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.
《Tetrahedron》1986,42(22):6225-6234
Ab initio molecular orbital calculations on the distonic radical cations CH2(CH2)nN+H3 and their conventional isomers CH3(CH2)nNH2+ (n = 0,1, 2 and 3) indicate a preference in each case for the distonic isomer. The energy difference appears to converge with increasing n towards a limit which is close to the energy difference between the component systems CH3·H2+CH3+NH3 (representing the distonic isomer) and CH3CH3+CH3NH2+ (representing the conventional isomer). The generality of this result is assessed by using results for the component systems CH3·Y+CH3X+H and CH3YH+CH3X+. (or CH3YH+. + CH3X) to predict the relative energies of the distonic ions ·Y(CH2)nX+H and their conventional isomers HY(CH2)nX+. (X = NH2, OH, F, PH2, SH, Cl; Y = CH2, NH, O) and testing the predictions through explicit calculations for systems with n = 0,1 and 2. Although the predictions based on component systems are often close to the results of direct calculations, there are substantial discrepancies in a number of cases; the reasons for such discrepancies are discussed. Caution must be exercised in applying this and related predictive schemes. For the systems examined in the present study, the conventional radical cation is predicted in most cases to lie lower in energy than its distonic isomer. It is found that the more important factors contributing to a preference for distonic over conventional radical cations are the presence in the system of a group(X) with high proton affinity and the absence of a group (X, Y or perturbed (C—C) with low ionization energy.  相似文献   

2.
High‐level ab initio calculations show that the formation of radicals, by the homolytic bond fission of Y?R (Y=F, OH, NH2; R=CH3, NH2, OH, F, SiH3, PH2, SH, Cl, NO) bonds is dramatically favored by the association of the molecule with BeX2 (X=H and Cl) derivatives. This finding is a consequence of two concomitant effects, the significant activation of the Y?R bond after the formation of the beryllium bond, and the huge stabilization of the F. (OH., NH2.) radical upon BeX2 attachment. In those cases where R is an electronegative group, the formation of the radicals is not only exergonic, but spontaneous.  相似文献   

3.
A computational study at CCSD(T) theoretical level has been carried out on radical cation [(PH2X)2]·+ homodimers. Four stable minima configurations have been found for seven substituted phosphine derivatives, X = H, CH3, CCH, NC, OH, F and Cl. The most stable minimum presents an intermolecular two-center three-electron P···P bond except for X = CCH. The other three minima correspond to an alternative P···P pnicogen bonded complex, to a P···X contact and the last one to the complex resulting from a proton transfer, PH3X+:PHX·. The complexes obtained have been compared with those of the corresponding neutral ones, (PH2X)2, and the analogous protonated ones, PH3X+:PH2X, recently described in the literature. The spin and charge densities of the complexes have been examined. The electronic characteristics of the complexes have been analyzed with the NBO and AIM methods. The results obtained for the spin density, charge and NBO are coherent for all the complexes.  相似文献   

4.
The resonance energies (REs) of neutral three membered ring analogs of the cyclopropenyl cation, computed using block localized wave function (BLW) methods, reveal considerable variations. The RE's of cyclopropenes substituted with exocyclic double bonded groups C?X, (X = O, NH, CH2, S, PH, SiH2) increase with the electronegativity of X in the same row (SiH2 < PH < S and CH2 < NH < O). The extra cyclic resonance energies (ECREs) (an energetic measure of aromaticity based on comparisons with the RE's of acyclic models) of these derivatives range from +10.5 kcal/mol for cyclopropenone (X = O) (somewhat aromatic; the benzene ECRE is 29.3 kcal/mol) to ?2.4 kcal/mol (slightly antiaromatic) for X = SiH2. Additional disubstitution of the C?C double bond by X′ groups (X′ = CH3, NH2, OH, SiH3, PH2, SH) increases the REs considerably, but has only small effects on the ECREs. Even the ECRE of deltic acid (X = O, X′ = OH) is only increased to +13.3 kcal/mol. The conclusion based on ECRE's, that all 12 of the three membered rings are only marginally aromatic/antiaromatic, is supported by the satisfactorily plot (R2 = 0.92) of ECRE against values of NICS(0)πzz (a superior nucleus chemical independent shift magnetic index of aromaticity), which range only from ?6.1 ppm (diatropic) for deltic acid (cf., ?35.5 ppm for benzene and ?14.2 ppm for the parent cyclopropenium ion) to +8.9 ppm (paratropic) for the silicon derivative, X = SiH2, X′ = SiH3. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

5.
A 4 K matrix ESR study shows that the molecular radical cations of isopropyl formate and acetate, produced radiolytically in halocarbon matrices at 4.2 K, undergo spontaneous rearrangement due to a selective intramolecular hydrogen shift from the tertiary CH bond in the isopropyl group to the carbonyl oxygen atom giving RC+(OH)OC(CH3)2, where R = H or CH3. The radical cation of tert-butyl acetate undergoes further fragmentation at the ester CO bond following a similar rearrangement to give an isobutene radical cation in CFCl3.  相似文献   

6.
Examination of the reactions of the long-lived (>0.5-s) radical cations of CD3CH2COOCH3 and CH3CH2COOCD3 indicates that the long-lived, nondecomposing methyl propionate radical cation CH3CH2C(O)OCH 3 isomerizes to its enol form CH3CH=C(OH)OCH 3 H isomerization ? ?32 kcal/mol) via two different pathways in the gas phase in a Fourier-transform ion cyclotron resonance mass spectrometer. A 1,4-shift of a β-hydrogen of the acid moiety to the carbonyl oxygen yields the distonic ion ·CH2CH2C+ (OH)OCH3 that then rearranges to CH3CH=C(OH)OCH 3 probably by consecutive 1,5- and 1,4-hydrogen shifts. This process is in competition with a 1,4-hydrogen transfer from the alcohol moiety to form another distonic ion, CH3CH2C+(OH)OCH 2 · , that can undergo a 1,4-hydrogen shift to form CH3CH=C(OH)OCH 3 . Ab initio molecular orbital calculations carried out at the UMP2/6-31G** + ZPVE level of theory show that the two distonic ions lie more than 16 kcal/mol lower in energy than CH3CH2C(O)OCH 3 . Hence, the first step of both rearrangement processes has a great driving force. The 1,4-hydrogen shift that involves the acid moiety is 3 kcal/mol more exothermic (ΔH isomerization=?16 kcal/mol) and is associated with a 4-kcal/mol lower barrier (10 kcal/mol) than the shift that involves the alcohol moiety. Indeed, experimental findings suggest that the hydrogen shift from the acid moiety is likely to be the favored channel.  相似文献   

7.
Abstract

Letcher and Van Wazers suggestion[1] that the 31P NMR chemical shifts of phosphines might be related to the substituent electronegativity, EN(X)[2], has not been verified subsequently by the available experimental data[3,4]. We now have explored this relationship systematically by means of reliable[5] ab initio magnetic property calculations[6] on a comprehensive set of molecules: PXY2 (Y= H, F, CH3, Cl) and PXYZ (Y= H, Cl and Z= F) with X= H, CH3, NH2, OH, F, SiH3, PH2, SH, Cl.  相似文献   

8.
The MP2 ab initio quantum chemistry methods were utilized to study the halogen‐bond and pnicogen‐bond system formed between PH2X (X = Br, CH3, OH, CN, NO2, CF3) and BrY (Y = Br, Cl, F). Calculated results show that all substituent can form halogen‐bond complexes while part substituent can form pnicogen‐bond complexes. Traditional, chlorine‐shared and ion‐pair halogen‐bonds complexes have been found with the different substituent X and Y. The halogen‐bonds are stronger than the related pnicogen‐bonds. For halogen‐bonds, strongly electronegative substituents which are connected to the Lewis acid can strengthen the bonds and significantly influenced the structures and properties of the compounds. In contrast, the substituents which connected to the Lewis bases can produce opposite effects. The interaction energies of halogen‐bonds are 2.56 to 32.06 kcal·mol?1; The strongest halogen‐bond was found in the complex of PH2OH???BrF. The interaction energies of pnicogen‐bonds are in the range 1.20 to 2.28 kcal·mol?1; the strongest pnicogen‐bond was found in PH2Br???Br2 complex. The charge transfer of lp(P) ? σ*(Br? Y), lp(F) ? σ*(Br? P), and lp(Br) ? σ*(X? P) play important roles in the formation of the halogen‐bonds and pnicogen‐bonds, which lead to polarization of the monomers. The polarization caused by the halogen‐bond is more obvious than that by the pnicogen‐bond, resulting in that some halogen‐bonds having little covalent character. The symmetry adapted perturbation theory (SAPT) energy decomposition analysis showes that the halogen‐bond and pnicogen‐bond interactions are predominantly electrostatic and dispersion, respectively.  相似文献   

9.
A density functional theory (DFT) study of carbon? hydrogen versus carbon? heteroatom bond activation is presented. Heteroatom groups (X) investigated include X = F, Cl, OH, SH, NH2, PH2. The activating model complex is a prototypical d0 zirconium‐imide. While C? X activation has a thermodynamic advantage over C? H activation, the former has been found to have a kinetic advantage. Implications for catalytic hydrocarbon functionalization and phosphine–ligand degradation are discussed. The present results for a high‐valent metal complex are compared/contrasted with low‐valent bond activating complexes. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  相似文献   

10.
The structure and electronic parameters of ClZ(CH3)2X molecules (Z = C, Si, Ge, X = CH3, OCH3) were calculated by the RHF/6–31G(d) and RHF/6–311G(d,p) methods with full geometry optimization; calculations of ClZ(CH3)2OCH3 molecules were also performed by the RHF/6–31G(d) method with partial geometry optimization. The 35Cl NQR frequencies calculated from the populations of less diffuse 3p constituents of valence p orbitals of chlorine [RHF/6–31G(d)] were in agreement with the experimental values. The 35Cl NQR frequencies for molecules with X = OCH3 are lower than those for molecules with X = CH3 (the Z atom being the same), due mainly to direct through-field polarization of the Z-Cl bond, induced by the effect of unshared electron pair of the oxygen atom in the trans position with respect to that bond. The difference in the 35Cl NQR frequencies decreases in going from Z = C to Z = Si, Ge, in parallel with variation of the Z-Cl bond polarization as the size of Z increases.  相似文献   

11.
《Polyhedron》1988,7(6):449-462
The complexes [ML*(NO)Cl(OR)] {L* = HB(3,5-Me2C3HN2)3; M= Mo, R = CH2CH2X, X = Cl, OMe or OEt; (CH2)nOH, n = 2, 5, 6; M = W, R = CH2CH2X, X = Cl, OMe or OEt; (CH2)nOH, n = 2–6; CH2(CF2)3CH2OH; CHMeCH2CMe2OH} and [ML*(NO)(OR)2] {M = Mo, R = CH2CH2X, X = Cl, OMe or OEt; (CH2)nOH, n = 2–6; M = W,R = CH2CH2X, X= Cl, OMe or OEt; (CH2)nOH, n = 2,4–6; CH2(CF2)3CH2OH} have been prepared from [ML*(NO)Cl2] and the appropriate alcohol in the presence of NEt3 or NaCO3, and have been characterized by IR, 1H NMR and mass spectroscopy.  相似文献   

12.
We have quantum chemically studied the iron-mediated C X bond activation (X = H, Cl, CH3) by d8-FeL4 complexes using relativistic density functional theory at ZORA-OPBE/TZ2P. We find that by either modulating the electronic effects of a generic iron-catalyst by a set of ligands, that is, CO, BF, PH3, BN(CH3)2, or by manipulating structural effects through the introduction of bidentate ligands, that is, PH2(CH2)nPH2 with n = 6–1, one can significantly decrease the reaction barrier for the C X bond activation. The combination of both tuning handles causes a decrease of the C H activation barrier from 10.4 to 4.6 kcal mol−1. Our activation strain and Kohn-Sham molecular orbital analyses reveal that the electronic tuning works via optimizing the catalyst–substrate interaction by introducing a strong second backdonation interaction (i.e., “ligand-assisted” interaction), while the mechanism for structural tuning is mainly caused by the reduction of the required activation strain because of the pre-distortion of the catalyst. In all, we present design principles for iron-based catalysts that mimic the favorable behavior of their well-known palladium analogs in the bond-activation step of cross-coupling reactions.  相似文献   

13.
The distonic radical cation C5H5N+?·CH2 can be generated by the reactions of neutral pyridine with the radical cations of cyclopropane, ethylene oxide, and ketene, as well as with the [C3H6]+ ion from fragmentation of tetrahydrofuran. The distonic product ion can be distinguished from isomeric methylpyridine radical cations because the former gives characteristic [M?CH2]+, [M ? CH2NCH]+, and a doubly charged ion, all of which are produced on collisional activation. Furthermore, the distonic species completely transfers CH2 + to more nucleophilic, substituted pyridines. These properties are all consistent with the assigned distonic structure. Another distonic isomer, the (3-methylene) pyridinium ion, can be distinguished from the (1-methylene)pyridinium ion on the basis of their different fragmentation behaviors. The latter ion exhibits higher stability (lower reactivity) than the prototypal [·CH2NH3 +], making available a distonic species whose bimolecular reactivity can be readily investigated.  相似文献   

14.
It has been established that transformations of azetidine radical cations observed in freonic matrices under the action of light with λ = 436 nm (T = 77 K) are associated with C-N bond cleavage which corresponds to the cyclic form yielding a mixture of open distonic C-centered radical cations of the following structure: ·CH2CH2CH=NH 2 +   相似文献   

15.
According to the X-ray diffraction data, the crystal and molecular structure of tris(2-hydroxyethyl) ammonium fluoride (F?N+H(CH2CH2OH)3, fluoroprotatrane, substantially differs from other halo protatranes X?N+H(CH2CH2OH)3 (X = Cl, Br, and I). At X = F, to the endo-molecular LP of the nitrogen atom the HF molecule having the minimum ionic radius in a series of X? anions is bonded. The geometry of fluoroprotatrane and the cation packing in the crystal are analyzed.  相似文献   

16.
《Polyhedron》2002,21(5-6):579-585
Substituent effects on the potential energy surface of XGeAs (X=H, Li, Na, BeH, MgH, BH2, AlH2, CH3, SiH3, NH2, PH2, OH, SH, F, and Cl) were investigated by using B3LYP and CCSD(T) methods. The isomers include structures with formal double (GeAsX) and triple (XGeAs) bonds to germanium–arsenic, so a direct comparison of these types of species is possible. Our model calculations indicate that electropositively substituted GeAsX species are thermodynamically and kinetically more stable than their isomeric XGeAs molecules. Moreover, the theoretical findings suggest that F, OH, NH2, and CH3 substitution prefer to shift the double bond (GeAsX) by forming a triple bond (XGeAs).  相似文献   

17.
The reactions indicated in the title have been studied in terms of direct processes and complex formation. Quantum-chemical methods have been applied to the passage of an acid (H+, CH, X+) from CH3X to CH3X, and the abstraction of a radical (H· CH, X·) from CH3X by CH3X. It has been shown that a complex represented by a dimer of a methyl-halide radical cation, (CH3X), with a two-center three-electron bond X? X, has fairly high stability. These investigations were based on non-empirical quantum-chemical calculations, the results being systematically compared with experimental determinations. Some calculations included all electrons (X=F, Cl, Br), others were based on relativistic pseudopotentials (X=F through At). The two sets of calculations agree qualitatively with each other and with experimental observations.  相似文献   

18.
Cationic alkoxycarbene complexes of platinum(II) have been isolated in the reactions of trans-[(PR3)2PtX(R′OH)]PF6 (X  H or Me; R′  Me or Et) with Me3SiCCR′′ (R′′  H, Me or SiMe3). In these reactions cleavage of the carbon-silicon bond by the nucleophilic attack of alcohol has been observed. These carbene complexes have been characterized by elemental analyses and by IR, 1H and 13C NMR spectral data. 13C NMR chemical shift data for carbene carbon atoms suggest that the carbene carbon may be very positively charged.  相似文献   

19.
Various p-substituted benzyl p-hydroxyphenyl methyl sulfonium salts ( 2 ) were synthesized and their initiator activities were evaluated in bulk polymerization of glycidyl phenyl ether (PGE). The order of the activity was found to be 2b (X = CH3) > 2a (X = H) ≈ 2c (X = Cl) > 2d (X = NO2), indicating that the introduction of an electron-donating group enhanced the activity. In Hammett's plots, the logarithm of the ratio of the polymerization rates (log kx/kH) was correlated with σ+ρ better than with σp and a negative ρ+ value (-1.18) was obtained. Reaction of 2a with benzyl mercaptan mainly gave dibenzyl sulfide and p-hydroxyphenyl methyl sulfide. The obtained results seemed to demonstrate that the OH group of the aryl group yielded no proton as initiator for the polymerization, whereas the benzyl group caused the polymerization, which was initiated by the corresponding benzyl cation formed by C? S bond cleavage. © 1993 John Wiley & Sons, Inc.  相似文献   

20.
The unimolecular chemistry of the methyl carbamate radical cation, H2NCOOCH, 1, has been further investigated by a combination of mass spectrometry-based experiments (metastable ion (MI), collisional activation (CA), collision-induced dissociative ionization (CIDI), neutralization-reionization (NR) Spectrometry and 2H labelling) and ab initio molecular orbital calculations, executed at the MP3/6–31G*//4–31G level of theory and corrected for zero-point vibrational energies. Apart from the previously located maxima, i.e. H2NCOOCH3, 1, the distonic ion H2NC(OH)OCH3, 2, hydrogen-bridged ions [H2N? C?O…? H…?O?CH2], 5, and [H2N? CH?O…?…?H…?O?C? H], 7, there exist at least two other equilibrium structures, viz. the iminol ion H? N?C(OH)? OCH, la, and the hydrogen-bridged species [H2C?O…?H…?N(H)COH], 6a, which is closely related to ion 5. Although the iminol ion la lies only 30 kJ mol?1 above 1, our calculations indicate that the barriers for its formation either directly from ionized methyl carbamate 1 via a 1,3-hydrogen shift or indirectly via 1,4-hydrogen shifts from the distonic ion 2 are too high to allow the iminol ion to be involved in the unimolecular chemistry of ionized methyl carbamate. This explains the earlier observation that there are no H-D exchange reactions prior to decomposition of ionized labelled methyl carbamate, in contrast to the related ion methyl acetate. However, attempts to generate the iminol ion by loss of CH3CN from CH3CH?N? NHCOOCH3 produced the more stable distonic ion 2 instead, but it proved very difficult to assign its structure unequivocally because 2 can rapidly interconvert with 1 and so virtually identical dissociation characteristics ensue. Only by integration of results obtained from many experiments and from ab initio calculations could structure 2 be assigned. The distonic ion 2 can undergo two transformations: after stretching of the C? OCH2 bond the incipient formaldehyde can migrate within the electrostatic field of ionized hydroxyaminocarbene to the OH end to generate 5, but it can also migrate to the NH end to generate 6a. This explains the previous puzzling observation that H2NCOOCD forms both CD2OD· and CD2OH· in CA and NR experiments. The calculations and experiments indicate that, although the ion is exceedingly difficult to characterize, the distonic ion 2 is the key intermediate for all the observed dissociations of methyl carbamate.  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