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 0 毫秒
1.
This paper examines the unimolecular dissociation of propargyl (HCCCH2) radicals over a range of internal energies to probe the CH+HCCH and C+C2H3 bimolecular reactions from the radical intermediate to products. The propargyl radical was produced by 157 nm photolysis of propargyl chloride in crossed laser-molecular beam scattering experiments. The H-loss and H2 elimination channels of the nascent propargyl radicals were observed. Detection of stable propargyl radicals gave an experimental determination of 71.5 (+5-10) kcal/mol as the lowest barrier to dissociation of the radical. This barrier is significantly lower than predictions for the lowest barrier to the radical's dissociation and also lower than calculated overall reaction enthalpies. Products from both H2+HCCC and H+C3H2 channels were detected at energies lower than what has been theoretically predicted. An HCl elimination channel and a minor C-H fission channel were also observed in the photolysis of propargyl chloride.  相似文献   

2.
The reactions of C2(a3Piu) radicals with a series of alkanes have been studied at room temperature and 6.5 torr total pressure using the pulsed laser photolysis/laser-induced fluorescence technique. C2(a3Piu) radicals were generated by photolysis of C2Cl4 with the focused output from the fourth harmonic of a Nd: YAG laser at 266 nm. The relative concentration of C2(a3Piu) radicals was monitored on the (0,0) band of the C2(d3Pig <-- a3Piu) transition at 516.5 nm by laser-induced fluorescence. From the analysis of the relative concentration-time behavior of C2(a3Piu) under pseudofirst-order conditions, the rate constants for the reactions of C2(a3Piu) with alkanes (C1-C8) were determined. The rate constant increases linearly with the increasing of the number of CH2 groups in the alkanes. The experimental results indicate that the reaction of C2(a3Piu) with small alkanes (C1-C8) follows the typical hydrogen abstraction process. Based on the correlation of the experimental results with the bond dissociation energy of the alkanes, the reactions of C2(a3Piu) with small alkanes likely proceed via the mechanism of hydrogen abstraction.  相似文献   

3.
The absolute bimolecular rate constants for the reactions of C6H5 with 2‐methylpropane, 2,3‐dimethylbutane and 2,3,4‐trimethylpentane have been measured by cavity ringdown spectrometry at temperatures between 290 and 500 K. For 2‐methylpropane, additional measurements were performed with the pulsed laser photolysis/mass spectrometry, extending the temperature range to 972 K. The reactions were found to be dominated by the abstraction of a tertiary C H bond from the molecular reactant, resulting in the production of a tertiary alkyl radical: C6H5 + CH(CH3)3 → C6H6 + t‐C4H9 (1) (1) C6H5 + (CH3)2CHCH(CH3)2 → C6H6 + t‐C6H13 (2) (2) C6H5 + (CH3)2CHCH(CH3)CH(CH3)2 → C6H6 + t‐C8H17 (3) (3) with the following rate constants given in units of cm3 mol−1 s−1: k1 = 10(11.45 ± 0.18) e−(1512 ± 44)/T k2 = 10(11.72 ± 0.15) e−(1007 ± 124)/T k3 = 10(11.83 ± 0.13) e−(428 ± 108)/T © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 645–653, 1999  相似文献   

4.
The potential energy surface for the reaction of the ground‐state carbon atom [C(3Pj)] with the propargyl radical [HCCCH2(X2B1)] is investigated using the G2M(RCC,MP2) method. Numerous local minima and transition states for various isomerization and dissociation pathways of doublet C4H3 are studied. The results show that C(3Pj) attacks the π system of the propargyl radical at the acetylenic carbon atom and yields the n‐C4H3(2A′) isomer i3 after an 1,2‐H atom shift. This intermediate either splits a hydrogen atom and produces singlet diacetylene, [HCCCCH ( p1 )+H] or undergoes (to a minor amount) a 1,2‐H migration to i‐C4H3(2A′) i5 , which in turn dissociates to p1 plus an H atom. Alternatively, atomic carbon adds to the triple C?C bond of the propargyl radical to form a three‐member ring C4H3 isomer i1 , which ring opens to i3 . Diacetylene is concluded to be a nearly exclusive product of the C(3Pj)+HCCCH2 reaction. At the internal energy of 10.0 kcal/mol above the reactant level, Rice–Ramsperger–Kassel–Marcus calculations show about 91.7% of HCCCCH comes from fragmentation of i3 and 8.3% from i5 . The other possible minor channels are identified as HCCCC+H2 and C2H+HCCH. © 2001 John Wiley & Sons, Inc. J Comput Chem 22: 1522–1535, 2001  相似文献   

5.
Copper‐catalyzed Sonogashira‐type reactions were dramatically accelerated by introducing a catalytic amount of polycyclic aromatic hydrocarbon additive. This novel catalytic system features low copper loading (0.5 mol% < Cu < 5 mol%), broad reaction scope and remarkable substrate tolerance. Both aromatic and aliphatic terminal alkynes as well as diverse aryl iodides were employed in this transformation, affording respectable yields of the desired products. The novel Cu(OTf)2/pyrene system was subsequently employed to synthesize phenylacetylene‐based fluorescent compounds. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

6.
The kinetics of the gas-phase reactions of the OH radical with (C2H5O)3PO and (CH3O)2P(S)Cl and of the reactions of NO3 radicals and O3 with (CH3O)2P(S)Cl have been studied at room temperature. Using a relative rate technique, the rate constants determined for the reactions of the OH radical with (C2H5O)3PO and (CH3O)2P(S)Cl at 296 ± 2 K and 740 torr total pressure of air were (5.53 ± 0.35) × 10?11 and (5.96 ± 0.38) × 10?11 cm3 molecule?1 s?1, respectively. Upper limits to the rate constants for the NO3 radical and O3 reactions with (CH3O)2P(S)Cl of <3 × 10?14 cm3 molecule?1 s?1 and <2 × 10?19 cm3 molecule?1 s?1, respectively, were obtained. These data are compared and discussed with previous literature data for organophosphorus compounds.  相似文献   

7.
The kinetics of C6H5 reactions with C2H6 (1) and neo‐C5H12 (2) have been studied by the pulsed laser photolysis/mass spectrometric method using C6H5COCH3 as the phenyl precursor at temperatures between 565 and 1000 K. The rate constants were determined by kinetic modeling of the absolute yields of C6H6 at each temperature. Another major product, C6H5CH3, formed by the recombination of C6H5 and CH3, could also be quantitatively modeled using the known rate constant for the reaction. A weighted least‐squares analysis of the two sets of data gave k1 = 1011.32±0.05 exp[−(2236 ± 91)/T] cm3 mol−1 s−1 and k2 = 1011.37±0.03 exp[−(1925 ± 48)/T] cm3 mol−1 s−1 for the temperature range studied. The result of our sensitivity analysis clearly supports that the yields of C6H6 and C6H5CH3 depend primarily on the abstraction reactions and C6H5 + CH3, respectively. From the absolute rate constants for the two reactions we obtained the value for the H‐abstraction from a primary C‐H bond, k‐CH = 1010.40±0.06 exp(−1790 ± 102/T) cm3 mol−1 s−1. This result is utilized for analysis of other kinetic data measured for C6H5 reactions with alkanes in solution as well as in the gas phase. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 33: 64–69, 2001  相似文献   

8.
Multiple mechanisms of backbiting and β‐scission reactions in free‐radical polymerization of methyl acrylate are modeled using different levels of theory, and the rigid‐rotor harmonic‐oscillator (RRHO) and hindered‐rotor (HR) approximations. We identify the most cost‐effective computational method(s) for studying the reactions and assess the effects of different factors (e.g., functional type and chain length) on thermodynamic quantities, and then identify the most likely mechanisms with first‐principles thermodynamic calculations and simulations of nuclear magnetic resonance (NMR) spectra. To this end, the composite method G4(MP2)‐6X is used to calculate the energy barrier of a representative backbiting reaction. This calculated barrier is then compared with values obtained using density functional theory (DFT) (B3LYP, M06‐2X, and PBE0) and a wavefunction‐based quantum chemistry method (MP2) to establish the benchmark method. Our study reveals that the barriers predicted using B3LYP, M06‐2X, and G4(MP2)‐6X are comparable. The entropies calculated using the RRHO and HR approximations are also comparable. DFT calculations indicate that the 1:5 backbiting mechanism with a six‐membered ring transition state and 1:7 backbiting with an eight‐membered ring transition state are energetically more favored than 1:3 backbiting and 1:9 backbiting mechanisms. The thermodynamic favorability of 1:5 versus 1:7 backbiting depends on the live polymer chain length. The activation energies and rate constants of the left and right β‐scission reactions are nearly equal. The calculated and experimental 13C and 1H NMR chemical shifts of polymer chains affected by backbiting and β‐scission reactions agree with each other, which provides further evidence in favor of the proposed mechanisms. © 2013 Wiley Periodicals, Inc.  相似文献   

9.
The molecules of methyl 3‐(2‐nitrophenylhydrazono)butanoate, C11H13N3O4, (I), and methyl 3‐(2,4‐dinitrophenylhydrazono)butanoate, C11H12N4O6, (II), both prepared from methyl 3‐oxobutanoate and the corresponding nitrophenylhydrazine, exhibit polarized molecular electronic structures; in each of (I) and (II), the molecules are linked into chains by a single C—H...O hydrogen bond. The molecules of 5‐hydroxy‐3‐methyl‐1‐phenyl‐1H‐pyrazole, C10H10N2O, (III), prepared by the reaction of methyl 3‐oxobutanoate and phenylhydrazine, are linked into chains by a single O—H...N hydrogen bond. The reaction between methyl 3‐oxobutanoate and 3‐nitrophenylhydrazine yields 5‐hydroxy‐3‐methyl‐1‐(3‐nitrophenyl)‐1H‐pyrazole, (IV), which when crystallized from acetone yields 4‐isopropylidene‐3‐methyl‐1‐(3‐nitrophenyl)‐1H‐pyrazol‐5(4H)‐one, C13H13N3O3, (V).  相似文献   

10.
We have employed high-level coupled cluster methods including connected triple excitations to study the possibility of symmetry-breaking in the (2)B(2) ground state of the c-C(3)C(2)H radical. Specifically, we find that spin-restricted open-shell Hartree-Fock (ROHF) reference orbitals yield a C(2v) structure, whereas spin-unrestricted Hartree-Fock (UHF) and Brueckner orbitals lead to a symmetry-broken C(s) minimum-energy geometry. Equation-of-motion coupled cluster singles and doubles method for ionized states yields a C(s) structure with a double-zeta basis set, but not with a triple-zeta basis set. Through a detailed analysis of the orbital instability/near-instability behavior of each type of Hartree-Fock reference, we have determined that the UHF reference wave function is more reliable than the ROHF reference in this case and that the Born-Oppenheimer potential surface for c-C(3)C(2)H exhibits a symmetry broken C(s) global minimum. This result is supported by excited-state computations, which indicate that a second-order (pseudo) Jahn-Teller interaction is responsible for the symmetry-breaking.  相似文献   

11.
Polyethylene‐block‐poly(methyl methacrylate) (PE‐b‐PMMA) was successfully synthesized through the combination of metallocene catalysis with living radical polymerization. Terminally hydroxylated polyethylene, prepared by ethylene/allyl alcohol copolymerization with a specific zirconium metallocene/methylaluminoxane/triethylaluminum catalyst system, was treated with 2‐bromoisobutyryl bromide to produce terminally esterified polyethylene (PE‐Br). With the resulting PE‐Br as an initiator for transition‐metal‐mediated living radical polymerization, methyl methacrylate polymerization was subsequently performed with CuBr or RuCl2(PPh3)3 as a catalyst. Then, PE‐b‐PMMA block copolymers of different poly(methyl methacrylate) (PMMA) contents were prepared. Transmission electron microscopy of the obtained block copolymers revealed unique morphological features that depended on the content of the PMMA segment. The block copolymer possessing 75 wt % PMMA contained 50–100‐nm spherical polyethylene lamellae uniformly dispersed in the PMMA matrix. Moreover, the PE‐b‐PMMA block copolymers effectively compatibilized homopolyethylene and homo‐PMMA at a nanometer level. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 3965–3973, 2003  相似文献   

12.
The kinetics of the C2H5 + Cl2, n‐C3H7 + Cl2, and n‐C4H9 + Cl2 reactions has been studied at temperatures between 190 and 360 K using laser photolysis/photoionization mass spectrometry. Decays of radical concentrations have been monitored in time‐resolved measurements to obtain reaction rate coefficients under pseudo‐first‐order conditions. The bimolecular rate coefficients of all three reactions are independent of the helium bath gas pressure within the experimental range (0.5–5 Torr) and are found to depend on the temperature as follows (ranges are given in parenthesis): k(C2H5 + Cl2) = (1.45 ± 0.04) × 10?11 (T/300 K)?1.73 ± 0.09 cm3 molecule?1 s?1 (190–359 K), k(n‐C3H7 + Cl2) = (1.88 ± 0.06) × 10?11 (T/300 K)?1.57 ± 0.14 cm3 molecule?1 s?1 (204–363 K), and k(n‐C4H9 + Cl2) = (2.21 ± 0.07) × 10?11 (T/300 K)?2.38 ± 0.14 cm3 molecule?1 s?1 (202–359 K), with the uncertainties given as one‐standard deviations. Estimated overall uncertainties in the measured bimolecular reaction rate coefficients are ±20%. Current results are generally in good agreement with previous experiments. However, one former measurement for the bimolecular rate coefficient of C2H5 + Cl2 reaction, derived at 298 K using the very low pressure reactor method, is significantly lower than obtained in this work and in previous determinations. © 2007 Wiley Periodicals, Inc. Int J Chem Kinet 39: 614–619, 2007  相似文献   

13.
Kinetics and mechanisms for the reactions of HNO with CH3 and C6H5 have been investigated by ab initio molecular orbital (MO) and transition‐state theory (TST) and/or Rice‐Ramsperger‐Kassel‐Marcus/Master Equation (RRKM/ME) calculations. The G2M(RCC, MP2)//B3LYP/6‐31G(d) method was employed to evaluate the energetics for construction of their potential energy surfaces and prediction of reaction rate constants. The reactions R + HNO (R = CH3 and C6H5) were found to proceed by two key product channels giving (1) RH + NO and (2) RNO + H, primarily by direct abstraction and indirect association/decomposition mechanisms, respectively. As both reactions initially occur barrierlessly, their rate constants were evaluated with a canonical variational approach in our TST and RRKM/ME calculations. For practical applications, the rate constants evaluated for the atmospheric‐pressure condition are represented by modified Arrhenius equations in units of cm3 mol?1 s?1 for the temperature range 298–2500 K: κ1A = 1.47 × 1011 T 0.76 exp[?175/ T ], κ2A = 8.06 × 103 T 2.40 exp[?3100/ T ], κ1B = 3.78 × 105 T 2.28 exp[230/ T ], and κ2B = 3.79 × 109 T 1.19 exp[?4800/ T ], where A and B represent CH3 and C6H5 reactions, respectively. Based on the predicted rate constant at 1 atm pressure for R + HNO → RNO + H, we estimated their reverse rate constants for R + HNO production from H + RNO in units of cm3 mol?1 s?1: κ?2A′ = 7.01 × 1010 T 0.84 exp[120/ T ] and κ?2B′ = 2.22 × 1019 T ?1.01 exp[?9700/ T ]. The heats of formation at 0 K for CH3NO, CH3N(H)O, CH3NOH, C6H5N(H)O, and C6H5NOH have been estimated to be 18.6, 18.1, 22.5, 47.2, and 50.7 kcal mol?1 with an estimated ±1 kcal mol?1 error. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 261–274, 2005  相似文献   

14.
The kinetics of addition reactions between methyl α-eleostearate which forms the main chain of tung oil and cresols when catalyzed by an acid, p-toluene sulfonic acid, have been studied. The addition reactions, carried out with any one of the o-, m-, and p-cresols were shown to be first order with regard to both methyl α-eleostearate and cresol concentrations. The reactions were additions of two cresol molecules to one methyl α-eleostearate molecule, and it was presumed that they proceed in the two steps given below in which the first step in rate-determining. (1) (2) (E: methyl/α-eleostearate, C: cresols) The apparent reaction rate constants (L/mol min) were found to be 0.046 for o-cresol, 0.038 for m-cresol, and 0.033 for p-cresol. The apparent activation energies (kcal/mol) were found to be 0.95, 3.66, and 4.05, in the cases of o-, m-, and p-cresols, respectively.  相似文献   

15.
ABA block copolymers of methyl methacrylate and methylphenylsilane were synthesized with a methodology based on atom transfer radical polymerization (ATRP). The reaction of samples of α,ω‐dihalopoly(methylphenylsilane) with 2‐hydroxyethyl‐2‐methyl‐2‐bromoproprionate gave suitable macroinitiators for the ATRP of methyl methacrylate. The latter procedure was carried out at 95 °C in a xylene solution with CuBr and 2,2‐bipyridine as the initiating system. The rate of the polymerization was first‐order with respect to monomer conversion. The block copolymers were characterized with 1H NMR and 13C NMR spectroscopy and size exclusion chromatography, and differential scanning calorimetry was used to obtain preliminary evidence of phase separation in the copolymer products. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 30–40, 2003  相似文献   

16.
Using a relative rate method, rate constants have been measured at 296 ± 2 K for the gas‐phase reactions of OH radicals with 1,2‐butanediol, 2,3‐butanediol, 1,3‐butanediol, and 2‐methyl‐2,4‐pentanediol, with rate constants (in units of 10?12 cm3 molecule?1 s?1) of 27.0 ± 5.6, 23.6 ± 6.3, 33.2 ± 6.8, and 27.7 ± 6.1, respectively, where the error limits include the estimated overall uncertainty of ±20% in the rate constant for the reference compound. Gas chromatographic analyses showed the formation of 1‐hydroxy‐2‐butanone from 1,2‐butanediol, 3‐hydroxy‐2‐butanone from 2,3‐butanediol, 1‐hydroxy‐3‐butanone from 1,3‐butanediol, and 4‐hydroxy‐4‐methyl‐2‐pentanone from 2‐methyl‐2,4‐pentanediol, with formation yields of 0.66 ± 0.11, 0.89 ± 0.09, 0.50 ± 0.09, and 0.47 ± 0.09, respectively, where the indicated errors are the estimated overall uncertainties. Pathways for the formation of these products are presented, together with a comparison of the measured and estimated rate constants and product yields. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 310–316, 2001  相似文献   

17.
18.
We developed a simple route to prepare stabilized micelles and nanovesicles in aqueous solutions. A hydrophobic poly(succinimide) (PSI) was conjugated with the hydrophilic poly(ethylene glycol) (PEG) as a new type of cross‐linkable unit. Spherical aggregates were formed when dissolving the amphiphilic PEG682b‐PSI130 copolymer in aqueous solutions directly, and polymer nanovesicles were prepared by a precipitation‐dialysis method using PEG455b‐PSI130 copolymer. Bifunctional primary amine was added to the micelle or nanovesicle solutions to prepare cross‐linked structures via aminolysis reaction of the succinimide units. The degree of cross‐linking was controlled by adjusting the molar ratio of the cross‐linker to the succinimide units. Increasing the degree of cross‐linking leads to the compaction of the micelle core thus reduced diameter. The cross‐linked polymer micelles or nanovesicles maintained their morphology in extremely diluted solutions because of their structural stability. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2010  相似文献   

19.
20.
We found that a novel Pd(0)‐catalyzed polycondensation of methyl propargyl carbonate ( 1 ) and bisphenol analogs 4 such as 4,4′‐dihydroxydiphenyl ether (4b ) proceeded successfully and gave polyethers 5 having exomethylene groups. The polycondensation was affected by the kind of bisphenol analogs used. The molar ratio of 1 and 4 also affected the polymerization behavior. Polyethers 5 with higher molecular weights were obtained by polymerization with 0.5 equiv of 4 toward 1 . The use of 1 equiv of 4 gave poor results. When 4,4′‐dihydroxybenzophenone ( 4d ) was employed as a nucleophile, benzenesulfonamide ( 13 ) was needed as an additive for obtaining the corresponding polyether with higher molecular weight. The Pd(0)‐catalyzed condensation of 1 and p‐hydroxyacetophenone was conducted as a model reaction. The desired compound and three by‐products were produced. Two of the by‐products were formed by reaction of 1 and methanol liberated from 1 . The other one appeared to be formed via hydrogenolysis. The formation of the by‐products suggested that some terminal reactions took place in the polycondensation of 1 and 4 . © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 2250–2261, 2008  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