• Optimizing Millisecond Time Scale Near-Infrared Emission in Polynuclear Chrome(III)-Lanthanide(III) Complexes
    L. Aboshyan-Sorgho, H. Nozary, A. Aebischer, J.-C.G. Bünzli, , K.R. Kittilstved, A. Hauser, S.V. Eliseeva, S. Petoud and C. Piguet
    Journal of the American Chemical Society, 134 (30) (2012), p12675-12684
    DOI:10.1021/ja304009b | unige:22645 | Abstract | Article HTML | Article PDF
 
This work illustrates a simple approach for optimizing long-lived near-infrared lanthanide-centered luminescence using trivalent chromium chromophores as sensitizers. Reactions of the segmental ligand L2 with stoichiometric amounts of M(CF3SO3)2 (M = Cr, Zn) and Ln(CF3SO3)3 (Ln = Nd, Er, Yb) under aerobic conditions quantitatively yield the D3-symmetrical trinuclear [MLnM(L2)3](CF3SO3)n complexes (M = Zn, n = 7; M = Cr, n = 9), in which the central lanthanide activator is sandwiched between the two transition metal cations. Visible or NIR irradiation of the peripheral Cr(III) chromophores in [CrLnCr(L2)3]9+ induces rate-limiting intramolecular intermetallic Cr→Ln energy transfer processes (Ln = Nd, Er, Yb), which eventually produces lanthanide-centered near-infrared (NIR) or IR emission with apparent lifetimes within the millisecond range. As compared to the parent dinuclear complexes [CrLn(L1)3]6+, the connection of a second strong-field [CrN6] sensitizer in [CrLnCr(L2)3]9+ significantly enhances the emission intensity without perturbing the kinetic regime. This work opens novel exciting photophysical perspectives via the buildup of non-negligible population densities for the long-lived doubly excited state [Cr*LnCr*(L2)3]9+ under reasonable pumping powers.
  • Thermodynamics, Structure and Properties of Polynuclear Lanthanide Complexes with a Tripodal Ligand: Insight into their Self-Assembly
    J. Hamacek, C. Besnard, T. Penhouet and
    Chemistry - A European Journal, 17 (24) (2011), p6753-6764
    DOI:10.1002/chem.201100173 | unige:17236 | Abstract | Article HTML | Article PDF
Self-assembly processes between a tripodal ligand and LnIII cations have been investigated by means of supramolecular analytical methods. At an equimolar ratio of components, tetranuclear tetrahedral complexes are readily formed in acetonitrile. The structural analysis of the crystallographic data shows a helical wrapping of binding strands around metallic cations. The properties of this series of highly charged 3D compounds were examined by using NMR spectroscopy and optical methods in solution and in the solid state. In the presence of excess metal, a new trinuclear complex was identified. The X-ray crystal structure elucidated the coordination of metallic cations with two ligands of different conformations. By varying the metal/ligand ratio, a global speciation of this supramolecular system has been evidenced with different spectroscopic methods. In addition, these rather complicated equilibria were successfully characterised with the thermodynamic stability constants. A rational analysis of the self-assembly processes was attempted by using the thermodynamic free energy model and the impact of the ligand structure on the effective concentration is discussed.
  • What governs nitrogen configuration in substituted aminophosphines?
    M.D. Wodrich, A. Vargas, , G. Merino and C. Corminboeuf
    Journal of Physical Organic Chemistry, 22 (2) (2008), p101-109
    DOI:10.1002/poc.1431 | unige:3181 | Abstract | Article PDF
The trigonal planar geometry of the nitrogen atom in commonly used phosphoramidite ligands is not in line with the traditional valence shell electron pair repulsion (VSEPR) model. In this work, the effects governing nitrogen configuration in several substituted aminophosphines, A2PNB2 (A or B = H, F, Cl, Br, Me, OMe, BINOP), are examined using modern computational analytic tools. The electron delocalization descriptions provided by both electron localization function (ELF) and block localized wavefunction analysis support the proposed relationships between conformation and negative hyperconjugative interactions. In the parent H2PNH2, the pyramidal nitrogen configuration results from nitrogen lone pair electron donation into the σ* P — H orbital. While enhanced effects are seen for F2PNMe2, placing highly electronegative fluorine substituents on nitrogen (i.e., Me2PNF2) eliminates delocalization of the nitrogen lone pair. Understanding and quantifying these effects can lead to greater flexibility in designing new catalysts.
  
  • Tuning the Polarization Along Linear Polyaromatic Strands for Rationally Inducing Mesomorphism in Lanthanide Nitrate Complexes
    E. Terazzi, L. Guénée, , G. Bernardinelli, B. Donnio, D. Guillon and C. Piguet
    Chemistry - A European Journal, 13 (6) (2007), p1674-1691
    DOI:10.1002/chem.200601389 | unige:3599 | Abstract | Article HTML | Article PDF
The opposite orientation of the ester spacers in the rodlike ligands L 4C12 (benzimidazole-OOC-phenyl) and L 5C12 (benzimidazole-COO-phenyl) drastically changes the electronic structure of the aromatic systems, without affecting their meridional tricoordination to trivalent lanthanides, LnIII, and their thermotropic liquid crystalline (i.e., mesomorphic) behaviors. However, the rich mesomorphism exhibited by the complexes [Ln(L 4C12)(NO3)3] (Ln=La-Lu) vanishes in [Ln(L 5C12)(NO3)3], despite superimposable molecular structures and comparable photophysical properties. Density functional theory (DFT) and time-dependant DFT calculations performed in the gas phase show that the inversion of the ester spacers has considerable effects on the electronic structure and polarization of the aromatic groups along the strands, which control residual intermolecular interactions responsible for the formation of thermotropic liquid-crystalline phases. As a rule of thumb, an alternation of electron-poor and electron-rich aromatic rings favors intermolecular interactions between the rigid cores and consequently mesomorphism, a situation encountered for L 4C12, L 5C12, [Ln(L 4C12)(NO3)3], but not for [Ln(L 5C12)(NO3)3]. The intercalation of an additional electron-rich diphenol ring on going from [Ln(L 5C12)(NO3)3] to [Ln(L 6C12)(NO3)3] restores mesomorphism despite an unfavorable orientation of the ester spacers, in agreement with our simple predictive model.
  • Tuning facial-meridional isomerisation in monometallic nine-coordinate lanthanide complexes with unsymmetrical tridentate ligands
    T. Le Borgne, P. Altmann, N. André, J.-C.G. Bünzli, G. Bernardinelli, , J. Weber and C. Piguet
    Dalton Transactions, (5) (2004), p723-733
    DOI:10.1039/b316035a | unige:3251 | Abstract | Article HTML | Article PDF
The unsymmetrical tridentate benzimidazole–pyridine–carboxamide units in ligands L1–L4 react with trivalent lanthanides, LnIII, to give the nine-co-ordinate triple-helical complexes [Ln(Li)3]3+ (i=1–4) existing as mixtures of C3-symmetrical facial and C1-symmetrical meridional isomers. Although the β13 formation constants are 3–4 orders of magnitude smaller for these complexes than those found for the D3-symmetrical analogues [Ln(Li)3]3+ (i=5–6) with symmetrical ligands, their formation at the millimolar scale is quantitative and the emission quantum yield of [Eu(L2)3]3+ is significantly larger. The fac-[Ln(Li)3]3+↔mer-[Ln(Li)3]3+ (i =1–4) isomerisation process in acetonitrile is slow enough for Ln=LuIII to be quantified by 1H NMR below room temperature. The separation of enthalpic and entropic contributions shows that the distribution of the facial and meridional isomers can be tuned by the judicious peripheral substitution of the ligands affecting the interstrand interactions. Molecular mechanics (MM) calculations suggest that one supplementary interstrand -stacking interaction stabilises the meridional isomers, while the facial isomers benefit from more favourable electrostatic contributions. As a result of the mixture of facial and meridional isomers in solution, we were unable to obtain single crystals of 13 complexes, but the X-ray crystal structures of their nine-co-ordinate precursors [Eu(L1)2(CF3SO3)2(H2O)](CF3SO3)(C3H5N)2(H2O) ( 6, C45H54EuF9N10O13S3, monoclinic, P21/c, Z=4) and [Eu(L4)2(CF3SO3)2(H2O)](CF3SO3)(C4H4O)1.5 ( 7, C51H66EuF9N8O15.5S3, triclinic, P, Z=2) provide crucial structural information on the binding mode of the unsymmetrical tridentate ligands.
  • Unusual regio-and enantioselective [1,2]-Stevens rearrangement of a spirobi[dibenzazepinium] cation
    L. Vial, M.-H. Gonçalves, , J. Weber, G. Bernardinelli and J. Lacour
    Synlett, (9) (2004), p1565-1568
    DOI:10.1055/s-2004-829069 | unige:3256 | Abstract | Article HTML | Article PDF
Highly symmetric spirobi[dibenzazepinium] cation 3 reacts with P4-t-Bu to form exclusively a ring-expanded tertiary amine; this unusual reactivity can be traced back to the geometry of the ylide.
  • The solution structure of homotrimetallic lanthanide helicates investigated with novel model-free multi-centre paramagnetic NMR methods
    N. Ouali, J.-P. Rivera, , J. Weber and C. Piguet
    Dalton Transactions, (7) (2003), p1251-1263
    DOI:10.1039/b212352e | unige:3505 | Abstract | Article HTML | Article PDF
 
The combination of one contact and three pseudo-contact contributions to the NMR hyperfine paramagnetic shift of each proton in the triple-stranded helicates [Ln3(L1)3]9+(Ln = Ce–Yb except Pm, Gd) produce intractable 1H NMR spectra whose assignment is limited by the large electronic contribution to the nuclear relaxation processes. The detailed analysis of the NMR spectra for the diamagnetic complexes [Ln3(L1)3]9+(Ln = La, Lu, Y) shows that the triple-helical structure found in the solid state is maintained in solution. Extension of the classical one-nucleus crystal-field dependent model-free method for paramagnetic D3-symmetrical homotrimetallic lanthanide complexes possessing two different metallic sites (i.e. two second-rank crystal-field parameters: B20central and B20terminal) allows (i) the complete interpretation of the paramagnetic signals for Ln = Ce–Yb and (ii) the detection of a concomitant abrupt change of the contact terms Fi and of the pseudo-contact terms Si=B20centralG1i+B20terminal(G2i+G3i) occurring near the middle of the lanthanide series. The derivation and application of a novel three-nuclei crystal-field independent method eventually demonstrates that the helicates [Ln3(L1)3]9+ adopt a single D3-symmetrical structure along the complete lanthanide series in solution, which ascribes the discontinuity observed for Si to a concomitant decrease of the two crystal-field parameters. Comparison with structural models is limited by the extreme sensitivity of the structural factors Cikl and Dikl to minor geometrical variations affecting the wrapping of the ligand strands, but calculations of the geometrical factors Gmi(m= 1–3) for [Ln3(L1)3]9+ in solution confirm the formation of a regular triple-helical structure. Generalization of this novel three-nuclei method for addressing the solution structure of rhombic lanthanide complexes is discussed.
  • Connecting Terminal Carboxylate Groups in Nine-Coordinate Lanthanide Podates: Consequences on the Thermodynamic, Structural, Electronic, and Photophysical Properties
    J.-M. Senegas, G. Bernardinelli, D. Imbert, J.-C.G. Bünzli, , J. Weber and C. Piguet
    Inorganic Chemistry, 42 (15) (2003), p4680-4695
    DOI:10.1021/ic034231t | unige:3236 | Abstract | Article HTML | Article PDF
The hydrolysis of terminal tbutyl-ester groups provides the novel nonadentate podand tris{2-[N-methylcarbamoyl-(6-carboxypyridine-2)-ethyl]amine} (L13) which exists as a mixture of slowly interconverting conformers in solution. At pH = 8.0 in water, its deprotonated form [L13 − 3H]3- reacts with Ln(ClO4)3 to give the poorly soluble and stable podates [Ln(L13 − 3H)] (log(β110) = 6.7−7.0, Ln = La−Lu). The isolated complexes [Ln(L13 − 3H)](H2O)7 (Ln = Eu, 8; Tb, 9; Lu, 10) are isostructural, and their crystal structures show Ln(III) to be nine-coordinate in a pseudotricapped trigonal prismatic site defined by the donor atoms of the three helically wrapped tridentate binding units of L13. The Ln−O(carboxamide) bonds are only marginally longer than the Ln−O(carboxylate) bonds in [Ln(L13 − 3H)], thus producing a regular triple helix around Ln(III) which reverses its screw direction within the covalent Me−TREN tripod. High-resolution emission spectroscopy demonstrates that (i) the replacement of terminal carboxamides with carboxylates induces only minor electronic changes for the metallic site, (ii) the solid-state structure is maintained in water, and (iii) the metal in the podate is efficiently protected from interactions with solvent molecules. The absolute quantum yields obtained for [Eu(L13 − 3H)] ( ΦtotEu = 1.8 × 10-3) and [Tb(L13 − 3H)] ( ΦtotEu = 8.9 × 10-3) in water remain modest and strongly contrast with that obtained for the lanthanide luminescence step (ΦEu = 0.28). Detailed photophysical studies assign this discrepancy to the small energy gap between the ligand-centered singlet (1ππ*) and triplet (3ππ*) states which limits the efficiency of the intersystem crossing process. Theoretical TDDFT calculations suggest that the connection of a carboxylate group to the central pyridine ring prevents the sizable stabilization of the triplet state required for an efficient sensitization process. The thermodynamic and electronic origins of the advantages (stability, lanthanide quantum yield) and drawbacks (solubility, sensitization) brought by the “carboxylate effect” in lanthanide complexes are evaluated for programming predetermined properties in functional devices.
  • Density Functional Study of a Helical Organic Cation
    D. Bas, , J. Weber and T.A. Wesolowski
    Chimia, 57 (4) (2003), p173-174
    DOI:10.2533/000942903777679460 | unige:3240 | Abstract | Article PDF
We report on the first stage of our theoretical study of the quino[2,3,4-kl]acridinium,1,13-dimethoxy-5,9-dipropyl-cation. This molecule, involved in the synthesis of novel triazaangulenium dyes of high chemical stability, is a chiral [4]-helicenium. The structure and the IR spectrum of the quino[2,3,4-kl]acridinium,1,13-dimethoxy-5,9-dimethyl-cation derived from theoretical calculations which use various density functional theory methods, are compared with the geometry derived from X-ray diffraction measurements and the experimental IR spectrum. Our study shows that the chosen variant of DFT methods (Becke88 for exchange, P86 for correlation, 3-21G** basis set) reproduces the experimental geometry within 0.004 Å and the IR frequencies within 15 cm−1.
  • Analysis of paramagnetic NMR spectra of triple-helical lanthanide complexes with 2,6-dipicolinic acid revisited : a new assignment of structural changes and crystal-field effects 25 years later
    N. Ouali, B. Bocquet, S. Rigault, , J. Weber and C. Piguet
    Inorganic Chemistry, 41 (6) (2002), p1436-1445
    DOI:10.1021/ic010801i | unige:3223 | Abstract | Article HTML | Article PDF
Variable-temperature 1H and 13C NMR measurements of the D3-symmetrical triple-helical complexes [Ln(L1-2H)3]3- (L1 = pyridine-2,6-dicarboxylic acid; Ln = La−Lu) show evidence of dynamic intermolecular ligand-exchange processes whose activation energies depend on the size of the metal ion. At 298 K, the use of diastereotopic probes in [Ln(L3-2H)3]3- (L3 = 4-ethyl-pyridine-2,6-dicarboxylic acid) shows that fast intramolecular P ↔ M interconversion between the helical enantiomers occurs on the NMR time scale. Detailed analyses of the paramagnetic NMR hyperfine shifts according to crystal-field independent techniques demonstrate the existence of two different helical structures, one for large lanthanides (Ln = La−Eu) and one for small lanthanides (Ln = Tb−Lu), in complete contrast with the isostructurality proposed 25 years ago. A careful reconsideration of the original crystal-field-dependent analysis shows that an abrupt variation of the axial crystal-field parameter A202> parallels the structural change leading to some accidental compensation effects that prevent the detection of structural variations according to the classical one-nucleus method. Crystal structures in the solid state and density functional theory calculations in the gas phase provide structural models that rationalize the paramagnetic NMR data. A regular triple-helical structure is found for small lanthanides (Ln = Tb−Lu) in which the terdentate chelating ligands are rigidly tricoordinated to the metals. A flexible and distorted structure is evidenced for Ln = La−Eu in which the central pyridine rings interact poorly with the metal ion. The origin of the simultaneous variation of structural parameters and crystal-field and hyperfine constants near the middle of the lanthanide series is discussed together with the use of crystal-field-independent techniques for the interpretation of paramagnetic NMR spectra in axial lanthanide complexes.
  • Aromatic bent-core liquid crystals : an opportunity for introducing terdentate binding units into mesophases
    H. Nozary, J.-P. Rivera, P. Tissot, C. Piguet, , J. Weber, G. Bernardinelli, J.-C.G. Bünzli, R. Deschenaux, B. Donnio and D. Guillon
    Chemistry of Materials, 14 (3) (2002), p1075-1090
    DOI:10.1021/cm011162c | unige:3222 | Abstract | Article HTML | Article PDF
Lipophilic linear semirigid side arms containing two or three successive phenyl rings separated by carboxylate spacers have been connected to the 5 or 6 positions of bent aromatic terdentate 2,6-bis(benzimidazol-2-yl)pyridine binding units to give extended V-shaped (L11) and I-shaped receptors (L12, L12b, and L13). The carboxylate spacers limit the flexibility of the side arms and provide crossed arrangements of the successive aromatic rings in the crystal structure of L12b (C63H61N5O10; triclinic, P↑, Z = 2) in agreement with semiempirical calculations performed on optimized gas-phase geometries. Moreover, the carboxylate spacers in L11−L13 prevent efficient electronic delocalization between the connected aromatic rings and act as weak π acceptors producing a slight increase of the energy of the 1ππ* and3ππ* levels centered on the terdentate binding unit. Intermolecular π-stacking interactions observed in the crystal of L12b are invoked to rationalize (i) the peculiar excimer emission ofL11 in the solid state and (ii) the rich and varied calamitic (I-shaped L12, L12b, and L13) and columnar (V-shaped L11) mesomorphism observed at high temperature. The ColR mesophase detected for L11 demonstrates that V-shaped bent terdentate binding units are compatible with liquid-crystalline behavior. Complexation of L11 with lanthanide(III) produces I-shaped complexes [Ln(L11)(NO3)3] (Ln = La, Eu, Gd, Tb, and Lu) possessing a large axial anisometry as found in the crystal structure of [Lu(L11)(CF3CO2)3(H2O)] (LuC81H87N5O17F9; triclinic, P↑,Z = 2), which exists in the solid state as H-bonded dimers. No mesomorphism is detected for the complexes as a result of the large perpendicular expansion brought by the metallic coordination site, but the high energy of the ligand-centered 3ππ* prevents Eu(5D0) → L11back transfer in the Eu(III) complex, which thus exhibits sizable red luminescence at room temperature, a crucial point for the design of luminescent materials.
  • Intermolecular interaction energies from the total energy bi-functional. A case study of carbazole complexes.
    T.A. Wesolowski, and J. Weber
    Journal of Chemical Physics, 116 (15) (2002), p6411-6421
    DOI:10.1063/1.1462613 | unige:3235 | Abstract | Article HTML | Article PDF
An approach in which the total energy of interacting subsystems is expressed as a bifunctional depending explicitly on two functions: electron densities of the two molecules forming a complex (ρ1 and ρ2) was used to determine the equilibrium geometry and the binding energy of several weak intermolecular complexes involving carbazole and such atoms or molecules as Ne, Ar, CH4, CO, and N2. For these complexes, the experimental dissociation energies fall within the range from 0.48 to 2.06 kcal/mol. Since the effect of the intermolecular vibrations on the dissociation energy is rather small, the experimental measurements provide an excellent reference set. The obtained interaction energies are in a good agreement with experiment and are superior to the ones derived from conventional Kohn–Sham calculations. A detailed analysis of relative contribution of the terms which are expressed using approximate functionals (i.e., exchange-correlationExc[ρ1+ρ2] and nonadditive kinetic energy Tsnad[ρ1,ρ2] = Ts[ρ1+ρ2]−Ts[ρ1]−Ts[ρ2]) is made. The nonvariational version of the applied formalism is also discussed.
  • Coupled-cluster calculations on ferrocene and its protonated derivatives : the final word on the mechanism of protonation of ferrocene ?
    M.J. Mayor-Lopez, H.P. Lüthi, H. Koch, and J. Weber
    Journal of Chemical Physics, 113 (18) (2000), p8009-8014
    DOI:10.1063/1.1316031 | unige:3344 | Abstract | Article HTML | Article PDF | Article PS (gzipped)
The mechanism of the protonation of ferrocene, the simplest model for the electrophilic attack on a metallocene, has been studied extensively. However, neither experiment nor computation have reached agreement on the details of the mechanism. The different model calculations applied [Hartree–Fock, Möller–Plesset, and density functional theory (HF, MP2, and DFT) with different functionals] come to contradicting conclusions. As a complement to our previous work, we report the results obtained for neutral and protonated ferrocene using the coupled-cluster method [CCSD(T)] with polarized double- and triple-zeta basis sets. These calculations show that the metal-protonated and the agostic forms represent minima on the potential energy surface, whereas the ring-protonated form is higher in energy with no minimum structure identified. With regard to the reaction, these results indicate an exo reaction path. The CCSD(T) results are in good agreement with the predictions made by the DFT calculations, whereas the results obtained from the Hartree–Fock and MP2 computations appear to be incorrect.
  • Binding of Organic Cations to a Cyclophane Host As Studied with Molecular Dynamics Simulations and Free Energy Calculations
    M.A.L. Eriksson, and P.A. Kollman
    Journal of Physical Chemistry B, 103 (21) (1999), p4474-4480
    DOI:10.1021/jp983513p | unige:2937 | Abstract | Article HTML | Article PDF
We have studied the binding of two organic cations, an iminium (IM) and a guanidinium (GU), to a cyclophane host P4--4Na+, using molecular dynamics simulations and free energy calculations. A proper treatment of the long-range electrostatic forces is essential for the stability of these highly charged complexes, and a simple cutoff at 12 Å results in an artifactual dissociation of the IM−P4--4Na+ complex. Since the host is highly aromatic and the guests cationic, cation−π interactions play an important role in the complex stability. In free energy calculations, using a simple additive force field, we calculate that the relative free energy of association of IM and GU binding to the host is 2.3 kcal/mol favoring IM, which is of the correct sign but 1.4 kcal/mol too small in magnitude. Differences in van der Waals interaction energies are mainly responsible for the different binding strengths, and the host adopts different shapes when accommodating IM compared to GU. To approximately estimate the contribution to the complex stability from the polarization energy, we calculated the in vacuo interaction energies in the two complexes, using a nonadditive force field, previously shown to accurately describe alkali cation−aromatic interaction energies in vacuo. Adding the contribution from the polarization energy upon forming the two complexes in this calculation to the estimate from the free energy calculation, we obtain an improved relative binding free energy (−4.0 kcal/mol), which is in close agreement with the experimental value of −3.7 kcal/mol.
  • Theoretical investigation of the adsorption of methanol on the (110) surface of γ-alumina
    D.A. De Vito, F. Gilardoni, L. Kiwi-Minsker, , S. Porchet, A. Renken and J. Weber
    Journal of Molecular Structure: THEOCHEM, 469 (1-3) (1999), p7-14
    DOI:10.1016/S0166-1280(98)00511-9 | unige:2742 | Abstract | Article HTML | Article PDF
The adsorption of methanol on the (110) surface of Îł-alumina was investigated using both ab initio and density functional theory quantum chemical methods. A [Al3O9H10]+ cluster model comprising one tetrahedral and two octahedral aluminum cations were used to describe the surface and the mechanism of adsorption of methanol. This has allowed us to rationalize the stable structures of adsorbate and the mode of bonding. The IR frequency shifts between the gas phase and the adsorbed species were also calculated and they exhibit good agreement with experiment.
  • Teaching Computational Chemistry
    J. Weber, and H.P. Lüthi
    in "Encyclopedia of Computational Chemistry" (P.v.R. Schleyer, N.L. Allinger, T. Clark, J. Gasteiger, P.A. Kollman, H.F. Schaefer and P.R. Schreiner, eds.), Wiley, Chichester, (1998), p2960
  • How much correlation can we expect to account for in density functional calculations ? Case studies of electrostatic properties of small molecules
    J. Weber, P. Jaber, P. Gulbinat and
    in "Strategies and Applications in Quantum Chemistry" (Y. Ellinger and M. Defranceschi, eds.), Kluwer, Dordrecht, (1996), p219
  • Solvation Free Energies of Amides and Amines: Disagreement between Free Energy Calculations and Experiment
    and P.A. Kollman
    Journal of the American Chemical Society, 117 (22) (1995), p6057-6063
    DOI:10.1021/ja00127a019 | unige:2988 | Abstract | Article PDF
We present molecular dynamidfree energy calculations on the molecules acetamide, N-methylacetamide, N,N-dimethylacetamide, ammonia, methylamine, dimethylamine, and trimethylamine. Unlike the experimental data, which suggest a very non-additive solvation free energy (N-methylacetamide and methylamine having the most negative free energy of solvation), the calculations all find that the free energy of solvation monotonically increases as a function of methyl addition. The disagreement with experiment is surprising, given the very good agreement (within 0.5 kcal/mol) with experiment for calculation of the solvation free energy of methane, ethane, propane, water, methanol, and dimethyl ether.
  • Teaching computational chemistry using computers
    J. Weber and
    Chimia, 49 (3) (1995), p77-83
    unige:2842 | Abstract | Article PDF
  • Comparative modelling studies on 3,6-substituted 1,2,4-trioxan-5-ones
    C.W. Jefford, G. Bernardinelli, M.C. Josso, and J. Weber
    Journal of Molecular Structure: THEOCHEM, 337 (1) (1995), p31-37
    DOI:10.1016/0166-1280(94)04110-E | unige:2861 | Abstract | Article PDF
The main geometric elements of three 3,6-substituted 1,2,4-trioxan-5-ones have been calculated by using molecular mechanics (MM2), and semiempirical (AM1, PM3) methods. The results are compared with those obtained by X-ray analysis.
  • Chemical reactivity and molecular shape interactions
    J. Weber, D. Stussi and
    in "New data challenges in our information age" CODATA, Ann Arbor, (1994)
  • The challenge of visualizing microscopic molecular worlds in chemical education
    J. Weber, A. Deloff, O. Schwalm, and P. Flükiger
    in "University education uses of visualization in scientific computing" North-Holland, Amsterdam, (1994), p9-24
  • Computer-aided design of organometallic compounds : a semi-empirical approach of their structure and reactivity
    J. Weber, F. Savary, D. Stussi and
    in "Computers and applied chemistry" Science Press, Beijing, (1993), p36-43
  • Molecular graphics and chemistry
    J. Weber, P. Flükiger and
    Educational media international, 29 (4) (1992), p247-253
  • Development and applications of an extended-Hückel-based reactivity index for organometallic complexes
    J. Weber, D. Stussi, P. Flükiger, and E.P. Kündig
    Comments on inorganic chemistry, 14 (1) (1992), p27-62
    DOI:10.1080/02603599208048656 | Abstract | Article PDF
  • Computer modeling of the activation processes of the aromatic nucleophilic substitution mechanism
    , P. Flükiger and J. Weber
    Journal de Chimie Physique et de Physico-Chimie Biologique, 89 (1992), p1723
  • Molecular graphics investigation of the addition of nucleophiles to (η4:butadiene) M(CO)3 complexes (M=Fe,Co+)
    J. Weber, and O. Eisenstein
    Journal of Molecular Structure: THEOCHEM, 254 (1992), p343-357
    DOI:10.1016/0166-1280(92)80079-2 | Abstract | Article PDF
  • Recent developments in molecular graphics : visualization of chemical structures and properties
    J. Weber, , P. Flükiger and A. Goursot
    The Visual Computer, 7 (2-3) (1991), p158-169
    DOI:10.1007/BF01901186 | Abstract | Article PDF
  • Calculation and visualization of a reactivity index for organometallics based on the extended Hückel model
    J. Weber, P. Flükiger, D. Stussi and
    Journal of Molecular Structure: THEOCHEM, 227 (1991), p175-185
    DOI:10.1016/0166-1280(91)85282-C | Abstract | Article PDF
  • Modeling of organometallic reactivity using a combination of extended Hueckel and molecular graphics techniques
    , P. Flükiger, J. Weber and E.P. Kündig
    Studies in Physical and Theoretical Chemistry, 71 (1990), p313-321
  • Structure and analysis of the binding energy of the copper carbonyl ion CuCO+ complex: an ab initio study
    and J. Weber
    Journal of Molecular Structure: THEOCHEM, 166 (1988), p247-252
    DOI:10.1016/0166-1280(88)80444-5 | Abstract | Article PDF
  • The modeling of nucleophilic and electrophilic additions to organometallic complexes using molecular graphics techniques
    J. Weber, P. Flükiger, , O. Schaad, A. Goursot and C. Daul
    Journal of Computer-Aided Molecular Design, 2 (3) (1988), p235-253
    DOI:10.1007/BF01531997 | Abstract | Article PDF
  • Theoretical and ESR/ENDOR single-crystal study of an azaallyl radical
    M.V.V.S. Reddy, A. Celalyan-Berthier, M. Geoffroy, , J. Weber and G. Bernardinelli
    Journal of the American Chemical Society, 110 (9) (1988), p2748-2753
    DOI:10.1021/ja00217a009 | Abstract | Article PDF

Google

 


Redisplay in format 

                 

    in encoding 

  
Format for journal references
Format for book references
Last update Tuesday March 26 2024