76 research outputs found

    Isolation of singlet carbene derived 2-phospha-1,3-butadienes and their sequential one-electron oxidation to radical cations and dications

    Get PDF
    A synthetic strategy for the 2-phospha-1,3-butadiene derivatives [{(IPr)C(Ph)}P(cAACMe^{Me})] (3a) and [{(IPr)C(Ph)}P(cAACCy^{Cy})] (3b) (IPr = C{(NDipp)CH}2_{2}, Dipp = 2,6-iPr2_{2}C6_{6}H3_{3}; cAACMe^{Me} = C{(NDipp)CMe2_{2}CH2_{2}CMe2_{2}}; cAACCy^{Cy} = C{(NDipp)CMe2_{2}CH2_{2}C(Cy)}, Cy = cyclohexyl) containing a C=C–P=C framework has been established. Compounds 3a and 3b have a remarkably small HOMO–LUMO energy gap (3a: 5.09; 3b: 5.05 eV) with a very high-lying HOMO (-4.95 eV for each). Consequently, 3a and 3b readily undergo one-electron oxidation with the mild oxidizing agent GaCl3_{3} to afford radical cations [{(IPr)C(Ph)}P(cAACR^{R})]GaCl4_{4} (R = Me 4a, Cy 4b) as crystalline solids. The main UV-vis absorption band for 4a and 4b is red-shifted with respect to that of 3a and 3b, which is associated with the SOMO related transitions. The EPR spectra of compounds 4a and 4b each exhibit a doublet due to coupling of the unpaired electron with the 31^{31}P nucleus. Further oneelectron removal from the radical cations 4a and 4b is also feasible with GaCl3_{3}, affording the dications [{(IPr)C(Ph)}P(cAACR^{R})](GaCl4_{4})2_{2} (R = Me 5a, Cy 5b) as yellow crystals. The molecular structures of compounds 3–5 have been determined by X-ray diffraction and analyzed by DFT calculations

    A Combined Spectroscopic and Computational Study on the Mechanism of Iron-Catalyzed Aminofunctionalization of Olefins Using Hydroxylamine Derived N-O Reagent as the "Amino" Source and "Oxidant".

    Get PDF
    Herein, we study the mechanism of iron-catalyzed direct synthesis of unprotected aminoethers from olefins by a hydroxyl amine derived reagent using a wide range of analytical and spectroscopic techniques (Mössbauer, Electron Paramagnetic Resonance, Ultra-Violet Visible Spectroscopy, X-ray Absorption, Nuclear Resonance Vibrational Spectroscopy, and resonance Raman) along with high-level quantum chemical calculations. The hydroxyl amine derived triflic acid salt acts as the "oxidant" as well as "amino" group donor. It activates the high-spin Fe(II) (St = 2) catalyst [Fe(acac)2(H2O)2] (1) to generate a high-spin (St = 5/2) intermediate (Int I), which decays to a second intermediate (Int II) with St = 2. The analysis of spectroscopic and computational data leads to the formulation of Int I as [Fe(III)(acac)2-N-acyloxy] (an alkyl-peroxo-Fe(III) analogue). Furthermore, Int II is formed by N-O bond homolysis. However, it does not generate a high-valent Fe(IV)(NH) species (a Fe(IV)(O) analogue), but instead a high-spin Fe(III) center which is strongly antiferromagnetically coupled (J = -524 cm-1) to an iminyl radical, [Fe(III)(acac)2-NH·], giving St = 2. Though Fe(NH) complexes as isoelectronic surrogates to Fe(O) functionalities are known, detection of a high-spin Fe(III)-N-acyloxy intermediate (Int I), which undergoes N-O bond cleavage to generate the active iron-nitrogen intermediate (Int II), is unprecedented. Relative to Fe(IV)(O) centers, Int II features a weak elongated Fe-N bond which, together with the unpaired electron density along the Fe-N bond vector, helps to rationalize its propensity for N-transfer reactions onto styrenyl olefins, resulting in the overall formation of aminoethers. This study thus demonstrates the potential of utilizing the iron-coordinated nitrogen-centered radicals as powerful reactive intermediates in catalysis

    Reversed Freeze Quench Method near the Solvent Phase Transition

    No full text
    Freeze quenching is a general method for trapping reaction intermediates on a (sub)­millisecond time scale. The method relies on a mixing and subsequent rapid freezing of solutions of reactants. If the reaction is limited by diffusion, it may be advantageous to initially mix the reactants under conditions where the reaction does not proceed, e.g., by mixing them at low temperature as solids. The temperature may then be raised close to the melting point of the solvent. Depending on the viscosity of the solvent, the temperature can be raised either by heating or by applying laser pulses of nanosecond length with concomitant conversion of light into heat. A reduction of the dead time and a good control of the reaction speed in comparison to the standard freeze quench technique has been achieved with this method. The feasibility of the method in combination with EPR spectroscopy is verified by examining the important prototypical reductions of benzoquinone and 2,6-dichlorophenolindophenol by ascorbate as representatives for two-step redox reactions. By using light pulses of a laser, the reaction could be driven with rates lowered by 4 orders of magnitude as compared to room temperature reaction rates. This has allowed the observation of previously unobserved radical intermediates: the reduction of DCPIP by ascorbate is found to be strongly pH dependent. It proceeds via two one-electron steps at low pH, whereas at neutral pH, the reduction of DCPIP by ascorbate proceeds in a 1:2 stoichiometry followed by a disproportionation of the ascorbate radicals

    Electronic Structure of the Lowest Triplet State of Flavin Mononucleotide

    No full text
    The electronic structure of flavin mononucleotide (FMN), an organic cofactor that plays a role in many important enzymatic reactions, has been investigated by electron paramagnetic resonance (EPR) spectroscopy, optical spectroscopy, and quantum chemistry. In particular, the triplet state of FMN, which is paramagnetic (total spin <i>S</i> = 1), allows an investigation of the zero field splitting parameters <i>D</i> and <i>E</i>, which are directly related to the two singly occupied molecular orbitals. Triplet EPR spectra and optical absorption spectra at different pH values in combination with time dependent density functional theory (TDDFT) reveal that the highest occupied orbital (HOMO) and lowest unoccupied orbital (LUMO) of FMN are largely unaffected by changes in the protonation state of FMN. Rather, the orbital structure of the lower lying doubly occupied orbitals changes dramatically. Additional EPR experiments have been carried out in the presence of AgNO<sub>3</sub>, which allows the formation of an Ag–FMN triplet state with different zero field splitting parameters and population and depopulation rates. Addition of AgNO<sub>3</sub> only induces small changes in the optical spectrum, indicating that the Ag<sup>+</sup> ion only contributes to the zero field splitting by second order spin–orbit coupling and leaves the orbital structure unaffected. By a combination of the three employed methods, the observed bands in the UV/vis spectra of FMN at different pH values are assigned to electronic transitions

    Raman Spectroscopy as a Method to Investigate Catalytic Intermediates: CO<sub>2</sub> Reducing [Re(Cl)(bpy-R)(CO)<sub>3</sub>] Catalyst

    No full text
    Complexes of the type [Re­(Cl)­(bpy-R)­(CO)<sub>3</sub>] (<b>1</b>, bpy = bipyridine, R = <sup><i>t</i></sup>Bu, H, CF<sub>3</sub>) show high catalytic activity for electrochemical CO<sub>2</sub> reduction. Application of Raman spectroscopy to these complexes as well as to the doubly reduced species [Re­(bpy-R)­(CO)<sub>3</sub>]<sup>−</sup> (<b>3</b>), which are the postulated active species, and the monoreduced complex [Re­(Cl)­(bpy-CF<sub>3</sub>)­(CO)<sub>3</sub>]<sup>−</sup> (<b>2</b>) and comparison with state-of-the-art quantum chemical calculations allows accurate investigation of electronic structures as well as geometries. For doubly reduced complexes, calculations point out a formal closed-shell singlet state only compatible with a formal {Re<sup>I</sup>(bpy-R)<sup>2–</sup>} moiety. In contrast, based on molecular orbital analysis and the change of the <i>actual</i> charge distribution during the overall two-electron reduction, the system is better described as {Re<sup>0</sup>(bpy-R<sup>•</sup>)<sup>−</sup>}. Interestingly, the Raman spectra of the monoreduced and doubly reduced complexes with the CF<sub>3</sub>-substituted bpy ligand are virtually identical, which points to the same overall electronic structure of the bpy species in both complexes. Additional Raman experiments and calculations of [Re­(COOH)­(bpy)­(CO)<sub>3</sub>] (<b>4</b>) and [Re­(bpy)­(CO)<sub>4</sub>]­OTf (<b>5</b>), which are proposed to be intermediates of the catalytic cycle for CO<sub>2</sub> reduction, confirm the presence of neutral bpy showing that the reducing equivalent stored at the bidentate ligand is involved in the activation of CO<sub>2</sub>. As such, Raman spectroscopy combined with quantum chemical calculations is an ideal tool to investigate catalysts with redox active ligands, since the spectra give local information about the electronic and geometric structure of the molecule

    Electronic Structure of the Cysteine Thiyl Radical: A DFT and Correlated ab Initio Study

    No full text
    The electronic structure and the unusual EPR parameters of sulfur-centered alkyl thiyl radical from cysteine are investigated by density functional theory (DFT) and correlated ab initio calculations. Three geometry-optimized, staggered conformations of the radical are found that lie within 630 cm-1 in energy. The EPR g-values are sensitive to the energy difference between the nearly-degenerate singly occupied orbital and one of the lone-pair orbitals (excitation energies of 1732, 1083, and 3429 cm-1 from Multireference Configuration Interaction calculations for the structures corresponding to the three minima), both of which are almost pure sulfur 3p orbitals. Because of the near degeneracy, the second order correction to the g tensor, which is widely used to analyze g-values of paramagnetic systems, is insufficient to obtain accurate g-values of the cysteine thiyl radical. Instead, an expression for the g tensor must be used in which third order corrections are taken into account. The near-degeneracy can be affected to roughly equal extents by changes in the structure of the radical and by hydrogen bonds to the sulfur. The magnitude of the hyperfine coupling constants for the β protons of the cysteine thiyl radical is found to depend on the structure of the radical. On the basis of a detailed comparison between experimental and calculated g-values and hyperfine coupling constants an attempt is made to identify the structure of thiyl radicals and the number of hydrogen bonds to the sulfur

    Sterically Stabilized End-On Superoxocopper(II) Complexes and Mechanistic Insights Into Their Reactivity With O-H, N-H and C-H Substrates.

    No full text
    Instability of end-on superoxocopper(II) complexes, with respect to conversion to the corresponding peroxo-bridged complexes, has largely constrained their study to very low temperatures (< -80°C). This limits their kinetic capacity to oxidize substrates. In response, we have developed a series of ligand systems bearing bulky aryl substituents that are primarily directed away from the metal centre, Ar3-TMPA (Ar = tpb, dpb, dtbpb), and used them to support [Cu(Ar3-TMPA)(NCMe)]+ copper(I) complexes. Solutions of all three react with O2 to yield [Cu(η1-O2•−)(Ar3-TMPA)]+ complexes that are stable against dimerization at all temperatures. Full binding of O2 is observed at sub-ambient temperatures and can be reversed by warming. The onset of oxygenation is ligand dependent, but can be observed at 25°C in the case of Ar = tpb and dpb. Furthermore, all three [Cu(η1-O2•−)(Ar3-TMPA)]+ complexes are stable against self-decay at temperatures ≤ -20°C. This provides a wide temperature window over which these complexes can be studied, which was exploited by performing extensive reaction kinetics measurements for [Cu(η1-O2•−)(tpb3-TMPA)]+ with a broad range of O-H, N-H, and C-H bond substrates. This includes correlation of second order rate constants (k2 values) versus oxidation potentials (Eox) for a range of phenols (i.e., a Marcus plot), construction of Eyring plots, and temperature-dependent kinetic isotope effect (KIE) measurements. The data obtained indicates that reaction with all substrates proceeds via H-atom transfer (HAT) to [Cu(η1-O2•−)(tpb3-TMPA)]+. In addition, evidence suggests that HAT reaction with the phenols studied proceeds with significant charge transfer, and that it involves full tunelling of both H and D atoms in the case of 1,2-diphenylhydrazine (DPH) and 4-methoxy-2,6-di-tert-butylphenol (MeO-ArOH). Consistent with expectations for HAT, large entropic barriers (ΔS‡) were measured for the substrates MeO-ArOH, DPH, triphenylhydrazine (TPH), and 1-benzyl-1,4-dihydronicotinamide (BNAH). Despite having the lowest X-H bond dissociation energy (BDE) amongst these substrates, the C-H substrate BNAH exhibits both the largest ΔS‡ and the second largest enthalpic barrier (ΔH‡) to reaction. This is congruent with the expectation that oxidation of C-H bonds is kinetically challenging and the experimental observation that [Cu(η1-O2•−)(tpb3-TMPA)]+ is only able to oxidize very weak C-H bonds, whereas it can oxidize moderately strong N-H bonds

    Theoretical Spectroscopy of the Ni-II Intermediate States in the Catalytic Cycle and the Activation of [NiFe] Hydrogenases

    Get PDF
    [NiFe] hydrogenases catalyze the reversible oxidation of dihydrogen. The corresponding catalytic cycle involves a formidable number of redox states of the Ni-Fe active site; these can be distinguished experimentally by the IR stretching frequencies of their CN and CO ligands coordinated to iron. These spectroscopic fingerprints serve as sensitive probes for the intrinsic electronic structure of the metal core and, indirectly, for the structural composition of the active site. In this study, density functional theory (DFT) was used to calculate vibrational frequencies, by focusing on the EPR-silent intermediate states that contain divalent metal centers. By using the well-characterized Ni-C and Ni-B states as references, we identified candidates for the Ni-SIr, Ni-SIa, and Ni-R states by matching the predicted relative frequency shifts with experimental results. The Ni-SIr and Ni-SIa states feature a water molecule loosely bound to nickel and a formally vacant bridge. Both states are connected to each other through protonation equilibria; that is, in the Ni-SIa state one of the terminal thiolates is protonated, whereas in Ni-SIr this thiolate is unprotonated. For the reduced Ni-R state two feasible models emerged: in one, H2 coordinates side-on to nickel, and the second features a hydride bridge and a protonated thiolate. The Ni-SU state remains elusive as no unequivocal correspondence between the experimental data and calculated frequencies of the models was found, thus indicating that a larger structural rearrangement might occur upon reduction from Ni-A to Ni-SU and that the bridging ligand might dissociate
    • …
    corecore