首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The optimality of MO basis sets of Gaussian functions, when constructed from AO basis sets optimized for the neutral atom or for atom ions, is investigated. A formal charge parameter Q is defined and used to adjust the AO basis sets to the molecular environment, by virtue of a simple quadratic expression. Calculations on a series of C1 hydrocarbons (CH2, CH3, CH3+, CH3?, CH4) using 3G basis sets indicate considerable variations in the optimum Q value with the molecular species. The proposed method offers a simple alternative technique to a full molecular basis set optimization.  相似文献   

2.
Ab initio molecular orbital theory using basis sets up to 6-311G* *, with electron correlation incorporated via configuration interaction calculations with single and double substitutions, has been used to study the structures and energies of the C3H2 monocation and dication. In agreement with recent experimental observations, we find evidence for stable cyclic and linear isomers of [C3H2]+ ˙. The cyclic structure (, a) represents the global minimum on the [C3H2]+ ˙ potential energy surface. The linear isomer (, b) lies somewhat higher in energy, 53 kJ mol?1 above a. The calculated heat of formation for [HCCCH]+ ˙ (1369 kJ mol?1) is in good agreement with a recent experimental value (1377 kJ mol?1). For the [C3H2]2+ dication, the lowest energy isomer corresponds to the linear [HCCCH]2+ singlet (h). Other singlet and triplet isomers are found not to be competitive in energy. The [HCCCH]2+ dication (h) is calculated to be thermodynamically stable with respect to deprotonation and with respect to C? C cleavage into CCH+ + CH+. The predicted stability is consistent with the frequent observation of [C3H2]2+ in mass spectrometric experiments. Comparison of our calculated ionization energies for the process [C3H2]+ ˙ → [C3H2]2+ with the Qmin values derived from charge-stripping experiments suggests that the ionization is accompanied by a significant change in structure.  相似文献   

3.
We have measured the synchrotron‐induced photofragmentation of isolated 2‐deoxy‐D ‐ribose molecules (C5H10O4) at four photon energies, namely, 23.0, 15.7, 14.6, and 13.8 eV. At all photon energies above the molecule′s ionization threshold we observe the formation of a large variety of molecular cation fragments, including CH3+, OH+, H3O+, C2H3+, C2H4+, CHxO+ (x=1,2,3), C2HxO+ (x=1–5), C3HxO+ (x=3–5), C2H4O2+, C3HxO2+ (x=1,2,4–6), C4H5O2+, C4HxO3+ (x=6,7), C5H7O3+, and C5H8O3+. The formation of these fragments shows a strong propensity of the DNA sugar to dissociate upon absorption of vacuum ultraviolet photons. The yields of particular fragments at various excitation photon energies in the range between 10 and 28 eV are also measured and their appearance thresholds determined. At all photon energies, the most intense relative yield is recorded for the m/q=57 fragment (C3H5O+), whereas a general intensity decrease is observed for all other fragments— relative to the m/q=57 fragment—with decreasing excitation energy. Thus, bond cleavage depends on the photon energy deposited in the molecule. All fragments up to m/q=75 are observed at all photon energies above their respective threshold values. Most notably, several fragmentation products, for example, CH3+, H3O+, C2H4+, CH3O+, and C2H5O+, involve significant bond rearrangements and nuclear motion during the dissociation time. Multibond fragmentation of the sugar moiety in the sugar–phosphate backbone of DNA results in complex strand lesions and, most likely, in subsequent reactions of the neutral or charged fragments with the surrounding DNA molecules.  相似文献   

4.
Optimum geometries and stabilization energies are determined for complexes of H2O, NH3, CH4, C2H4, CO, and N2 with metal cations including Li+, Na+, K+, Rb+, Be2+, Mg2+, Ca2+, Zn2+, and Al3+, for the complex (HO)2PO 2 ...Mg2+ and for the complexes of water with F, Cl, and Br by SCF calculations employing the MINI-1 minimal gaussian basis sets. The Boys-Bernardi method was used to evaluate the superposition error. Comparison with the extended basis set results revealed that the MINI-1 set gives uniformly good results for a broad variety of ionic complexes and therefore should be preferred to other small basis sets.  相似文献   

5.
The gas phase heats of formation of several organosulfur cations were determined from thiirane, thietane and tetrahydrothiophene precursor molecules by photoionization mass spectrometry. Heats of formation at 0 K and 298 K are reported for the following ions: [H2CS], [H3CS]+, [C2H3S]+, [C2H4S], [C3H5S]+, [C3H6S], [C4H7S]+ and [C4H8S]. The [C4H7S]+ (m/z 87), [C2H4S] (m/z 60), [C2H3S]+ (m/z 59), [C4H7]+ (m/z 55), [C4H6] (m/z 54) and [CH2S] (m/z 46) ions are produced from metastable tetrahydrothiophene ions at photon energies between 10.2 and 10.7 eV. Decay rates of internal energy selected parent ions to the m/z 60, 59, 55 and 54 fragments were measured by threshold photoelectron-photoion coincidence, the results of which are compared to statistical theory (RRKM/QET) calculations. The [C2H4S] ion from tetrahydrothiophene is found to have the thioacetaldehyde structure. From the measured [C2H4S] onset a ΔH = 50±8 kJ mol?1 was calculated for the thioacetaldehyde molecule.  相似文献   

6.
The structures of the 1:1 proton‐transfer compounds of isonipecotamide (piperidine‐4‐carboxamide) with 4‐nitrophthalic acid [4‐carbamoylpiperidinium 2‐carboxy‐4‐nitrobenzoate, C6H13N2O8+·C8H4O6, (I)], 4,5‐dichlorophthalic acid [4‐carbamoylpiperidinium 2‐carboxy‐4,5‐dichlorobenzoate, C6H13N2O8+·C8H3Cl2O4, (II)] and 5‐nitroisophthalic acid [4‐carbamoylpiperidinium 3‐carboxy‐5‐nitrobenzoate, C6H13N2O8+·C8H4O6, (III)], as well as the 2:1 compound with terephthalic acid [bis(4‐carbamoylpiperidinium) benzene‐1,2‐dicarboxylate dihydrate, 2C6H13N2O8+·C8H4O42−·2H2O, (IV)], have been determined at 200 K. All salts form hydrogen‐bonded structures, viz. one‐dimensional in (II) and three‐dimensional in (I), (III) and (IV). In (I) and (III), the centrosymmetric R22(8) cyclic amide–amide association is found, while in (IV) several different types of water‐bridged cyclic associations are present [graph sets R42(8), R43(10), R44(12), R33(18) and R64(22)]. The one‐dimensional structure of (I) features the common `planar' hydrogen 4,5‐dichlorophthalate anion, together with enlarged cyclic R33(13) and R43(17) associations. In the structures of (I) and (III), the presence of head‐to‐tail hydrogen phthalate chain substructures is found. In (IV), head‐to‐tail primary cation–anion associations are extended longitudinally into chains through the water‐bridged cation associations, and laterally by piperidinium–carboxylate N—H...O and water–carboxylate O—H...O hydrogen bonds. The structures reported here further demonstrate the utility of the isonipecotamide cation as a synthon for the generation of stable hydrogen‐bonded structures. An additional example of cation–anion association with this cation is also shown in the asymmetric three‐centre piperidinium–carboxylate N—H...O,O′ interaction in the first‐reported structure of a 2:1 isonipecotamide–carboxylate salt.  相似文献   

7.
A theoretical study of the halogenated cations of mono-, di-, tri- and tetramethyl-substituted ethylenes, C3H6X+, C4H8X+, C5H10X+ and C6H12X+, X=F, Cl, Br, have been studied at the ab initio MP2 and density functional B3LYP levels of theory implementing 6-311++G(d,p) basis set. The potential energy surfaces of all molecules under investigation have been scanned and the 13C and 1H NMR chemical shifts for all the bridged halonium ions studied have been calculated using the GIAO method at the B3LYP level. The calculated halogen binding energies in the halonium ions have been correlated with the experimental rates of chlorination and bromination of the corresponding alkenes. The computed hydride affinities and the NICS values for the bridged cations show that the bromo cations are more stable than the analogous chloro and fluoro cations.  相似文献   

8.
The title compounds are salts of the general form (Q+)2[Zn(dmit)2]2?, where dmit corresponds to the ligand (C3S5)? present in both and Q+ to the counter‐cations (nBu4N)+ [or C16H36N+] and (Ph4As)+ [or C24H20As+], respectively. In the first case, Zn is in the 4e special positions of space group C2/c and hence the [Zn(dmit)2]2? dianion possesses twofold axial crystallographic symmetry. Including these, there are now 11 known examples of [Zn(dmit)2]2? or its analogues, with O replacing the exocyclic thione S, and [Zn(dmio)2]2? dianions in nine structures with various Q. Comparison of these reveals a remarkable variation in details of the conformation which the dianion may adopt in terms of Zn coordination, equivalence of the Zn—S bond lengths, displacement of Zn from the plane of the ligand and overall dianion shape.  相似文献   

9.
This study undertakes a theoretical investigation into uncommon hydrogen bonds between the ethyl cation (C2H5 +) and π hydrocarbons. Firstly, it considers the hyperconjugation effect of the ethyl cation, in which the non-localized hydrogen (H+) is taken to be a pseudoatom bound to the carbons of the methyl groups. The goal of the research is to use this electronic phenomenon to gain a better understanding of the (H+···π) and (H+···p-π) hydrogen bonds, which are considered uncommon because they are formed through the interaction of the H+ of the ethyl cation with the π bonds of the acetylene (C2H2) and ethene (C2H4), as well as with the pseudo-π bond of the cyclopropane (C3H6). In view of this, B3LYP/6-311++G(d,p) calculations were used to determine the geometries of the C2H5 +···C2H2, C2H5 +···C2H4, and C2H5 +···C3H6 hydrogen-bonded complexes. Deformations of the bond lengths and bond angles of these systems were analyzed geometrically. Examination of the stretch frequencies and absorption intensities of the (H+···π) and (H+···p-π) hydrogen bonds has revealed red-shifts in π and p-π bonds. After structural modeling and vibrational characterization, analysis of the charge transfer following the ChelpG approach and subsequently quantification of the hydrogen bond energies (basis sets superpostition error and zero point vibrational energies being considered) were used to predict the strength of the (H+···π) and (H+···p-π) hydrogen bonds. In addition, the molecular topography was estimated using the quantum theory of atoms in molecules (QTAIM). QTAIM was chosen because of a desire to understand the (H+···π) and (H+···p-π) hydrogen bonds chemically on the basis of the quantity of charge density and interpretation of Laplacian fields. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

10.
In this study ab initio Car–Parrinello molecular dynamics simulations, extended transition state (ETS)‐natural orbitals for chemical valence (NOCV) and QTAIM bonding analyses, were performed to characterize the ansa‐bridged molybdocene complexes [(C5H4)2XMe2MoH3]+ for X = C, Si, Ge, Sn, Pb, and nonbridged Cp2MoH system. The results have shown that the [(C5H4)2CMe2MoH(H2)]+ complex exhibits nonclassical dihydrogen/hydride (H2/H) conformation (97.6% of time of simulation), contrary to trihydride (H3) structure noted for nonbridged Cp2MoH (86.9%) and ansa‐bridged [(C5H4)2SnMe2MoH3]+ (84.8%), [(C5H4)2PbMe2MoH3]+ (84.9%) systems. Further, [(C5H4)2SiMe2MoH3]+ and [(C5H4)2GeMe2MoH3]+ complexes, appeared to exist in both conformations (H2/H—55.4%, H3—44.6% for Si‐based system and H2/H—36.2%, H3—63.8 % for germanium congener). It has been proven that the “steric availability” of the metal center, measured by the changes in the Cp? Mo? Cp angle (α), determines the existence of a given conformation—namely, the smaller value of the angle (molybdenum is sterically more accessible) the larger preference for the formation of dihydrogen/hydride structure. ETS‐NOCV method allowed to conclude that increase in the Cp? Mo? Cp angle (from α ca. 120° to α ca. 150°) leads to the enhancement of donation from H2 and back‐donation from Mo to the σ*(H? H), what consequently leads to breaking of the H? H bond and formation of the trihydride structure. Systematical increase in the charge depletion from the σ‐bonding orbital of H2 can be related to the reduction of the energy gap between the major orbitals involved in this contribution, namely highest occupied molecular orbital (HOMO) of H2 with lowest unoccupied molecular orbital (LUMO) of [MoHL]+; ΔE = 0.0868 a.u. [for L =(C5H4)2C], ΔE = 0.0827a.u. [for L = (C5H4)2Si] ΔE = 0.0638 a.u. [for L = Cp2]. Further, the relatively low energetic barrier to hydrogen exchange (ΔE# = 3.3 kcal/mol) for carbon‐bridged complex, [(C5H4)2CMe2MoHc(HaHb)]+ → [(C5H4)2 CMe2MoHa(HbHc)]+, is related to strengthening of the Mo–H bonds when going from the substrate to the transition state (TS). Notably higher barrier to hydrogen rotation (ΔE# = 10.1 kcal/mol) in [(C5H4)2CMe2MoH(H2)]+ is due to lowering in the electrostatic stabilization as well as weakening of the donation (H2 → Mo charge transfer) and practically lack‐of back‐donation (Mo → H2) in the rotated TS. © 2012 Wiley Periodicals, Inc.  相似文献   

11.
Chromium(III)-isonicotinate complexes, cis-[Cr(C2O4)2(N-inic)(H2O)]- and [Cr(C2O4)(H2O)3-OH-Cr(C2O4)2(O-inic)]-(N-inic)(H2 (N-inic = N-bonded and O-inic = O-bonded isonicotinic acid) were obtained and characterized in solution. Kinetics of acid-catalyzed isonicotinate ligand liberation were studied spectrophotometrically in the 0.1–1.0 m HClO4 range, at I=1.0 m. The dependencies of the pseudo-first order rate constant on [H+] were established: kobs = k0+kHQH[H+] and kobs = kHQH[H+] for the N-inic and O-inic complex, respectively, where k0 and kH are the rate constants of the spontaneous and the acid-catalyzed reaction paths, and QH is the protonation constant of the carboxylic group in isonicotinic ligand. The obtained results indicate that N-bonded isonicotinic acid liberation occurs mainly via a spontaneous reaction path and is much slower than O-bonded inic liberation. The mechanisms for these processes are proposed.  相似文献   

12.
The Na[Cr(PDA)2] · 2H2O complex (PDA1 = dipicolinic acid anion) and its aquation product, [Cr(PDA)(H2O)3]+, were prepared and characterized. The electronic spectra demonstrate that the bis(dipicolinato) complex undergoes very fast partial dechelation during dissolution. In acidic media, pH controlled, rapid protolytic and ring opening processes lead to coexistence of complexes with one tridentate (PDA) and the other bi- or mono-dentate (PDA). The kinetics of PDA ligand liberation were followed spectrophotometrically within the 0.1–2.0 M HClO4 range at I = 2.0 M. The observed first-order rate constant depends on [H+] according to the equation: k obs = A[H+]/(1 + B[H+] + C[H+]2). A reaction course via the uncharged [Cr(PDA)(HPDA)(H2O)2]0 complex is proposed. The observed rate increase, followed by rate retardation with [H+] increase, is attributed to the unreactive [Cr(PDA)(H2PDA)(H2O)2]+ complex. In terms of the proposed mechanism, A, B, C parameters have been defined as: A = k 1 Q 1, B = Q 1, C = Q 1 Q 2 where k 1 is the rate constant of the CrIII-carboxylato oxygen bond-breaking in the monodentate HPDA ligand, Q 1 is a composite value describing protolytic and dechelation processes and Q 2 is the protonation constant of the uncharged [Cr(PDA)(HPDA)(H2O)2]0 complex.  相似文献   

13.
Ab initio molecular orbital calculations with split-valence plus polarization basis sets and incorporating valence-electron correlation have been performed to determine the equilibrium structure of ethyloxonium ([CH3CH2OH2]+) and examine its modes of unimolecular dissociation. An asymmetric structure (1) is predicted to be the most stable form of ethyloxonium, but a second conformational isomer of Cs symmetry lies only 1.4 kJ mol?1 higher in energy than 1. Four unimolecular decomposition pathways for 1 have been examined involving loss of H2, CH4, H2O or C2H4. The most stable fragmentation products, lying 65 kJ mol?1 above 1, are associated with the H2 elimination reaction. However, large barriers of 257 and 223 kJ mol?1 have to be surmounted for H2 and CH4 loss, respectively. On the other hand, elimination of either C2H4 or H2O from ethyloxonium can proceed without a barrier to the reverse associations and, with total endothermicities of 130 and 160 kJ mol?1, respectively, these reactions are expected to dominate at lower energies. A second important equilibrium structure on the surface is a hydrogen-bridged complex, lying 53 kJ mol?1 above 1. This complex is involved in the C2H4 elimination reaction, acts as an intermediate in the proton-transfer reaction connecting [C2H5]+ +H2O and C2H4 + [H3O]+ and plays an important role in the isotopic scrambling that has been observed experimentally in the elimination of either H2O or C2H4 from ethyloxonium. The proton affinity of ethanol was calculated as 799 kJ mol?1, in close agreement with the experimental value of 794 kJ mol?1.  相似文献   

14.
Four organic salts, namely benzamidinidium orotate (2,6‐dioxo‐1,2,3,6‐tetrahydropyrimidine‐4‐carboxylate) hemihydrate, C7H9N2+·C5H3N2O4·0.5H2O (BenzamH+·Or), (I), benzamidinium isoorotate (2,4‐dioxo‐1,2,3,4‐tetrahydropyrimidine‐5‐carboxylate) trihydrate, C7H9N2+·C5H3N2O4·3H2O (BenzamH+·Isor), (II), benzamidinium diliturate (5‐nitro‐2,6‐dioxo‐1,2,3,6‐tetrahydropyrimidin‐4‐olate) dihydrate, C7H9N2+·C4H2N3O5·2H2O (BenzamH+·Dil), (III), and benzamidinium 5‐nitrouracilate (5‐nitro‐2,4‐dioxo‐1,2,3,4‐tetrahydropyrimidin‐1‐ide), C7H9N2+·C4H2N3O4 (BenzamH+·Nit), (IV), have been synthesized by a reaction between benzamidine (benzenecarboximidamide or Benzam) and the appropriate carboxylic acid. Proton transfer occurs to the benzamidine imino N atom. In all four acid–base adducts, the asymmetric unit consists of one tautomeric aminooxo anion (Or, Isor, Dil and Nit) and one monoprotonated benzamidinium cation (BenzamH+), plus one‐half (which lies across a twofold axis), three and two solvent water molecules in (I), (II) and (III), respectively. Due to the presence of protonated benzamidine, these acid–base complexes form supramolecular synthons characterized by N+—H...O and N+—H...N (±)‐charge‐assisted hydrogen bonds (CAHB).  相似文献   

15.
Two CrIII–picolinato complexes were obtained and characterized in solution. The [Cr(C2O4)(pyac)2] and [Cr(C2O4)2(pyac)]2– ions (pyac = picolinic acid anion) in acidic solutions undergo a reversible one-end CrIII–picolinato chelate ring opening via CrIII—N bond breaking. The reaction rate was determined spectrophotometrically in the 0.1–1.0 M HClO4 range at I = 1.0 M. The observed pseudo-first order rate constant depends on [H+] according to the equation: k obs = a + b[H+] + c/[H+]. A reaction mechanism, which assumes participation of the protonated and unprotonated forms of the reactants, has been proposed. The kinetic parameters a, b, c have been defined as a = k 1, b = k 2 Q 1, c = k –1/Q 2, where k 1, k –1,k 2 are rate constants for the forward and reverse processes and Q 1, Q 2 are the protolytic equilibrium constants in the term of the proposed mechanism. The activation parameters have been determined and discussed.  相似文献   

16.
New chromium(III) complexes, [Cr(C2O4)2(2-hnic)]2− and [Cr(C2O4)2(3-hpic)]2− (where 2-hnic = O,O′-bonded 2-hydroxynicotinic acid and 3-hpic = N,O-bonded 3-hydroxypicolinic acid), were obtained and characterized in solution. The acid-catalyzed aquation of the both complexes leads to liberation of the appropriate pyridinecarboxylic acid and formation of cis-[Cr(C2O4)2(H2O)2]. Kinetics of these reactions were studied spectrophotometrically in the 0.1–1.0 M HClO4 range, at I = 1.0 M. In the case of [Cr(C2O4)2(2-hnic)]2−, a slow chelate-ring opening at the Cr–O (phenolate) bond is followed by a fast Cr–O (carboxylate) bond breaking. The rate law: kobs = kHQH[H+] was established, where kH is the acid-catalyzed rate constant and QH is the protonation constant of the coordinated phenolate oxygen atom. In the case of [Cr(C2O4)2(3-hpic)]2−, the reversible chelate-ring opening at Cr–N bond is followed by the rate determining step – the one-end bonded ligand liberation. The rate law for the first step was determined: kobs = k1+k−1/Q1[H+], where k1 and k−1 are the rate constants of the chelate-ring opening and closure and Q1 is the protonation constant of the pyridine nitrogen atom. The aquation mechanisms are proposed and the effect of ligand coordination mode on complex reactivity is discussed.  相似文献   

17.
Bis(5‐chloro‐8‐hydroxyquinolinium) tetrachloridopalladate(II), (C9H7ClNO)2[PdCl4], (I), catena‐poly[dimethylammonium [[dichloridopalladate(II)]‐μ‐chlorido]], {(C2H8N)[PdCl3]}n, (II), ethylenediammonium bis(5‐chloroquinolin‐8‐olate), C2H10N22+·2C9H5ClNO, (III), and 5‐chloro‐8‐hydroxyquinolinium chloride, C9H7ClNO+·Cl, (IV), were synthesized with the aim of preparing biologically active complexes of PdII and NiII with 5‐chloroquinolin‐8‐ol (ClQ). Compounds (I) and (II) contain PdII atoms which are coordinated in a square‐planar manner by four chloride ligands. In the structure of (I), there is an isolated [PdCl4]2− anion, while in the structure of (II) the anion consists of PdII atoms, lying on centres of inversion, bonded to a combination of two terminal and two bridging Cl ligands, lying on twofold rotation axes, forming an infinite [–μ2‐Cl–PdCl2–]n chain. The negative charges of these anions are balanced by two crystallographically independent protonated HClQ+ cations in (I) and by dimethylammonium cations in (II), with the N atoms lying on twofold rotation axes. The structure of (III) consists of ClQ anions, with the hydroxy groups deprotonated, and centrosymmetric ethylenediammonium cations. On the other hand, the structure of (IV) consists of a protonated HClQ+ cation with the positive charge balanced by a chloride anion. All four structures are stabilized by systems of hydrogen bonds which occur between the anions and cations. π–π interactions were observed between the HClQ+ cations in the structures of (I) and (IV).  相似文献   

18.
Conducting organic polymers (COPs) are made of a conjugated polymer backbone supporting a certain degree of oxidation. These positive charges are compensated by the doping anions that are introduced into the polymer synthesis along with their accompanying cations. In this work, the influence of these cations on the stoichiometry and physicochemical properties of the resulting COPs have been investigated, something that has previously been overlooked, but, as here proven, is highly relevant. As the doping anion, metallacarborane [Co(C2B9H11)2] was chosen, which acts as a thistle. This anion binds to the accompanying cation with a distinct strength. If the binding strength is weak, the doping anion is more prone to compensate the positive charge of the polymer, and the opposite is also true. Thus, the ability of the doping anion to compensate the positive charges of the polymer can be tuned, and this determines the stoichiometry of the polymer. As the polymer, PEDOT was studied, whereas Cs+, Na+, K+, Li+, and H+ as cations. Notably, with the [Co(C2B9H11)2] anions, these cations are grouped into two sets, Cs+ and H+ in one and Na+, K+, and Li+ in the second, according to the stoichiometry of the COPs: 2:1 EDOT/[Co(C2B9H11)2] for Cs+ and H+, and 3:1 EDOT/[Co(C2B9H11)2] for Na+, K+, and Li+. The distinct stoichiometries are manifested in the physicochemical properties of the COPs, namely in the electrochemical response, electronic conductivity, ionic conductivity, and capacitance.  相似文献   

19.
OH+ is an extraordinarily strong oxidant. Complexed forms (L? OH+), such as H2OOH+, H3NOH+, or iron–porphyrin‐OH+ are the anticipated oxidants in many chemical reactions. While these molecules are typically not stable in solution, their isolation can be achieved in the gas phase. We report a systematic survey of the influence on L on the reactivity of L? OH+ towards alkanes and halogenated alkanes, showing the tremendous influence of L on the reactivity of L? OH+. With the help of with quantum chemical calculations, detailed mechanistic insights on these very general reactions are gained. The gas‐phase pseudo‐first‐order reaction rates of H2OOH+, H3NOH+, and protonated 4‐picoline‐N‐oxide towards isobutane and different halogenated alkanes CnH2n+1Cl (n=1–4), HCF3, CF4, and CF2Cl2 have been determined by means of Fourier transform ion cyclotron resonance meaurements. Reaction rates for H2OOH+ are generally fast (7.2×10?10–3.0×10?9 cm3 mol?1 s?1) and only in the cases HCF3 and CF4 no reactivity is observed. In contrast to this H3NOH+ only reacts with tC4H9Cl (kobs=9.2×10?10), while 4‐CH3‐C5H4N‐OH+ is completely unreactive. While H2OOH+ oxidizes alkanes by an initial hydride abstraction upon formation of a carbocation, it reacts with halogenated alkanes at the chlorine atom. Two mechanistic scenarios, namely oxidation at the halogen atom or proton transfer are found. Accurate proton affinities for HOOH, NH2OH, a series of alkanes CnH2n+2 (n=1–4), and halogenated alkanes CnH2n+1Cl (n=1–4), HCF3, CF4, and CF2Cl2, were calculated by using the G3 method and are in excellent agreement with experimental values, where available. The G3 enthalpies of reaction are also consistent with the observed products. The tendency for oxidation of alkanes by hydride abstraction is expressed in terms of G3 hydride affinities of the corresponding cationic products CnH2n+1+ (n=1–4) and CnH2nCl+ (n=1–4). The hypersurface for the reaction of H2OOH+ with CH3Cl and C2H5Cl was calculated at the B3 LYP, MP2, and G3m* level, underlining the three mechanistic scenarios in which the reaction is either induced by oxidation at the hydrogen or the halogen atom, or by proton transfer.  相似文献   

20.
Cyclic polysulfides isolated from higher plants, model compounds and their electron impact induced fragment ions have been investigated by various mass spectrometric methods. These species represent three sets of sulfur compounds: C3H6Sx (x=1?6), C2H4Sx (x=1?5) and CH2Sx (x=1?4). Three general fragmentation mechanisms are discussed using metastable transitions: (1) the unimolecular loss of structural parts (CH2S, CH2 and Sx); (2) fragmentations which involve ring opening reactions, hydrogen migrations and recyclizations of the product ions ([M? CH3]+, [M? CH3S]+, [M? SH]+ and [M? CS2]); and (3) complete rearrangements preceding the fragmentations ([M? S2H]+ and [M? C2H4]). The cyclic structures of [M] and of specific fragment ions have been investigated by comparing the collisional activation spectra of model ions. On the basis of these results the cyclic ions decompose via linear intermediates and then recyclizations of the product ions occur. The stabilities of the fragment ions have been determined by electron efficiency vs electron energy curves.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号