5 research outputs found

    Pathways to the Polymerization of Boron Monoxide Dimer To Give Low-Density Porous Materials Containing Six-Membered Boroxine Rings

    No full text
    Density functional theory has been used to examine the key mechanistic details of the polymerization of boron monoxide (BO) via its OB–BO dimer to give ultimately low-density porous polymeric (BO)<sub><i>n</i></sub> materials. The structures of such materials consist of planar layers of six-membered boroxine (B<sub>3</sub>O<sub>3</sub>) rings linked by boron–boron bonds. Initial cyclooligomerization of B<sub>2</sub>O<sub>2</sub> leads to a B<sub>4</sub>O<sub>4</sub> dimer with a four-membered B<sub>2</sub>O<sub>2</sub> ring, a B<sub>6</sub>O<sub>6</sub> trimer containing a six-membered B<sub>3</sub>O<sub>3</sub> (boroxine) ring, a B<sub>8</sub>O<sub>8</sub> tetramer containing an eight-membered B<sub>4</sub>O<sub>4</sub> ring, and even a B<sub>10</sub>O<sub>10</sub> pentamer containing a ten-membered B<sub>5</sub>O<sub>5</sub> ring. However, an isomeric B<sub>10</sub>O<sub>10</sub> structure containing two boroxine rings linked by a B–B bond is a much lower energy structure by ∼31 kcal/mol owing to the special stability of the aromatic boroxine rings. Rotation of the boroxine rings around the central B–B bond in this B<sub>10</sub>O<sub>10</sub> structure has a low rotation barrier suggesting that further oligomerization to give products containing either perpendicular or planar orientations of the B<sub>3</sub>O<sub>3</sub> rings is possible. However, the planar oligomers are energetically more favorable since they have fewer high-energy external BO groups bonded to the network of boroxine rings. The pendant boronyl groups are reactive sites that can be used for further polymerization. Mechanistic aspects of the further oligomerization of (BO)<sub><i>x</i></sub> systems to give a B<sub>24</sub>O<sub>24</sub> oligomer with a naphthalene-like arrangement of boroxine rings and a B<sub>84</sub>O<sub>84</sub> structure with a coronene-like arrangement of boroxine rings have been examined. Further polymerization of these intermediates by similar processes is predicted to lead ultimately to polymers consisting of planar networks of boroxine rings. The holes between the boroxine rings in such polymers suggests that they will be porous low-density materials. Applications of such materials as absorbents for small molecules are anticipated

    The Quest for Metal–Metal Quadruple and Quintuple Bonds in Metal Carbonyl Derivatives: Nb<sub>2</sub>(CO)<sub>9</sub> and Nb<sub>2</sub>(CO)<sub>8</sub>

    No full text
    The synthesis by Power and co-workers of the first metal–metal quintuple bond (<i>Science</i> <b>2005</b>, <i>310</i>, 844) is a landmark in inorganic chemistry. The 18-electron rule suggests that Nb<sub>2</sub>(CO)<sub>9</sub> and Nb<sub>2</sub>(CO)<sub>8</sub> are candidates for binary metal carbonyls containing metal–metal quadruple and quintuple bonds, respectively. Density functional theory (MPW1PW91 and BP86) indeed predicts structures having very short Nb–Nb distances of ∼2.5 Å for Nb<sub>2</sub>(CO)<sub>9</sub> and ∼2.4 Å for Nb<sub>2</sub>(CO)<sub>8</sub> as well as relatively large Nb–Nb Wiberg bond indices supporting these high formal Nb–Nb bond orders. However, analysis of the frontier molecular orbitals of these unbridged structures suggests formal NbNb triple bonds and 16-electron metal configurations. This contrasts with an analysis of the frontier orbitals in a model chromium­(I) alkyl linear CH<sub>3</sub>CrCrCH<sub>3</sub>, which confirms the generally accepted presence of chromium–chromium quintuple bonds in such molecules. The presence of NbNb triple bonds rather than quadruple or quintuple bonds in the Nb<sub>2</sub>(CO)<sub><i>n</i></sub> (<i>n</i> = 9, 8) structures frees up d­(<i>xy</i>) and d­(<i>x</i><sup>2</sup>–<i>y</i><sup>2</sup>) orbitals for dπ→pπ* back-bonding to the carbonyl groups. The lowest energy Nb<sub>2</sub>(CO)<sub><i>n</i></sub> structures (<i>n</i> = 9, 8) are not these unbridged structures but structures having bridging carbonyl groups of various types and formal Nb–Nb orders no higher than three. Thus, the two lowest energy Nb<sub>2</sub>(CO)<sub>9</sub> structures have NbNb triple bond distances of ∼2.8 Å and three semibridging carbonyl groups, leading to a 16-electron configuration rather than an 18-electron configuration for one of the niobium atoms. The lowest energy structure of the highly unsaturated Nb<sub>2</sub>(CO)<sub>8</sub> is unusual since it has a formal <i>single</i> Nb–Nb bond of length ∼3.1 Å and two four-electron donor η<sup>2</sup>-μ-CO groups, thereby giving each niobium atom only a 16-electron configuration

    Construction of the Tetrahedral Trifluorophosphine Platinum Cluster Pt<sub>4</sub>(PF<sub>3</sub>)<sub>8</sub> from Smaller Building Blocks

    No full text
    The experimentally known but structurally uncharacterized Pt<sub>4</sub>(PF<sub>3</sub>)<sub>8</sub> is predicted to have an <i>S</i><sub>4</sub> structure with a central distorted Pt<sub>4</sub> tetrahedron having four short PtPt distances, two long Pt–Pt distances, and all terminal PF<sub>3</sub> groups. The structures of the lower nuclearity species Pt­(PF<sub>3</sub>)<sub><i>n</i></sub> (<i>n</i> = 4, 3, 2), Pt<sub>2</sub>(PF<sub>3</sub>)<sub><i>n</i></sub> (<i>n</i> = 7, 6, 5, 4), and Pt<sub>3</sub>(PF<sub>3</sub>)<sub>6</sub> were investigated by density functional theory to assess their possible roles as intermediates in the formation of Pt<sub>4</sub>(PF<sub>3</sub>)<sub>8</sub> by the pyrolysis of Pt­(PF<sub>3</sub>)<sub>4</sub>. The expected tetrahedral, trigonal planar, and linear structures are found for Pt­(PF<sub>3</sub>)<sub>4</sub>, Pt­(PF<sub>3</sub>)<sub>3</sub>, and Pt­(PF<sub>3</sub>)<sub>2</sub>, respectively. However, the dicoordinate Pt­(PF<sub>3</sub>)<sub>2</sub> structure is bent from the ideal 180° linear structure to approximately 160°. Most of the low-energy binuclear Pt<sub>2</sub>(PF<sub>3</sub>)<sub><i>n</i></sub> (<i>n</i> = 7, 6, 5) structures can be derived from the mononuclear Pt­(PF<sub>3</sub>)<sub><i>n</i></sub> (<i>n</i> = 4, 3, 2) structures by replacing one of the PF<sub>3</sub> groups by a Pt­(PF<sub>3</sub>)<sub>4</sub> or Pt­(PF<sub>3</sub>)<sub>3</sub> ligand. In some of these binuclear structures one of the PF<sub>3</sub> groups on the Pt­(PF<sub>3</sub>)<sub><i>n</i></sub> ligand becomes a bridging group. The low-energy binuclear structures also include symmetrical [Pt­(PF<sub>3</sub>)<sub><i>n</i></sub>]<sub>2</sub> dimers (<i>n</i> = 2, 3) of the coordinately unsaturated Pt­(PF<sub>3</sub>)<sub><i>n</i></sub> (<i>n</i> = 3, 2). The four low-energy structures for the trinuclear Pt<sub>3</sub>(PF<sub>3</sub>)<sub>6</sub> include two structures with central equilateral Pt<sub>3</sub> triangles and two structures with isosceles Pt<sub>3</sub> triangles and various arrangements of terminal and bridging PF<sub>3</sub> groups. Among these four structures the lowest-energy Pt<sub>3</sub>(PF<sub>3</sub>)<sub>6</sub> structure has an unprecedented four-electron donor η<sup>2</sup>-μ<sub>3</sub>-PF<sub>3</sub> group bridging the central Pt<sub>3</sub> triangle through three Pt–P bonds and one Pt–F bond. Thermochemical studies on the aggregation of these Pt-PF<sub>3</sub> complexes suggest the tetramerization of Pt­(PF<sub>3</sub>)<sub>2</sub> to Pt<sub>4</sub>(PF<sub>3</sub>)<sub>8</sub> to be highly exothermic regardless of the mechanistic details

    Cyclization of Thiocarbonyl Groups in Binuclear Homoleptic Nickel Thiocarbonyls To Give Ligands Derived from Sulfur Analogues of Croconic and Rhodizonic Acids

    No full text
    The sulfur analogue of the well-known Ni­(CO)<sub>4</sub>, namely, Ni­(CS)<sub>4</sub>, has been observed spectroscopically in low temperature matrices but is not known as a stable species under ambient conditions. Theoretical studies show that Ni­(CS)<sub>4</sub> with monomeric CS ligands and tetrahedrally coordinated nickel is disfavored by ∼17 kcal/mol relative to unusual isomeric Ni­(C<sub>2</sub>S<sub>2</sub>)<sub>2</sub> structures. In the latter structures the CS ligands couple pairwise through C–C bond formation to give dimeric SCCS ligands, which bond preferentially to the nickel atom through their CS bonds rather than their CC bonds. Coupling of CS ligands in the lowest energy binuclear Ni<sub>2</sub>(CS)<sub><i>n</i></sub> (<i>n</i> = 7, 6, 5) structures results in cyclization to give remarkable C<sub><i>n</i></sub>S<sub><i>n</i></sub> (<i>n</i> = 5, 6) ligands containing five- and six-membered carbocyclic rings. Such ligands, which are the sulfur analogues of the well-known croconate (<i>n</i> = 5) and rhodizonate (<i>n</i> = 6) oxocarbon ligands, function as bidentate ligands to the central Ni<sub>2</sub> unit. Higher energy Ni<sub>2</sub>(CS)<sub><i>n</i></sub> (<i>n</i> = 7, 6, 5) structures contain dimeric C<sub>2</sub>S<sub>2</sub> ligands, which can bridge the central Ni<sub>2</sub> unit. Dimeric C<sub>2</sub>S<sub>2</sub> ligands rather than tetrathiosquare C<sub>4</sub>S<sub>4</sub> ligands are found in the lowest energy Ni<sub>2</sub>(CS)<sub>4</sub> structures

    Binuclear Cyclopentadienylmetal Methylene Sulfur Dioxide Complexes of Rhodium and Iridium Related to a Photochromic Metal Dithionite Complex

    No full text
    The photochromic dithionite complex Cp*<sub>2</sub>Rh<sub>2</sub>(μ-CH<sub>2</sub>)<sub>2</sub>(μ-O<sub>2</sub>SSO<sub>2</sub>) (Cp* = η<sup>5</sup>-Me<sub>5</sub>C<sub>5</sub>) is of interest because it undergoes an unusual fully reversible unimolecular photochemical rearrangement to the isodithionite complex Cp*<sub>2</sub>Rh<sub>2</sub>(μ-CH<sub>2</sub>)<sub>2</sub>(μ-O<sub>2</sub>SOSO). In order to obtain more insight into these systems, a comprehensive density functional theory study has been carried out on isomeric Cp<sub>2</sub>M<sub>2</sub>(CH<sub>2</sub>)<sub>2</sub>(SO<sub>2</sub>)<sub>2</sub> (M = Rh, Ir) derivatives. The experimentally observed rhodium complexes with coupled sulfur dioxide (SO<sub>2</sub>) units to give dithionite or isodithionite ligands are surprisingly high-energy kinetic isomers in our analysis, reflecting the need for dithionite rather than SO<sub>2</sub> for their synthesis. Many isomeric structures containing two separate SO<sub>2</sub> ligands are found to lie at lower energies than these dithionite and isodithionite complexes. In the lowest-energy Cp<sub>2</sub>M<sub>2</sub>(CH<sub>2</sub>)<sub>2</sub>(SO<sub>2</sub>)<sub>2</sub> isomers, the two methylene groups couple to form an ethylene ligand that can be either terminal or bridging. In slightly higher energy structures, a formal hydrogen shift is predicted to occur within the ethylene ligand to give a methylcarbene CH<sub>3</sub>CH ligand. Isomers with a bridging methylcarbene ligand are energetically preferred over isomers with a terminal methylcarbene ligand. Generation of the lower-energy Cp<sub>2</sub>Rh<sub>2</sub>(CH<sub>2</sub>)<sub>2</sub>(SO<sub>2</sub>)<sub>2</sub> isomers containing separate SO<sub>2</sub> ligands should be achievable through reactions of SO<sub>2</sub> with more highly reduced cyclopentadienylrhodium methylene complexes such as Cp*<sub>2</sub>Rh<sub>2</sub>(μ-CH<sub>2</sub>)<sub>2</sub>
    corecore