10 research outputs found

    Thermodynamic, Kinetic, Structural, and Computational Studies of the Ph<sub>3</sub>Snā€“H, Ph<sub>3</sub>Snā€“SnPh<sub>3</sub>, and Ph<sub>3</sub>Snā€“Cr(CO)<sub>3</sub>C<sub>5</sub>Me<sub>5</sub> Bond Dissociation Enthalpies

    No full text
    The kinetics of the reaction of Ph<sub>3</sub>SnH with excess ā€¢CrĀ­(CO)<sub>3</sub>C<sub>5</sub>Me<sub>5</sub> = ā€¢<b>Cr</b>, producing H<b>Cr</b> and Ph<sub>3</sub>Snā€“<b>Cr</b>, was studied in toluene solution under 2ā€“3 atm CO pressure in the temperature range of 17ā€“43.5 Ā°C. It was found to obey the rate equation <i>d</i>[Ph<sub>3</sub>Snā€“<b>Cr</b>]/<i>d</i>t = <i>k</i>[Ph<sub>3</sub>SnH]Ā­[ā€¢<b>Cr</b>] and exhibit a normal kinetic isotope effect (<i>k</i><sub>H</sub>/<i>k</i><sub>D</sub> = 1.12 Ā± 0.04). Variable-temperature studies yielded Ī”<i>H</i><sup>ā€”</sup> = 15.7 Ā± 1.5 kcal/mol and Ī”<i>S</i><sup>ā€”</sup> = āˆ’11 Ā± 5 cal/(molĀ·K) for the reaction. These data are interpreted in terms of a two-step mechanism involving a thermodynamically uphill hydrogen atom transfer (HAT) producing Ph<sub>3</sub>Snā€¢ and H<b>Cr</b>, followed by rapid trapping of Ph<sub>3</sub>Snā€¢ by excess ā€¢<b>Cr</b> to produce Ph<sub>3</sub>Snā€“<b>Cr</b>. Assuming an overbarrier of 2 Ā± 1 kcal/mol in the HAT step leads to a derived value of 76.0 Ā± 3.0 kcal/mol for the Ph<sub>3</sub>Snā€“H bond dissociation enthalpy (BDE) in toluene solution. The reaction enthalpy of Ph<sub>3</sub>SnH with excess ā€¢<b>Cr</b> was measured by reaction calorimetry in toluene solution, and a value of the Snā€“Cr BDE in Ph<sub>3</sub>Sn-<b>Cr</b> of 50.4 Ā± 3.5 kcal/mol was derived. Qualitative studies of the reactions of other R<sub>3</sub>SnH compounds with ā€¢<b>Cr</b> are described for R = <sup>n</sup>Bu, <sup>t</sup>Bu, and Cy. The dehydrogenation reaction of 2Ph<sub>3</sub>SnH ā†’ H<sub>2</sub> + Ph<sub>3</sub>SnSnPh<sub>3</sub> was found to be rapid and quantitative in the presence of catalytic amounts of the complex PdĀ­(IPr)Ā­(PĀ­(<i>p</i>-tolyl)<sub>3</sub>). The thermochemistry of this process was also studied in toluene solution using varying amounts of the Pd(0) catalyst. The value of Ī”<i>H</i> = āˆ’15.8 Ā± 2.2 kcal/mol yields a value of the Snā€“Sn BDE in Ph<sub>3</sub>SnSnPh<sub>3</sub> of 63.8 Ā± 3.7 kcal/mol. Computational studies of the Snā€“H, Snā€“Sn, and Snā€“Cr BDEs are in good agreement with experimental data and provide additional insight into factors controlling reactivity in these systems. The structures of Ph<sub>3</sub>Snā€“<b>Cr</b> and Cy<sub>3</sub>Snā€“<b>Cr</b> were determined by X-ray crystallography and are reported. Mechanistic aspects of oxidative addition reactions in this system are discussed

    High Quantum Yield Molecular Bromine Photoelimination from Mononuclear Platinum(IV) Complexes

    No full text
    PtĀ­(IV) complexes <i>trans</i>-PtĀ­(PEt<sub>3</sub>)<sub>2</sub>(R)Ā­(Br)<sub>3</sub> (R = Br, aryl and polycyclic aromatic fragments) photoeliminate molecular bromine with quantum yields as high as 82%. Photoelimination occurs both in the solid state and in solution. Calorimetry measurements and DFT calculations (PMe<sub>3</sub> analogs) indicate endothermic and endergonic photoeliminations with free energies from 2 to 22 kcal/mol of Br<sub>2</sub>. Solution trapping experiments with high concentrations of 2,3-dimethyl-2-butene suggest a radical-like excited state precursor to bromine elimination

    High Quantum Yield Molecular Bromine Photoelimination from Mononuclear Platinum(IV) Complexes

    No full text
    PtĀ­(IV) complexes <i>trans</i>-PtĀ­(PEt<sub>3</sub>)<sub>2</sub>(R)Ā­(Br)<sub>3</sub> (R = Br, aryl and polycyclic aromatic fragments) photoeliminate molecular bromine with quantum yields as high as 82%. Photoelimination occurs both in the solid state and in solution. Calorimetry measurements and DFT calculations (PMe<sub>3</sub> analogs) indicate endothermic and endergonic photoeliminations with free energies from 2 to 22 kcal/mol of Br<sub>2</sub>. Solution trapping experiments with high concentrations of 2,3-dimethyl-2-butene suggest a radical-like excited state precursor to bromine elimination

    Two-Step Binding of O<sub>2</sub> to a Vanadium(III) Trisanilide Complex To Form a Non-Vanadyl Vanadium(V) Peroxo Complex

    No full text
    Treatment of VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>1</b>) (Ar = 3,5-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>) with O<sub>2</sub> was shown by stopped-flow kinetic studies to result in the rapid formation of (Ī·<sup>1</sup>-O<sub>2</sub>)Ā­VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>2</b>) (Ī”<i>H</i><sup>ā§§</sup> = 3.3 Ā± 0.2 kcal/mol and Ī”<i>S</i><sup>ā§§</sup> = āˆ’22 Ā± 1 cal mol<sup>ā€“1</sup> K<sup>ā€“1</sup>), which subsequently isomerizes to (Ī·<sup>2</sup>-O<sub>2</sub>)Ā­VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>3</b>) (Ī”<i>H</i><sup>ā§§</sup> = 10.3 Ā± 0.9 kcal/mol and Ī”<i>S</i><sup>ā§§</sup> = āˆ’6 Ā± 4 cal mol<sup>ā€“1</sup> K<sup>ā€“1</sup>). The enthalpy of binding of O<sub>2</sub> to form <b>3</b> is āˆ’75.0 Ā± 2.0 kcal/mol, as measured by solution calorimetry. The reaction of <b>3</b> and <b>1</b> to form 2 equiv of Oī—¼VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>4</b>) occurs by initial isomerization of <b>3</b> to <b>2</b>. The results of computational studies of this rearrangement (Ī”<i>H</i> = 4.2 kcal/mol; Ī”<i>H</i><sup>ā§§</sup> = 16 kcal/mol) are in accord with experimental data (Ī”<i>H</i> = 4 Ā± 3 kcal/mol; Ī”<i>H</i><sup>ā§§</sup> = 14 Ā± 3 kcal/mol). With the aim of suppressing the formation of <b>4</b>, the reaction of O<sub>2</sub> with <b>1</b> in the presence of <sup><i>t</i></sup>BuCN was studied. At āˆ’45 Ā°C, the principal products of this reaction are <b>3</b> and <sup><i>t</i></sup>BuCĀ­(ī—»O)Ā­Nī—¼VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>5</b>), in which the bound nitrile has been oxidized. Crystal structures of <b>3</b> and <b>5</b> are reported

    Role of Axial Base Coordination in Isonitrile Binding and Chalcogen Atom Transfer to Vanadium(III) Complexes

    No full text
    The enthalpy of oxygen atom transfer (OAT) to VĀ­[(Me<sub>3</sub>SiNCH<sub>2</sub>CH<sub>2</sub>)<sub>3</sub>N], <b>1</b>, forming OVĀ­[(Me<sub>3</sub>SiNCH<sub>2</sub>CH<sub>2</sub>)<sub>3</sub>N], <b>1</b>ā€“O, and the enthalpies of sulfur atom transfer (SAT) to <b>1</b> and VĀ­(NĀ­[<i>t</i>-Bu]Ā­Ar)<sub>3</sub>, <b>2</b> (Ar = 3,5-C<sub>6</sub>H<sub>3</sub>Me<sub>2</sub>), forming the corresponding sulfides SVĀ­[(Me<sub>3</sub>SiNCH<sub>2</sub>CH<sub>2</sub>)<sub>3</sub>N], <b>1</b>ā€“S, and SVĀ­(NĀ­[<i>t</i>-Bu]Ā­Ar)<sub>3</sub>, <b>2</b>ā€“S, have been measured by solution calorimetry in toluene solution using dbabhNO (dbabhNO = 7-nitroso-2,3:5,6-dibenzo-7-azabicyclo[2.2.1]Ā­hepta-2,5-diene) and Ph<sub>3</sub>SbS as chalcogen atom transfer reagents. The Vā€“O BDE in <b>1</b>ā€“O is 6.3 Ā± 3.2 kcalĀ·mol<sup>ā€“1</sup> lower than the previously reported value for <b>2</b>ā€“O and the Vā€“S BDE in <b>1</b>ā€“S is 3.3 Ā± 3.1 kcalĀ·mol<sup>ā€“1</sup> lower than that in <b>2</b>ā€“S. These differences are attributed primarily to a weakening of the Vā€“N<sub>axial</sub> bond present in complexes of <b>1</b> upon oxidation. The rate of reaction of <b>1</b> with dbabhNO has been studied by low temperature stopped-flow kinetics. Rate constants for OAT are over 20 times greater than those reported for <b>2</b>. Adamantyl isonitrile (AdNC) binds rapidly and quantitatively to both <b>1</b> and <b>2</b> forming high spin adducts of VĀ­(III). The enthalpies of ligand addition to <b>1</b> and <b>2</b> in toluene solution are āˆ’19.9 Ā± 0.6 and āˆ’17.1 Ā± 0.7 kcalĀ·mol<sup>ā€“1</sup>, respectively. The more exothermic ligand addition to <b>1</b> as compared to <b>2</b> is opposite to what was observed for OAT and SAT. This is attributed to less weakening of the Vā€“N<sub>axial</sub> bond in ligand binding as opposed to chalcogen atom transfer and is in keeping with structural data and computations. The structures of <b>1</b>, <b>1</b>ā€“O, <b>1</b>ā€“S, <b>1</b>ā€“CNAd, and <b>2</b>ā€“CNAd have been determined by X-ray crystallography and are reported

    Two-Step Binding of O<sub>2</sub> to a Vanadium(III) Trisanilide Complex To Form a Non-Vanadyl Vanadium(V) Peroxo Complex

    No full text
    Treatment of VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>1</b>) (Ar = 3,5-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>) with O<sub>2</sub> was shown by stopped-flow kinetic studies to result in the rapid formation of (Ī·<sup>1</sup>-O<sub>2</sub>)Ā­VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>2</b>) (Ī”<i>H</i><sup>ā§§</sup> = 3.3 Ā± 0.2 kcal/mol and Ī”<i>S</i><sup>ā§§</sup> = āˆ’22 Ā± 1 cal mol<sup>ā€“1</sup> K<sup>ā€“1</sup>), which subsequently isomerizes to (Ī·<sup>2</sup>-O<sub>2</sub>)Ā­VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>3</b>) (Ī”<i>H</i><sup>ā§§</sup> = 10.3 Ā± 0.9 kcal/mol and Ī”<i>S</i><sup>ā§§</sup> = āˆ’6 Ā± 4 cal mol<sup>ā€“1</sup> K<sup>ā€“1</sup>). The enthalpy of binding of O<sub>2</sub> to form <b>3</b> is āˆ’75.0 Ā± 2.0 kcal/mol, as measured by solution calorimetry. The reaction of <b>3</b> and <b>1</b> to form 2 equiv of Oī—¼VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>4</b>) occurs by initial isomerization of <b>3</b> to <b>2</b>. The results of computational studies of this rearrangement (Ī”<i>H</i> = 4.2 kcal/mol; Ī”<i>H</i><sup>ā§§</sup> = 16 kcal/mol) are in accord with experimental data (Ī”<i>H</i> = 4 Ā± 3 kcal/mol; Ī”<i>H</i><sup>ā§§</sup> = 14 Ā± 3 kcal/mol). With the aim of suppressing the formation of <b>4</b>, the reaction of O<sub>2</sub> with <b>1</b> in the presence of <sup><i>t</i></sup>BuCN was studied. At āˆ’45 Ā°C, the principal products of this reaction are <b>3</b> and <sup><i>t</i></sup>BuCĀ­(ī—»O)Ā­Nī—¼VĀ­(NĀ­[<sup><i>t</i></sup>Bu]Ā­Ar)<sub>3</sub> (<b>5</b>), in which the bound nitrile has been oxidized. Crystal structures of <b>3</b> and <b>5</b> are reported

    Thermodynamic and Kinetic Study of Cleavage of the Nā€“O Bond of Nā€‘Oxides by a Vanadium(III) Complex: Enhanced Oxygen Atom Transfer Reaction Rates for Adducts of Nitrous Oxide and Mesityl Nitrile Oxide

    No full text
    Thermodynamic, kinetic, and computational studies are reported for oxygen atom transfer (OAT) to the complex VĀ­(NĀ­[<i>t</i>-Bu]Ā­Ar)<sub>3</sub> (Ar = 3,5-C<sub>6</sub>H<sub>3</sub>Me<sub>2</sub>, <b>1</b>) from compounds containing Nā€“O bonds with a range of BDEs spanning nearly 100 kcal mol<sup>ā€“1</sup>: PhNO (108) > SIPr/MesCNO (75) > PyO (63) > IPr/N<sub>2</sub>O (62) > MesCNO (53) > N<sub>2</sub>O (40) > dbabhNO (10) (Mes = mesityl; SIPr = 1,3-bisĀ­(diisopropyl)Ā­phenylimidazolin-2-ylidene; Py = pyridine; IPr = 1,3-bisĀ­(diisopropyl)Ā­phenylimidazol-2-ylidene; dbabh = 2,3:5,6-dibenzo-7-azabicyclo[2.2.1]Ā­hepta-2,5-diene). Stopped flow kinetic studies of the OAT reactions show a range of kinetic behavior influenced by both the mode and strength of coordination of the O donor and its ease of atom transfer. Four categories of kinetic behavior are observed depending upon the magnitudes of the rate constants involved: (I) dinuclear OAT following an overall third order rate law (N<sub>2</sub>O); (II) formation of stable oxidant-bound complexes followed by OAT in a separate step (PyO and PhNO); (III) transient formation and decay of metastable oxidant-bound intermediates on the same time scale as OAT (SIPr/MesCNO and IPr/N<sub>2</sub>O); (IV) steady-state kinetics in which no detectable intermediates are observed (dbabhNO and MesCNO). Thermochemical studies of OAT to <b>1</b> show that the Vā€“O bond in Oī—¼VĀ­(NĀ­[<i>t</i>-Bu]Ā­Ar)<sub>3</sub> is strong (BDE = 154 Ā± 3 kcal mol<sup>ā€“1</sup>) compared with all the Nā€“O bonds cleaved. In contrast, measurement of the Nā€“O bond in dbabhNO show it to be especially weak (BDE = 10 Ā± 3 kcal mol<sup>ā€“1</sup>) and that dissociation of dbabhNO to anthracene, N<sub>2</sub>, and a <sup>3</sup>O atom is thermodynamically favorable at room temperature. Comparison of the OAT of adducts of N<sub>2</sub>O and MesCNO to the bulky complex <b>1</b> show a faster rate than in the case of free N<sub>2</sub>O or MesCNO despite increased steric hindrance of the adducts

    Thermodynamic and Kinetic Study of Cleavage of the Nā€“O Bond of Nā€‘Oxides by a Vanadium(III) Complex: Enhanced Oxygen Atom Transfer Reaction Rates for Adducts of Nitrous Oxide and Mesityl Nitrile Oxide

    No full text
    Thermodynamic, kinetic, and computational studies are reported for oxygen atom transfer (OAT) to the complex VĀ­(NĀ­[<i>t</i>-Bu]Ā­Ar)<sub>3</sub> (Ar = 3,5-C<sub>6</sub>H<sub>3</sub>Me<sub>2</sub>, <b>1</b>) from compounds containing Nā€“O bonds with a range of BDEs spanning nearly 100 kcal mol<sup>ā€“1</sup>: PhNO (108) > SIPr/MesCNO (75) > PyO (63) > IPr/N<sub>2</sub>O (62) > MesCNO (53) > N<sub>2</sub>O (40) > dbabhNO (10) (Mes = mesityl; SIPr = 1,3-bisĀ­(diisopropyl)Ā­phenylimidazolin-2-ylidene; Py = pyridine; IPr = 1,3-bisĀ­(diisopropyl)Ā­phenylimidazol-2-ylidene; dbabh = 2,3:5,6-dibenzo-7-azabicyclo[2.2.1]Ā­hepta-2,5-diene). Stopped flow kinetic studies of the OAT reactions show a range of kinetic behavior influenced by both the mode and strength of coordination of the O donor and its ease of atom transfer. Four categories of kinetic behavior are observed depending upon the magnitudes of the rate constants involved: (I) dinuclear OAT following an overall third order rate law (N<sub>2</sub>O); (II) formation of stable oxidant-bound complexes followed by OAT in a separate step (PyO and PhNO); (III) transient formation and decay of metastable oxidant-bound intermediates on the same time scale as OAT (SIPr/MesCNO and IPr/N<sub>2</sub>O); (IV) steady-state kinetics in which no detectable intermediates are observed (dbabhNO and MesCNO). Thermochemical studies of OAT to <b>1</b> show that the Vā€“O bond in Oī—¼VĀ­(NĀ­[<i>t</i>-Bu]Ā­Ar)<sub>3</sub> is strong (BDE = 154 Ā± 3 kcal mol<sup>ā€“1</sup>) compared with all the Nā€“O bonds cleaved. In contrast, measurement of the Nā€“O bond in dbabhNO show it to be especially weak (BDE = 10 Ā± 3 kcal mol<sup>ā€“1</sup>) and that dissociation of dbabhNO to anthracene, N<sub>2</sub>, and a <sup>3</sup>O atom is thermodynamically favorable at room temperature. Comparison of the OAT of adducts of N<sub>2</sub>O and MesCNO to the bulky complex <b>1</b> show a faster rate than in the case of free N<sub>2</sub>O or MesCNO despite increased steric hindrance of the adducts

    Thermodynamic and Kinetic Study of Cleavage of the Nā€“O Bond of Nā€‘Oxides by a Vanadium(III) Complex: Enhanced Oxygen Atom Transfer Reaction Rates for Adducts of Nitrous Oxide and Mesityl Nitrile Oxide

    No full text
    Thermodynamic, kinetic, and computational studies are reported for oxygen atom transfer (OAT) to the complex VĀ­(NĀ­[<i>t</i>-Bu]Ā­Ar)<sub>3</sub> (Ar = 3,5-C<sub>6</sub>H<sub>3</sub>Me<sub>2</sub>, <b>1</b>) from compounds containing Nā€“O bonds with a range of BDEs spanning nearly 100 kcal mol<sup>ā€“1</sup>: PhNO (108) > SIPr/MesCNO (75) > PyO (63) > IPr/N<sub>2</sub>O (62) > MesCNO (53) > N<sub>2</sub>O (40) > dbabhNO (10) (Mes = mesityl; SIPr = 1,3-bisĀ­(diisopropyl)Ā­phenylimidazolin-2-ylidene; Py = pyridine; IPr = 1,3-bisĀ­(diisopropyl)Ā­phenylimidazol-2-ylidene; dbabh = 2,3:5,6-dibenzo-7-azabicyclo[2.2.1]Ā­hepta-2,5-diene). Stopped flow kinetic studies of the OAT reactions show a range of kinetic behavior influenced by both the mode and strength of coordination of the O donor and its ease of atom transfer. Four categories of kinetic behavior are observed depending upon the magnitudes of the rate constants involved: (I) dinuclear OAT following an overall third order rate law (N<sub>2</sub>O); (II) formation of stable oxidant-bound complexes followed by OAT in a separate step (PyO and PhNO); (III) transient formation and decay of metastable oxidant-bound intermediates on the same time scale as OAT (SIPr/MesCNO and IPr/N<sub>2</sub>O); (IV) steady-state kinetics in which no detectable intermediates are observed (dbabhNO and MesCNO). Thermochemical studies of OAT to <b>1</b> show that the Vā€“O bond in Oī—¼VĀ­(NĀ­[<i>t</i>-Bu]Ā­Ar)<sub>3</sub> is strong (BDE = 154 Ā± 3 kcal mol<sup>ā€“1</sup>) compared with all the Nā€“O bonds cleaved. In contrast, measurement of the Nā€“O bond in dbabhNO show it to be especially weak (BDE = 10 Ā± 3 kcal mol<sup>ā€“1</sup>) and that dissociation of dbabhNO to anthracene, N<sub>2</sub>, and a <sup>3</sup>O atom is thermodynamically favorable at room temperature. Comparison of the OAT of adducts of N<sub>2</sub>O and MesCNO to the bulky complex <b>1</b> show a faster rate than in the case of free N<sub>2</sub>O or MesCNO despite increased steric hindrance of the adducts

    Ligand-Directed Reactivity in Dioxygen and Water Binding to <i>cis</i>-[Pd(NHC)<sub>2</sub>(Ī·<sup>2</sup>ā€‘O<sub>2</sub>)]

    No full text
    Reaction of [PdĀ­(IPr)<sub>2</sub>] (IPr = 1,3-bisĀ­(2,6-diisopropylphenyl)Ā­imidazol-2-ylidene) and O<sub>2</sub> leads to the surprising discovery that at low temperature the initial reaction product is a highly labile peroxide complex <i>cis</i>-[PdĀ­(IPr)<sub>2</sub>(Ī·<sup>2</sup>-O<sub>2</sub>)]. At temperatures ā‰³ āˆ’40 Ā°C, <i>cis</i>-[PdĀ­(IPr)<sub>2</sub>(Ī·<sup>2</sup>-O<sub>2</sub>)] adds a second O<sub>2</sub> to form <i>trans</i>-[PdĀ­(IPr)<sub>2</sub>(Ī·<sup>1</sup>-O<sub>2</sub>)<sub>2</sub>]. Squid magnetometry and EPR studies yield data that are consistent with a singlet diradical ground state with a thermally accessible triplet state for this unique bis-superoxide complex. In addition to reaction with O<sub>2</sub>, <i>cis</i>-[PdĀ­(IPr)<sub>2</sub>(Ī·<sup>2</sup>-O<sub>2</sub>)] reacts at low temperature with H<sub>2</sub>O in methanol/ether solution to form <i>trans</i>-[PdĀ­(IPr)<sub>2</sub>(OH)Ā­(OOH)]. The crystal structure of <i>trans</i>-[PdĀ­(IPr)<sub>2</sub>(OOH)Ā­(OH)] is reported. Neither reaction with O<sub>2</sub> nor reaction with H<sub>2</sub>O occurs under comparable conditions for <i>cis</i>-[PdĀ­(IMes)<sub>2</sub>(Ī·<sup>2</sup>-O<sub>2</sub>)] (IMes = 1,3-bisĀ­(2,4,6-trimethylphenyl)Ā­imidazol-2-ylidene). The increased reactivity of <i>cis</i>-[PdĀ­(IPr)<sub>2</sub>(Ī·<sup>2</sup>-O<sub>2</sub>)] is attributed to the enthalpy of binding of O<sub>2</sub> to [PdĀ­(IPr)<sub>2</sub>] (āˆ’14.5 Ā± 1.0 kcal/mol) that is approximately one-half that of [PdĀ­(IMes)<sub>2</sub>] (āˆ’27.9 Ā± 1.5 kcal/mol). Computational studies identify the cause as interligand repulsion forcing a wider Cā€“Pdā€“C angle and tilting of the NHC plane in <i>cis</i>-[PdĀ­(IPr)<sub>2</sub>(Ī·<sup>2</sup>-O<sub>2</sub>)]. Areneā€“arene interactions are more favorable and serve to further stabilize <i>cis</i>-[PdĀ­(IMes)<sub>2</sub>(Ī·<sup>2</sup>-O<sub>2</sub>)]. Inclusion of dispersion effects in DFT calculations leads to improved agreement between experimental and computational enthalpies of O<sub>2</sub> binding. A complete reaction diagram is constructed for formation of <i>trans</i>-[PdĀ­(IPr)<sub>2</sub>(Ī·<sup>1</sup>-O<sub>2</sub>)<sub>2</sub>] and leads to the conclusion that kinetic factors inhibit formation of <i>trans</i>-[PdĀ­(IMes)<sub>2</sub>(Ī·<sup>1</sup>-O<sub>2</sub>)<sub>2</sub>] at the low temperatures at which it is thermodynamically favored. Failure to detect the predicted T-shaped intermediate <i>trans</i>-[PdĀ­(NHC)<sub>2</sub>(Ī·<sup>1</sup>-O<sub>2</sub>)] for either NHC = IMes or IPr is attributed to dynamic effects. A partial potential energy diagram for initial binding of O<sub>2</sub> is constructed. A range of low-energy pathways at different angles of approach are present and blur the distinction between pure ā€œside-onā€ or ā€œend-onā€ trajectories for oxygen binding
    corecore