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 15 毫秒
1.
N-Substituted 3-carbamoylpyridinium salts were reduced by glyceraldehyde to give 1,4-dihydronicotinamide derivatives, which may be regarded as a model for oxidation by glyceraldehyde-3-phosphate dehydrogenase.  相似文献   

2.
3.
《Tetrahedron》1986,42(4):961-973
NAD(P)H/NAD(P)+ and related compounds have intrinsic chirality against the axis alone the C3–Ccarbonyl single bond. A model compound which has a stable conformation with respect to this chirality was synthesized and the conformational relationship between the carbonyl dipole and the reacting hydrogen was studied. It has been concluded that there is a relationship between conformations of these groups both in reduction and in oxidation provided the substrate is a neutral species. The result is discussed in relation with the hypothesis proposed previously on the basis of quantum chemical calculations.  相似文献   

4.
Strongly enhanced N2 first positive emission N2(B 3Πg → A 3Σ+u) has been observed on addition of N atoms into a flowing mixture of Cl and HN3. The dependence of the emission intensity on N atom concentration gave a rate constant for the reaction N + N3 → N2(B 3Πg) + N2(X 1Σ+g) of i(1.6 ± 1.1) × 10?11 cm3 molecule?1 s?1. That for the reaction Cl + HN3 → HCl + N3 is (8.9 ± 1.0) × 10?13 cm3 molecule?1 s?1 from the decay of the emission. Comparison of the emission intensity in ClHN3 with that in ClHN3N gave the rate constant of the reaction N3 + N3 → N2(B 3Πg) + 2N2(X 1Σ+g) as 1.4 × 10?12 cm3 molecule?1 s?1 on the assumption that N + N3 yields only N2(B 3Πg) + N2(X 1Σ+g).  相似文献   

5.
The vibrational analysis of the CN(B2Σ+ → X2Σ+) emission sensitized by Hg(63P0) metastables has shown that the energy transfer process, Hg(63P0) + CN(X2Σ+) → Hg(61S0) + CN(B2Σ+), populates the CN(B2Σ+) state in a non-Franck-Condon fashion. The relative vibrational populations for the ν = 0 to 4 states are 1.00, 0.56 ± 0.06, 0.26 ± 0.03, 0.11 ± 0.03 and 0.04 ± 0.01, respectively. Long-range attractive interaction between the Hg(63P0) atom and the CN(X2Σ+) radical is evidenced by the observed high rotational excitation of the CN(B2Σ+) radical following the energy transfer process.  相似文献   

6.
Absolute cross sections were measured for beam attenuation, MX+(1Σg+) chemiionization and MX(A2Π) chemiluminescence. The latter disagree strikingly with predictions based on adiabatic correlations. Information theoretic analysis shows some channels to be statistically, other highly non-statistically populated. A qualitative MO model is in accord with these findings.  相似文献   

7.
The chemiluminescent interaction of Xe(3P2,0) and BrCN has been studied under crosscd-beam conditions at collision energies ranging up to 70 kj mol?1. The CN(B → X) fluorescence spectrum, the excitation function for its production and the fluorescence polarisation - or rather its absence - have been determined. The results can be explained by a two-stage harpooning mechanism involving an inert-gas cyanide (Xe+CN?)1 intermediate but not by a “sensitisation” mechanism proceeding through electronic energy transfer.  相似文献   

8.
Collisions of B+(3P) with H2(X 1Σg+) have been studied repeatedly using molecular-beam techniques. The theoretical interpretation of the results suffered from missing information about the potential-energy surfaces for this system. This paper reports on diatomics-in-molecules calculations for a wide range of nuclear geometries. The stationary points on the minimum-energy paths are determined. Symmetrically orthogonalized and non-hermitean diatomics-in-molecules versions differ only slightly at the minimum-energy path. The potential-energy surfaces of the 13A′ and 23A′ states show dramatical changes with the BHH angle, leading possibly to adiabatic as well as non-adiabatic elementary processes. The 13A′' and 13A′ states are nearly degenerate for geometries ranging from the entrance channel to the interaction region. The most favourable configurations of approach on these potential-energy surfaces are those with C2, symmetry.  相似文献   

9.
The electronic energy transfer process Hg(6 3P0) + OH(X2Πi, υ = 0,K) → Hg(6 1S0) + OH(A 2Σ+, υ,K) has been studied by the sensitized fluorescence method. A rather broad spectrum of rotational population, Nυ′K, was obtained under conditions of minimum relaxation, which illustrates the non-resonant and non-optical nature of this energy transfer process. The fractions of the exoergicity, above electronic excitation of OH(A 2Σ+, υ = 0, K = 0), going into vibrational, rotational and translational excitation are 0.11, 0.31, and 0.58, respectively. A statistical mode of energy partitioning, such as would result from long-lived complex formation, seems to account well for these observations.  相似文献   

10.
The HCl+ (A) vibrational state distributions from the title reaction were studied over the 48–110 meV collision energy range. The 10 and 20 ratios increase by factors of about 1.3 and 2.5, respectively. The branching ratios appear to be determined primarily by Franck-Condon factors, and by the endoergicity of the ν′ = 2 level.  相似文献   

11.
In this work we studied the collinear F(2 P 1/2) + H2 system in a strong CO2 laser field. It was found that the trajectory. surface hopping method (TSHM) yields entirely different results than the exact quantum-mechanical treatment. A curvecrossing model was devised to explain the discrepancy.  相似文献   

12.
A new aspects of the role of the solvent mode in the photoinduced electron-transfer process of electron donor and acceptor system in polar solvents has been exploited. Taking into account the important fact that the vibrational frequency of the solvent mode in the initial neutral state of the reactants is considerably smaller than that in the final ionic state, we have derived a new formula for the energy-gap dependence of the electron-transfer rate. In this formulation, the activation energy is greatly reduced and the electron-transfer rate is almost independent of the energy gap over a wide down-hill energy region. This qualitative feature explains the experimental results for the relation between the bimolecular quenching rate constant kw and the standard free-energy change ΔG° associated with electron transfer in the “anomalous region”.  相似文献   

13.
The binding of the Li+, Na+, K+, Mg2+, and Co2+ ions by 21,31-diphenyl-12,42-dioxo-7,10,13-trioxa-1,4(3,1)-diquinoxaline-2(2,3),3(3,2)-diindolysine-cyclopentadecaphane containing two indolysine fragments, two quinoxaline fragments, and 3,6,9-trioxyundecane spacer in the acetonitrile/0.1 M Bu4NBF4 environment is studied by the method of cyclic voltammetry. It is demonstrated that the Li+, Na+, K+, and Co2+ ions are not bound by this macrocycle, whereas selective redox-switchable binding is observed for the Mg2+ ions. The macrocycle binds the Mg2+ ions way more efficiently as compared with its radical cation and dication. The indolysinequinoxaline fragments play the determining role in the binding. Original Russian Text ? V.V. Yanilkin, N.V. Nastapova, V.A. Mamedov, A.A. Kalinin, V.P. Gubskaya, 2007, published in Elektrokhimiya, 2007, Vol. 43, No. 7, pp. 808–814.  相似文献   

14.
The vibrational distribution of CO produced from the electronic-to-vibrational energy transfer reaction: Na(32P) + CO(X1Σ+, υ=0)→Na(32S) + CO(X1Σ+, υ?8) has been determined by means of infrared resonance absorption measurements employing a cw CO laser. A flash-lamp-pumped dye laser is used to excite the ground state Na to the 32P12 and 32P32 states. The CO molecules formed in the reaction were found to be vibrationally excited up to the limits of available electronic energies carried by the excited Na atoms, and the vibrational population exhibits a maximum at υ=2. The efficiency of E→V energy transfer was determined to be 35%. Our present results were found to be consistent with the impulsive (half-collision) and curve-crossing models.  相似文献   

15.
Lifetimes of selected vibrational levels of the predissociated Ã2Σ+ and Ã2Π electronic states of N2O+ and COS+, re- spectively, have been measured. These values have been used in conjunction with previous data on fluorescence quantum yields to obtain predissociation rates for the various vibrational levels.  相似文献   

16.
Interaction potentials for CaCl(X 2Σ+)-Ar and KCl(X 1Σ+)-Ar have been determined. They include a Gordon-Kim electron-gas repulsive part smoothly joined to the long-range van der Waals potential. The van der waals potential for KClAr was taken from Meyer and Toennies. For CaClAr, the necessary molecular parameter were estimated from the Rittner model, which predicts both the dipole and quadrupole moments fairly accurately. The CaClAr interaction potential is quite different from that of KClAr. Due to the outer 4s electron on the Ca+ ion. the CaClAr potential exhibits a deep minimum in the odd-order Legendre terms which is expected to have a large effect on the cross sections for collisional rotational excitation. The KClAr potential determined here also shows significant differences in the repulsive and well regions from that predicted by Meyer and Toennies using a site-site model for the repulsive contribution.  相似文献   

17.
Steady-state and transient photokinetic and spectroscopic measurements on aqueous Eu(NO3)3 show different affinities of 7F, 5D1 and 5D0 Eu3+aq towards nitrate ion. This may be rationalised by differences in the inner- and outer-shell hydration structures between 5DO, 5D1 Eu3+(aq) and 7F Eu3+(aq). Nitrate penetration into the inner-shell of Eu3+(aq), and inner-coordination (EuNO2+3)* exciplex formation, occur solely in the long-lived 5DO level of Eu3+(aq).  相似文献   

18.
A MRD CI procedure has been used to calculate several electronic states of the hydroperoxyl radical. The basis set is of double-zeta plus polarization quality augmented with s- and p-type bond and Rydberg functions. The vertical excitation energies of the lowest eight doublet and six quartet states are reported. Oscillator strengths for transitions form the ground to upper doublet states were calculated. A cut of the potential energy surfaces along the OOH fragmentation pathway is used to discuss the mechanisms of HO2 photodissociation below 6.4 eV. Arguments are presented which indicate O(1D) rather than O(3P) is the primary dissociation product, and so support the experimental findings rather than theory in the conflict raised earlier on this matter. Ostensibly the dissociation proceeds diabatically on the surface of the initially populated 2A″(1a″ → 2a″) state yielding OH(X2II) + O(1D).  相似文献   

19.
The reactions of the lowest metastable states of Ar, Kr and Xe with XeF2 were studied in a flowing afterglow apparatus; XeF emission (from D2Π12 and B 2Π+ states) was observed in all cases. The total rate constants (cm3 molecule?1 s?1) for XeF* formation were determined as 75 × 10?11 ? Xe(3P2);64 × 10?11 ? Kr(3P2) and 20 × 10?11 ? Ar(3P0,2). The reactions of Ar(3P0,2) and Kr(3P2) with XeF2 also gave ArF* and KrF*, respectively. Analysis of these emissions indicates that at least two different mechanisms are operative: reactive quenching by the ionic—covalent curve-crossing mechanism and excitation transfer. The Ar(3P0,2 + XeF2 reaction is a sufficiently strong source of XeF(D—X) emission that the main features of the XeF(D2Π12 ? X2Σ+) system could be photographed and tentative assignments of these vibrational bands are given. The XeF(D → B) emission could not be observed and the ratio of the D—X versus the D—B transition probability must be > 1000 : 1.  相似文献   

20.
Triplet methylene, CH2(3B1), and methyl radicals were produced by flash photolysis of a mixture of ketene and azomethane. A computer fit of the product ratios, using the known rate constants for CH2 + CH2, and CH3 + CH3, requires a rate constant of 5.0 × 10?11 cm3 molecule?1s?1 for the reaction CH2 + CH3 ? C2H4 + H.  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