthesized to explore gem-dibromoethene moieties as ā€œmaskedā€ polyynes. However, the reductive Fritschā€“Buttenbergā€“Wiechell rearrangement to form the desired polyyne rotaxanes has not yet been achieved. X-ray crystallographic analysis on six [2]Ā­rotaxanes and two [3]Ā­rotaxanes provides insight into the noncovalent interactions in these systems. Differential scanning calorimetry (DSC) reveals that the longer polyyne rotaxanes (C16, C18, and C24) decompose at higher temperatures than the corresponding unthreaded polyyne axles. The stability enhancement increases as the polyyne becomes longer, reaching 60 Ā°C in the C24 rotaxane

    Gyroscope-Like Complexes Based on Dibridgehead Diphosphine Cages That Are Accessed by Three-Fold Intramolecular Ring Closing Metatheses and Encase Fe(CO)<sub>3</sub>, Fe(CO)<sub>2</sub>(NO)<sup>+</sup>, and Fe(CO)<sub>3</sub>(H)<sup>+</sup> Rotators

    No full text
    Reactions of <i>trans</i>-FeĀ­(CO)<sub>3</sub>(PĀ­((CH<sub>2</sub>)<sub><i>m</i></sub>CHī—»CH<sub>2</sub>)<sub>3</sub>)<sub>2</sub> (<i>m</i> = <b>a</b>/4; <b>b</b>/5, <b>c</b>/6, <b>e</b>/8) and Grubbsā€™ catalyst (12ā€“24 mol %, CH<sub>2</sub>Cl<sub>2</sub>, reflux) give the cage-like trienes <i>trans</i>-FeĀ­(CO)<sub>3</sub>(PĀ­((CH<sub>2</sub>)<sub><i>m</i></sub>CHī—»CHĀ­(CH<sub>2</sub>)<sub><i>m</i></sub>)<sub>3</sub>P) (<b>3a</b>ā€“<b>c</b>,<b>e</b>, 60ā€“81%). Hydrogenations (ClRhĀ­(PPh<sub>3</sub>)<sub>3</sub>, 60ā€“80 Ā°C) yield the title compounds <i>trans</i>-FeĀ­(CO)<sub>3</sub>(PĀ­((CH<sub>2</sub>)<sub><i>n</i></sub>)<sub>3</sub>P) (<b>4a</b>ā€“<b>c</b>,<b>e</b>, 74ā€“86%; <i>n</i> = 2<i>m</i> + 2), which have idealized <i>D</i><sub>3<i>h</i></sub> symmetry. A crystal structure of <b>4c</b> suggests enough van der Waals clearance for the FeĀ­(CO)<sub>3</sub> moiety to rotate within the three PĀ­(CH<sub>2</sub>)<sub>14</sub>P linkages; structures of <i>E</i>,<i>E</i>,<i>E</i>-<b>3a</b> show rotation to be blocked by the shorter PĀ­(CH<sub>2</sub>)<sub>4</sub>CHī—»CHĀ­(CH<sub>2</sub>)<sub>4</sub>P linkages. Additions of NO<sup>+</sup>BF<sub>4</sub><sup>ā€“</sup> give the isoelectronic and isosteric cations [FeĀ­(CO)<sub>2</sub>(NO)Ā­(PĀ­((CH<sub>2</sub>)<sub><i>n</i></sub>)<sub>3</sub>P)]<sup>+</sup>BF<sub>4</sub><sup>ā€“</sup> (<b>5a</b>ā€“<b>c</b><sup>+</sup>BF<sub>4</sub><sup>ā€“</sup>; 81ā€“98%). Additions of [HĀ­(OEt<sub>2</sub>)<sub>2</sub>]<sup>+</sup>BAr<sub>f</sub><sup>ā€“</sup> (BAr<sub>f</sub> = BĀ­(3,5-C<sub>6</sub>H<sub>3</sub>(CF<sub>3</sub>)<sub>2</sub>)<sub>4</sub>) afford the metal hydride complexes <i>mer</i>,<i>trans</i>-[FeĀ­(CO)<sub>3</sub>(H)Ā­(PĀ­((CH<sub>2</sub>)<sub><i>n</i></sub>)<sub>3</sub>P)]<sup>+</sup>BAr<sub>f</sub><sup>ā€“</sup> (<b>6a</b>ā€“<b>c</b>,<b>e</b><sup>+</sup>BAr<sub>f</sub><sup>ā€“</sup>; 98ā€“99%). The behavior of the rotators in the preceding complexes is probed by VT NMR. At ambient temperature in solution, <b>5a</b>,<b>b</b><sup>+</sup>BF<sub>4</sub><sup>ā€“</sup> and <b>6a</b><sup>+</sup>BAr<sub>f</sub><sup>ā€“</sup> show two sets of PĀ­(CH<sub>2</sub>)<sub><i>n</i>/2</sub> <sup>13</sup>C NMR signals (2:1), whereas <b>5c</b><sup>+</sup>BF<sub>4</sub><sup>ā€“</sup> and <b>6b</b>,<b>c</b><sup>+</sup>BAr<sub>f</sub><sup>ā€“</sup> show only one. At higher temperatures, the signals of <b>5b</b><sup>+</sup>BF<sub>4</sub><sup>ā€“</sup> coalesce; at lower temperatures, those of <b>5c</b><sup>+</sup>BF<sub>4</sub><sup>ā€“</sup> and <b>6b</b><sup>+</sup>BAr<sub>f</sub><sup>ā€“</sup> decoalesce. These data give Ī”<i>H</i><sup>ā§§</sup>/Ī”<i>S</i><sup>ā§§</sup> values (kcal/mol and eu) of 8.3/ā€“28.4 and 9.5/ā€“6.5 for FeĀ­(CO)<sub>2</sub>(NO)<sup>+</sup> rotation (<b>5b</b>,<b>c</b><sup>+</sup>) and 6.1/ā€“23.5 for FeĀ­(CO)<sub>3</sub>(H)<sup>+</sup> rotation (<b>6b</b><sup>+</sup>). <sup>13</sup>C CP/MAS NMR spectra show that the FeĀ­(CO)<sub>3</sub> moiety in polycrystalline <b>4c</b> (but not <b>4a</b>) undergoes rapid rotation between āˆ’60 and 95 Ā°C. Approaches to minimizing these barriers and developing molecular gyroscopes are discussed

    Legislative Documents

    No full text
    Also, variously referred to as: House bills; House documents; House legislative documents; legislative documents; General Court documents
    corecore