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.
Two series of neopentylbenzenes with one or two substituents on the benzyl group have been synthesized. In one series the substituents were H, F, Cl, Br, I, OCH3, OCOCH3, OSi(CH3)3 CH3 and CH2CH3, and in the other OH and R [R ? H, CH3, CH2CH3, (CH2)3CH3, CH(CH3)2 and C(CH3)3]. Barriers to internal C? C and C? C rotation have been estimated by 13C NMR band shape methods. Estimated barriers were found to increase as the size of the substituent increases. The results are discussed in terms of possible initial and transition states, based on summations of results from molecular mechanics (MM) calculations, using the Allinger MMP1 program. Barriers estimated experimentally are compared with results from other systems found in the literature.  相似文献   

2.
The effects of several substituents (? BH2, ? BF2, ? AlH2, ? CH3, ? C6H5, ? CN, ? COCH3, ? CF3, ? SiH3, ? NH2, ? NH3+, ? NO2, ? PH2, ? OH, ? OH2+, ? SH, ? F, ? Cl, ? Br) on the Bergman cyclization of (Z)‐1,5‐hexadiyne‐3‐ene (enediyne, 3 ) were investigated at the Becke–Lee–Yang–Parr (BLYP) density functional (DFT) level employing a 6‐31G* basis set. Some of the substituents (? NH3+, ? NO2, ? OH, ? OH2+, ? F, ? Cl, ? Br) are able to lower the barrier (up to a minimum of 16.9 kcal mol?1 for difluoro‐enediyne 7rr ) and the reaction enthalpy (the cyclization is predicted to be exergonic for ? OH2+ and ? F) compared to the parent system giving rise to substituted 1,4‐dehydrobenzenes at physiological temperatures. © 2001 John Wiley & Sons, Inc. J Comput Chem 22: 1605–1614, 2001  相似文献   

3.
The effects of substituents (X) on the structures and stabilities of CH2X? anions for groups comprised of fourth- and fifth-period main group elements (X = K, CaH, GaH2, GeH3, AsH2, SeH, Br, Rb, SrH, InH2, SnH3, SbH2, TeH, and I) have been investigated by ab initio pseudopotential calculations. Full geometry optimizations have been carried out on the CH2X? anions and the corresponding neutral parent molecules, CH3X, at HF/DZP + and MP2/DZP + levels. Results for substituents from the second (X = Li? F) and third (X = Na? Cl) periods provide comparisons of substituent effects of the main group elements of the first four rows of the periodic table on methyl anions. Frequency calculations characterize the nature of stationary points and show pyramidal CH2X? anion structures to be the most stable unless π acceptor interactions (e.g., with BH2, AlH2, GaH2, and InH2 favor planar geometries. The CH2X? stabilization energies [at QCISD(T)/DZP + /MP2/DZP + + ZPE level for X = K? I and QCISD(T)/6?31 + G*/MP2/6?31 + G* + ZPE level] for X = Li? Cl) also show strong π-stabilizing effects for the same substituents. With the exception of CH3 and NH2, all substituents stabilize methyl anions, although the σ stabilization by OH and F is small. The SiH3? PH2? SH? Cl, GeH3? AsH2? SeH? Br, and SnH3? SbH2? TeH? I sets of substituents give stabilization energies between 19 and 30 kcal/mol. The stability of methyl anions substituted by the halogens and the chalcogens (X = OH, SH, SeH, and TeH) increases down a group in accord with the increasing substituent polarizability, while for π acceptors (BH2, AlH2, GaH2, and InH2) the stability decreases down a group in line with their π-accepting ability. © 1994 by John Wiley & Sons, Inc.  相似文献   

4.
CCSD(T) calculations have been used for identically nucleophilic substitution reactions on N‐haloammonium cation, X? + NH3X+ (X = F, Cl, Br, and I), with comparison of classic anionic SN2 reactions, X? + CH3X. The described SN2 reactions are characterized to a double curve potential, and separated charged reactants proceed to form transition state through a stronger complexation and a charge neutralization process. For title reactions X? + NH3X+, charge distributions, geometries, energy barriers, and their correlations have been investigated. Central barriers ΔE for X? + NH3X+ are found to be lower and lie within a relatively narrow range, decreasing in the following order: Cl (21.1 kJ/mol) > F (19.7 kJ/mol) > Br (10.9 kJ/mol) > I (9.1 kJ/mol). The overall barriers ΔE relative to the reactants are negative for all halogens: ?626.0 kJ/mol (F), ?494.1 kJ/mol (Cl), ?484.9 kJ/mol (Br), and ?458.5 kJ/mol (I). Stability energies of the ion–ion complexes ΔEcomp decrease in the order F (645.6 kJ/mol) > Cl (515.2 kJ/mol) > Br (495.8 kJ/mol) > I (467.6 kJ/mol), and are found to correlate well with halogen Mulliken electronegativities (R2 = 0.972) and proton affinity of halogen anions X? (R2 = 0.996). Based on polarizable continuum model, solvent effects have investigated, which indicates solvents, especially polar and protic solvents lower the complexation energy dramatically, due to dually solvated reactant ions, and even character of double well potential in reactions X? + CH3X has disappeared. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

5.
The geometries and interaction energies of complexes of pyridine with C6F5X, C6H5X (X=I, Br, Cl, F and H) and RFI (RF=CF3, C2F5 and C3F7) have been studied by ab initio molecular orbital calculations. The CCSD(T) interaction energies (Eint) for the C6F5X–pyridine (X=I, Br, Cl, F and H) complexes at the basis set limit were estimated to be ?5.59, ?4.06, ?2.78, ?0.19 and ?4.37 kcal mol?1, respectively, whereas the Eint values for the C6H5X–pyridine (X=I, Br, Cl and H) complexes were estimated to be ?3.27, ?2.17, ?1.23 and ?1.78 kcal mol?1, respectively. Electrostatic interactions are the cause of the halogen dependence of the interaction energies and the enhancement of the attraction by the fluorine atoms in C6F5X. The values of Eint estimated for the RFI–pyridine (RF=CF3, C2F5 and C3F7) complexes (?5.14, ?5.38 and ?5.44 kcal mol?1, respectively) are close to that for the C6F5I–pyridine complex. Electrostatic interactions are the major source of the attraction in the strong halogen bond although induction and dispersion interactions also contribute to the attraction. Short‐range (charge‐transfer) interactions do not contribute significantly to the attraction. The magnitude of the directionality of the halogen bond correlates with the magnitude of the attraction. Electrostatic interactions are mainly responsible for the directionality of the halogen bond. The directionality of halogen bonds involving iodine and bromine is high, whereas that of chlorine is low and that of fluorine is negligible. The directionality of the halogen bonds in the C6F5I– and C2F5I–pyridine complexes is higher than that in the hydrogen bonds in the water dimer and water–formaldehyde complex. The calculations suggest that the C? I and C? Br halogen bonds play an important role in controlling the structures of molecular assemblies, that the C? Cl bonds play a less important role and that C? F bonds have a negligible impact.  相似文献   

6.
The following p-substituted N,N-bis-trimethlsilyl anilines p-X? C6H4? N[Si(CH3)3]2 are prepared by silylation of free amines: X = H, CH3, C2H5, CH3O, CH3CO, F, Cl, Br, J, CN, C6HS, (CH3)3SiO, and [(CH3)3Si]2N, and the isotopic derivatives C6H5? 15N[Si(CH3)3]2 and C6D5N[Si(CH3)3]2. The vibrational spectra are reported and assigned. The molecular symmetry of p-[(CH3)3Si]2N? C6H4? N[Si(CH3)3]2 is determined. The influence of the mass of the substituents X on the positions of the νsSiNSi vibrational frequencies is discussed.  相似文献   

7.
The substituent effect of electron‐withdrawing groups on electron affinity and gas‐phase basicity has been investigated for substituted propynl radicals and their corresponding anions. It is shown that when a hydrogen of the α‐CH3 group in the propynyl system is substituted by an electron‐withdrawing substituent, electron affinity increases, whereas gas‐phase basicity decreases. These results can be explained in terms of the natural atomic charge of the terminal acetylene carbon of the systems. The calculated electron affinities are 3.28 eV (?C?C? CH2F), 3.59 eV (?C?C? CH2Cl) and 3.73 eV (?C?C? CH2Br), and the gas‐phase basicities of their anions are 359.5 kcal/mol (?:C?C? CH2F), 354.8 kcal/mol (:C?C? CH2Cl) and 351.3 kcal/mol (?:C?C? CH2Br). It is concluded that the larger the magnitude of electron‐withdrawing, the greater is the electron affinity of radical and the smaller is the gas‐phase basicity of its anion. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009  相似文献   

8.
Halomethylation of polysulfone (PS) with C8H17OCH2X (X = Cl, Br) in the presence of SnX4 (X = Cl, Br) led to PS–CH2X (X = Cl or Br or both) (Scheme 1). Under controlled conditions, PS–CH2X could be isolated and retains the good film forming properties of PS itself. Interhalogen exchange reactions occur in the presence of SnX4 (X = Cl, Br) under anhydrous conditions (Scheme 1), or a quaternary ammonium phase transfer catalyst R*R3N+X?, under aqueous conditions (Scheme 2). The exchange reactions with R*R3N+X?, are favored when R = C8? C10, and with R = C4 only if n-octanol is added; otherwise gelation occurs. Exchange in CHCl3 is attributed to dehydrohalogenation (and generation of dichlorocarbene) of the solvent in the presence of tetrabutyl ammonium hydroxide. Further chemical modifications of PS–CH2X by reaction with strong nucleophiles, led to hydroxymethyl polysulfone, acetoxymethyl polysulfone, and t-butyl-oxymethyl polysulfone (Scheme 3). Hydroxymethyl polysulfone sometimes gels under basic hydrolytic conditions and is best obtained by methanolysis of PS–CH2-OAc. Both PS? CH2? OAc and PS? CH2O-t-Bu are very stable, and provide a way to generate PS? CH2Br on need by cleavage with HBr in acetic acid. Direct oxidations with DMSO or tetrabutyl ammonium dichromate (Scheme 4) or indirect oxidations (Scheme 5) produce polysulfone with pendent CHO, CO2R and PO3R groups. Finally, polysulfones with linker arms including, carboxy alkyl, hexaglycol or sulfonamido crowns are described (Scheme 6). The reaction products were characterized by 1H- and 13C-NMR. Double irradiation experiments, proved unequivocally, that the first substitution occurred on the B ring of the bisphenol A moiety (see Table I); the second substitution occurs on the A ring in position a. Thermogravimetric analysis generally shows for all modified polysulfones an extra transition at a lower temperature. The area of this band agrees generally with the values expected from calculated substitution degrees.  相似文献   

9.
Why is silicon hypervalent and carbon not? Or why is [Cl? CH3? Cl]? labile with a tendency to localize one of its axial C? Cl bonds and to largely break the other one, while the isostructural and isoelectronic [Cl? SiH3? Cl]? forms a stable pentavalent species with a delocalized structure featuring two equivalent Si? Cl bonds? Various hypotheses have been developed over the years focusing on electronic and steric factors. Here, we present the so‐called ball‐in‐a‐box model, which tackles hypervalence from a new perspective. This model reveals the key role of steric factors and provides a simple way of understanding the above phenomena in terms of different atom sizes. Our bonding analyses are supported by computation experiments in which we probe, among other things, the shape of the SN2 potential‐energy surface of Cl? attacking a carbon atom in the series of substrates CH3Cl, .CH2Cl, ..CHCl, and ...CCl. Our findings for ClCH3Cl? and ClSiH3Cl? are generalized to other Group 14 central atoms (Ge, Sn, and Pb) and axial substituents (F).  相似文献   

10.
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   

11.
Gas‐phase anionic reactions X? + CH3SY (X, Y = F, Cl, Br, I) have been investigated at the level of B3LYP/6‐311+G (2df,p). Results show that the potential energy surface (PES) of gas‐phase reactions X? + CH3SY (X, Y = Cl, Br, I) has a quadruple‐well structure, indicating an addition–elimination (A–E) pathway. The fluorine behaves differently in many respects from the other halogens and the reactions F? + CH3SY (Y = F, Cl, Br, I) correspond to deprotonation instead of substitution. The gas‐phase reactions X? + CH3SF (X = Cl, Br, I), however, follow an A–E pathway other than the last two out going steps (COM2 and PR) that proceeds via a deprotonation. The polarizable continuum model (PCM) has been used to evaluate the solvent effects on the energetics of the reactions X? + CH3SY (X, Y = Cl, Br, I). The PES is predicted to be unimodal in the solvents of high polarity. © 2007 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  相似文献   

12.
The R? CH2? HO…H? X (R = SCl, Cl, SH, NO2, OMe, CHO, CN, C2H5, CH3, H; X = F, Cl, Br) complexes are considered here as the interest sample for the consideration of different measures of H‐bond strength. The intermolecular interaction energies are predicted by using MP2/6‐31++G(d,p) and B3LYP/6‐31++G(d,p) methods with basis set superposition error and zero‐point energy corrections. The results showed that intermolecular hydrogen bonds for complexes with HF are stronger than such interactions in complexes with HCl and HBr. Quantum theory of “Atoms in Molecules” and natural bond orbitals method were applied to analyzed H‐bond interactions. The gas phase thermodynamic properties of complexes were predicted using quantum mechanical computations. The obtained results showed a strong influence of the R and X substituents on the thermodynamic properties of complexes. Numerous correlations between topological, geometrical, thermodynamic properties and energetic parameters were also found. © 2011 Wiley Periodicals, Inc.  相似文献   

13.
The gas‐phase reactions of XH? (X=O, S) + CH3Y (Y=F, Cl, Br) span nearly the whole range of SN2 pathways, and show an intrinsic reaction coordinate (IRC) (minimum energy path) with a deep well owing to the CH3XH???Y? (or CH3S????HF) hydrogen‐bonded postreaction complex. MP2 quasiclassical‐type direct dynamics starting at the [HX???CH3???Y]? transition‐state (TS) structure reveal distinct mechanistic behaviors. Trajectories that yield the separated CH3XH+Y? (or CH3S?+HF) products directly are non‐IRC, whereas those that sample the CH3XH???Y? (or CH3S????HF) complex are IRC. The IRCIRC/non‐IRC ratios of 90:10, 40:60, 25:75, 2:98, 0:100, and 0:100 are obtained for (X, Y)=(S, F), (O, F), (S, Cl), (S, Br), (O, Cl), and (O, Br), respectively. The properties of the energy profiles after the TS cannot provide a rationalization of these results. Analysis of the energy flow in dynamics shows that the trajectories cross a dynamical bifurcation, and that the inability to follow the minimum energy path arises from long vibration periods of the X?C???Y bending mode. The partition of the available energy to the products into vibrational, rotational, and translational energies reveals that if the vibrational contribution is more than 80 %, non‐IRC behavior dominates, unless the relative fraction of the rotational and translational components is similar, in which case a richer dynamical mechanism is shown, with an IRC/non‐IRC ratio that correlates to this relative fraction.  相似文献   

14.
Silicon in [Cl? SiH3? Cl]? is hypervalent, whereas carbon in [Cl? CH3? Cl]? is not. We have recently shown how this can be understood in terms of the ball‐in‐a‐box model, according to which silicon fits perfectly into the box that is constituted by the five substituents, whereas carbon is too small and, in a sense, “drops to the bottom” of the box. But how does carbon acquire hypervalency in the isostructural and isoelectronic noble gas (Ng)/methyl cation complexes [Ng? CH3? Ng]+ (Ng=He and Ne), which feature a delocalized D3h‐symmetric structure with two equivalent C? Ng bonds? From Ng=Ar onwards, the [Ng? CH3? Ng]+ complex again acquires a propensity to localize one of its axial C? Ng bonds and to largely break the other one, and this propensity increases in the order Ng=Ar3Ng+ and, for comparison, CH3Ng+, NgHNg+, and NgH+. It appears that, at variance with [Cl? CH3? Cl]?, the carbon atom in [Ng? CH3? Ng]+ can no longer be considered as a ball in a box of the five substituents.  相似文献   

15.
The accelerated formation of 2,3-diphenylquinoxalines in microdroplets generated in a nebulizer has been investigated by competition experiments in which equimolar quantities of 1,2-phenylenediamine, C6H4(NH2)2, and a 4-substituted homologue, XC6H3(NH2)2 [X = F, Cl, Br, CH3, CH3O, CO2CH3, CF3, CN or NO2], or a 4,5-disubstituted homologue, X2C6H2(NH2)2 [X = F, Cl, Br, or CH3], compete to condense with benzil, (C6H5CO)2. Electron-donating substituents (X = CH3 and CH3O) accelerate the reaction; in contrast, electron-attracting substituents (X = F, Cl, Br and particularly CO2CH3, CN, CF3 and NO2) retard it. A structure–reactivity relationship in the form of a Hammett correlation has been found by analyzing the ratio of 2,3-diphenylquinoxaline and the corresponding substituted-2,3-diphenylquinoxaline, giving a ρ value of −0.96, thus confirming that the electron density in the aromatic ring of the phenylenediamine component is reduced in the rate-limiting step in this accelerated condensation. This correlation shows that the phenylenediamine acts as a nucleophile in the reaction.  相似文献   

16.
The reaction mechanism of the halogen (Cl and Br)-atom initiated oxidation of C2H4 was studied using the long path FTIR spectroscopic method in 700 torr of air at 296 ± 2 K. Among the major halogen-containing products were X? CH2CHO, X? CH2CH2OH, and X? CH2CH2OOH (X = Cl or Br) which were shown to be formed via the self-reaction of the X? CH2CH2OO radicals, i.e., 2X? CH2CH2OO → 2X? CH2CH2O + O2; (a) 2X? CH2CH2OO → X? CH2CHO + X? CH2CH2OH + O2 and (b) followed by X? CH2CH2O + O2 → X? CH2CHO + HO2 and X? CH2CH2OO + HO2 → X? CH2CH2OOH + O2. From the observed yields of X? CH2CHO and X? CH2CH2OH the branching ratios for reactions (a) and (b) were determined to be ka/kb = 1.35 ± 0.07(2σ) for both X = Cl and Br. In addition, the O2-dependence of the rate constant for the Br + C2H4 reaction was determined by the relative rate technique as a function of O2 partial pressure from 140 to 700 torr at 700 torr total pressure of N2/O2 diluent. Rate constants for the reactions of Cl-atoms with Cl-CH2CHO and Br-atoms with Br-CH2CHO were also determined to be [4.3 ± 0.2(2sigma;)] × 10?11 and less than or equal to [1.83 ± 0.11(2σ)] × 10?13 cm3 molecule?1 s?1, respectively.  相似文献   

17.
Total NMR band shape fitting methods have provided accurate energy data for inversion barriers at sulphur and selenium in complexes of types cis-[MX2L] (M = PdII, PtII; X = Cl, Br, I; L = MeS(CH2)2SMe, MeS(CH2)3SMe, o-(SMe)2C6H3Me, cis-MeSCH=CHSMe) and [PtXMe{MeE(CH2)2E′Me}] (E= E′= S or Se and E = S, E′= Se; X = Cl, Br, I). Barrier energies were found to decrease by 10–12 kJ mol?1 in going from aliphatic through aromatic to olefinic ligand back-bone. This can be explained in terms of (3p - 2p) π conjugation between the inverting centre and the ligand back-bone. The effects of ligand ring size, nature of halogen atom and the metal oxidation state on the barrier energies are discussed.  相似文献   

18.
The dielectric relaxation rates of 18 phenols and 3 thiophenols have been determined in paraffin solutions at 4.2 or 77 K and used to obtain relative values of the tunnel splitting of the ground torsional state. Where comparison is possible the results are compatible with vapour phase microwave spectroscopic data. The barriers to hydroxyl rotation are calculated and are fairly consistent with barriers derived from far infrared data. Para-substituents F, CH3, Cl, C(CH3)3, Br and I in order of decreasing effectiveness, lower the barrier. COOH and CHO raise it by 350–400 cm?1.  相似文献   

19.
Positron-annihilation lifetime spectra have been measured for mixtures of CH3Cl and CH3Br in cyclohexane and of CH3Cl in benzene. The ortho-positronium (Ps) yield decreased monotonically from 38% and 43% in cyclohexane and benzene respectively to 11% in pure CH3Cl and 6% in pure CH3Br. The strength of the inhibition of Ps formation by CH3Br was ten times that of CH3Cl in cyclohexane, because the CH3Br? anion debrominates rapidly, while CH3Cl? is long-lived (= 30 ns) compared to the maximum time of Ps formation of 400–500 ps. as shown in radiation chemistry. The positron can pick off the electron from the CH3X? anions to form Ps. while it forms a bound state with the halides. X?. CH3Cl was a roughly three times weaker Ps inhibitor in benzene than in cyclohexane, which shows that CH3Cl? does not dechlorinate in times comparable to or shorter than 400–500 ps in benzene. An improved model for the explanation of Ps formation in mixtures, where the Ps yield versus electron scavenger concentration has a minimum, is proposed and discussed.  相似文献   

20.
The thermal functions S0T, -(G0T-H0O)/T and (H0T-H0O) have been calculated from structural and spectroscopic data for the gaseous organometallics C5H5BeX (X = Cl, Br and BH4), C5H5MX3 (M = Ti and Ge; X = Cl, Br and I) and CH3TiX3 (X = Cl, Br and I). The rotational barriers of the C5H5 and CH3 groups have been evaluated and discussed.  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