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.
For the polymerization of n-octadecene-1 with catalysts derived from titanium tetrachloride and triethylaluminum, it has been shown that, in addition to polymerization of the olefin, the formation of isomerized olefin occurs. The latter is predominantly trans-n-octadecene-2 and its formation is favored by increase in Al:Ti mole ratio, in catalyst concentration, and in reaction temperature. It has also been shown that 1,1-disubstituted ethylene is present in the nonpolymeric reaction products. The influence of added trans-n-octadecene-2 or trans-n-octadecene-7 on the polymerization of n-octadecene-1 has been studied, and it is shown that the 2-isomer has the more pronounced effect on polymer yield and intrinsic viscosity. It has also been shown that no significant copolymerization of terminal with nonterminal octenes or octadecenes occurs under these conditions. Results indicate that, in polymerizations of this kind, the interaction of catalyst with isomerized monomer is probably an important factor in determining polymer yield and molecular weight. The isomerization reaction is also of interest as a general preparative method for trans-2-olefins.  相似文献   

2.
The directionality (Dn) of a polymerization mechanism can be defined as the average orientation of the reaction or the degree of randomness of head-to-tail, head-to-head, tail-to-tail or tail-to-head reactions within the polymer at the n-th monomer or reaction (n ≫ 2). Directionality of a polymer type can be defined as the average orientation inside a polymer. If the directionality is random, Dn = 0.5. For full directionality, Dn = 1. It is demonstrated that these situations correspond to step and chain polymerization, respectively. Directionality can be as important for the functional properties of polymers as length and composition distributions and the functionality of the original monomers.  相似文献   

3.
Rate constants and activation parameters are reported for the decarboxylation of malonic acid in seven normal alkanols (butanol-l to decanol-l inclusive). It is found that the enthalpy of activation of the reaction is a linear function of the number of carbon atoms in the hydrocarbon chain of tthe solvent, expressed by the equation: ΔH = –600n + 30,000, where n is thenumber of carbon atoms in the chain. Also an equation is developed relatingthe rate constant for the decarboxylation of malonic acid in normal alkanols to n (the number of carbon atoms in the chain): log K = 10.854283 – 0.3212674n + (131.136876n – 6556.5438)/T + log T. With the aid of this equation rate constants may be calulated for the decarboxylationof malonic acid in any alcohol at any temperature which agree with experimental values to within the limit of error of the experiments. A comparison of the data obtained in the present research for the decarboxylation of malonic acid in normal alkanols with previously reported data for the reaction in amines indicates that for reaction taking place in alcohols the transition state probably contains two molecules of solvent but only one for the reaction in amines.  相似文献   

4.
The reaction of glycerol with tert-butyl alcohol in the liquid phase on acid Amberlyst-type ion-exchange resins was studied. The influence of temperature, mole ratio n(TBA)/n(G), water and swelling of gel, and macroreticular type of polymer catalysts on etherification reaction was investigated. The most favourable reaction temperature is 75°C. The conversion of glycerol and yield of glycerol tert-butyl ethers has increased with the mole ratio n(TBA)/n(G). Dry form of macroreticular catalysts provided the best results. Etherification reaction of glycerol with isobutylene in non-aqueous conditions gives the highest yield of desired ethers. The influence of water was studied. The gel forms of ion-exchange resins have very low catalytic activity. It can be concluded that water has an inhibition effect on ion-exchange resins. By comparing the gel and macroreticular forms of Amberlyst ion-exchange resins it can be concluded that very acid forms of macroreticular ion-exchange resins with a high degree of crosslinking are more active catalysts for the studied reaction due to their pores which are sufficiently large so that the voluminous tert-butyl ethers of glycerol can be formed. It was estimated that tert-butyl alcohol as tert-butylation agent is not suitable for etherification of glycerol with the formation of di-and triethers.  相似文献   

5.
The urethane reactions of 1,2-propanediol, 1,3-propanediol, and n-propanol with phenyl isocyanate were respectively carried out in nitrogenous solvents. In situ FT-IR was used to monitor the reactions, and rate constants were determined. It was shown that the reaction rate of 1,2-propanediol was fastest, followed by the reaction rates of 1,3-propanediol and n-propanol. After that, activation energy (Ea), activation enthalpy (ΔH), and activation entropy (ΔS) were calculated. It was found that these thermodynamic parameters for 1,2-propanediol and 1,3-propanediol are very similar, but they were very different from those of n-propanol, which is very useful to understand the urethane reaction mechanism.  相似文献   

6.
The efficiency of dye-sensitised nanocrystalline solar cells is limited in part by the back reaction of photo-injected electrons with tri-iodide ions present in the electrolyte. Competition between this back reaction and the collection of electrons by diffusion to the substrate contact can be described in terms of the electron diffusion length Ln=(Dnτn)1/2, where Dn is the electron diffusion coefficient and τn is the electron lifetime determined by the rate of reaction of electrons with tri-iodide. Dn and τn have been determined over five orders of magnitude of illumination intensity using intensity-modulated photocurrent and intensity-modulated photovoltage spectroscopy. It has been found that τn decreases with light intensity, whereas Dn increases. As a consequence, the electron diffusion coefficient Ln is only weakly intensity dependent, and the incident photon to current conversion efficiency (IPCE) is predicted to be almost independent of intensity. The experimental IPCE agrees well with the predicted values. The results suggest that the kinetics of the back reaction of electrons with tri-iodide couple may be second order in electron density.  相似文献   

7.
The energy, the Gibbs energy of the reaction OH-·(H2O) n- 1 + H2O = OH-·(H2O) n are calculated by the Monte-Carlo method with a large canonical ensemble for n = 1, ..., 20. The ion-waternonpair interaction potential was obtained by numerical fitting of calculated Gibbs energy and entropy of (H2O)n clusters (n = 1, ..., 5) to experimental ones. A good fit to experiment both of the internal energy and the Gibbs energy can be obtained in terms of a model allowing for nonpair interaction. It is shown that constructing an ion-water interaction potential without allowance for the entropy factor can lead to considerable errors in the Gibbs energy of cluster formation and in the nucleation rate.  相似文献   

8.
Hong Wang  Lin Wu 《中国化学》2011,29(4):735-740
Density functional calculations have been carried out on a series of fluorinated empty cages XnFn(n=2–20) with X?Si, Ge, and Sn. It indicates that the fullerene‐like cage structure with pentagons turns out to be the most stable with n increasing, and the stability of the XnFn isomers increases with the number of five‐membered rings. The HOMO‐LUMO gap for Ge (n=6, 10) cages is found to be even larger than the values for Si cages, though in bulk Ge has a smaller band gap than Si. Moreover, calculation of the Gibbs free energy of oligomerization reaction of SiF→1/n (SiF)n showed that this reaction is exothermic even at 900 K, indicating the favorability of their formation from the SiF monomer.  相似文献   

9.
The kinetics of the N,N′,N″-pentamethyl dipropylene triamine (PMPT)-catalyzed reaction of phenyl isocyanate with n-butanol was studied in acetonitrile between 26.5 and 50°C by measuring the NCO disappearance as well as the formation of the various reaction products by means of the standard dibutylamine back-titration method and the high-performance liquid chromatography (HPLC) method. The resulting products from the phenyl isocyanate and n-butanol reaction were found to be N-butyl phenylcarbamate, N-butyl-α,γ-diphenylallophanate, and triphenylisocyanurate. Trimer formed at the expense of carbamate formation even at a high OH/NCO ratio. Allophanate appeared to be an intermediate in the formation of trimer. PMPT was found to be a urethane and trimerization catalyst for the model reaction of phenyl isocyanate with n-butanol in acetonitrile. The PMPT-catalyzed reaction of phenyl isocyanate with n-butanol in the presence of water in acetonitrile at 50°C was also investigated. The resulting reaction products consisted of n-butyl phenylcarbamate, n-butyl-β,γ-diphenylallophanate, triphenylisocyanurate, sym-triphenylbiuret, and N,N′-diphenylurea. The presence of water retarded the disappearance of NCO groups as well as the trimer formation. Aniline (the product of phenyl isocyanate and water) was detected in the reaction of equivalent amounts of phenyl isocyanate and water in acetonitrile.  相似文献   

10.
The intercalation process ofn-butylamine was investigated. The adsorption ofn-butylamine in interlamellar space had stepwise character in case of both crystalline forms of zirconium phosphate. The intercalatedn-butylamine existed at low concentration as bilayered complex. The reaction heat was determined by a microcalorimetric method. It was found that about 90% of it refers to the neutralization ofn-butylamine and only about 10% is related with surface adsorption (ion exchange). The steps of adsorption are 6.0 J/g and 1.0 J/g reaction heat values, respectively. The enthalpy balance of total process in dilute solution system (c 0=3.0 vol%) is 14.67 kJ/mol. The calculated value for ion adsorption (exchange) was 1.37 kJ/mol.  相似文献   

11.
Rates of reactions can be expressed as dn/dt = kf(n), where n is moles of reaction, k is a rate constant, and f(n) is a function of the properties of the sample. Instrumental measurement of rates requires c(dn/dt) = ckf(n), where c is the proportionality constant between the measured variable and the rate of reaction. When the product of instrument time constant, τ, and k is ? 1, the reaction is much slower than the time response of the instrument and measured rates are unaffected by instrument response. When τ k < 1, = 1, or >1, the reaction rate and instrument response rate are sufficiently comparable that measured rates are significantly affected by instrument response and correction for instrument response must be done to obtain accurate reaction kinetics. This paper describes a method for simultaneous determination of τ, k, c, and instrument baseline by fitting equations describing the combined instrument response and rate law to rates observed as a function of time. When τ cannot be neglected, correction for instrument response has previously been done by truncating early data or by use of the Tian equation. Both methods can lead to significant errors that increase as τk increases. Inclusion of instrument baseline as a fitting parameter significantly reduced variability in k and c compared with use of measured instrument baselines. The method was tested with data on the heat rate from acid‐catalyzed hydrolysis of sucrose collected with three types of calorimeters. In addition, to demonstrate the generality of this method of data analysis, equations including τ, k, c, and instrument baseline are derived for the relation between the reaction rate and the observed rate for first order, second order (first in each reactant), nth order in one reactant, autocatalytic, Michaelis–Menten kinetics, and the Ng equation. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 43: 53–61, 2011  相似文献   

12.
The same regioselectivity can be obtained in the CuI catalyzed allylic coupling of n‐butylzinc reagents prepared by either pre‐transmetallation or in situ transmetallation of Grignard reagents in the presence of allylic partner and catalyst. n‐Butylzinc bromide and di‐n‐butylzinc undergo γ‐selective allylation whereas tri‐n‐butylzincate gives preferential α‐selectivity. The regioselectivity obtained in the reaction of n‐butyl bromide and E‐crotyl chloride in the presence of Mg and CuCN is parallel to the coupling of preformed n‐butylmagnesium bromide. It is remarkable that the regiochemical outcome of copper catalyzed alkyl‐allyl coupling can be controlled by using Grignard reagents prepared under Barbier conditions and alkylzincs prepared by in situ transmetallation. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

13.
2-Pyrrolidones have aroused enormous interest as a useful structural moiety in drug discovery; however, not only does their syntheses suffer from low selectivity and yield, but also it requires high catalyst loadings. The radical cyclization of 1,n-enynes and 1,n-dienes has demonstrated to be an attractive method for the synthesis of 2-pyrrolidones due to its mild reaction conditions, fewer steps, higher atom economy, excellent functional group compatibility, and high regioselectivity. Furthermore, radical receptors with unsaturated bonds (i. e. 1,n-enynes and 1,n-dienes) play a crucial role in realizing radical cyclization because of the ability to selectively introduce one or more radical sources. In this review, we discuss representative examples of methods involving the radical cyclization of 1,n-enynes and 1,n-dienes published in the last five years and discuss each prominent reaction design and mechanism, providing favorable tools for the synthesis of valuable 2-pyrrolidone for a variety of applications.  相似文献   

14.
An improved method of isomolar series for determining the composition of an aqueous binary spectrophotometric system is proposed. It is based on the determination by dual-wavelength spectrophotometry of the equilibrium concentration of the chromogenic agent L added in excess, the absorbance of which has been found to be a linear function ofn/m, the composition index of the reaction product M m L n in the system, Hence it is a direct and reliable method, applicable also to systems in which the absorption spectra of L and M m L n overlap and can be resolved by instrumental masking.  相似文献   

15.
The reaction of adsorbed oxygen (Oads) with aliphatic alcohols n-C n H2n + 1OH with n = 2–5 is studied by the method of transients of open-circuit potential in combination with potentiodynamic pulses. It is shown that these alcohols react with Oads by a mechanism the same as for CH3OH. Kinetic parameters of these reactions are determined in ranges of high and medium surface coverages with Oads. These data together with analogous results obtained earlier for CH3OH were studied with the aim of elucidating how the length of the hydrocarbon chain affects the kinetics of interaction of alcohols with Oads. The complex variations of the reaction rate with n (with a maximum) are explained by several factors among which the energy of the C–H bond at α-carbon atom and the degree hydration of alcohols should be singled out.  相似文献   

16.
The carriers of the diffuse interstellar bands (DIBs) are still largely unknown although polycyclic aromatic hydrocarbons, carbon chains, and fullerenes are likely candidates. A recent analysis of the properties of n-acenes of general formula C4n+2H2n+4 suggested that these could be potential carriers of some DIBs. Dehydrogenation reactions of n-acenes after absorption of an interstellar UV photon may result in dehydroacenes. Here the reaction energies and barriers for formation of n-cyclacenes from 2,3-didehydroacenes (n-DDA) by intramolecular Diels–Alder reaction to dihydro-etheno-cyclacenes (n-DEC) followed by ejection of ethyne by retro-Diels–Alder reactions are analyzed using thermally assisted occupation density functional theory (TAO-DFT) for n=10–20. It is found that the barriers for each of the steps depend on the ring strain of the underlying n-cyclacene, and that the ring strain of n-DEC is about 75 % of that of the corresponding n-cyclacene. In each case, ethyne extrusion is the step with the highest energy barrier, but these barriers are smaller than CH bond dissociation energies, suggesting that formation of cyclacenes is an energetically conceivable fate of n-acenes after multiple absorption of UV photons.  相似文献   

17.
Rate constants and activation parameters are reported for the decarboxylation of methylmalonic acid and n-octadecylmalonic acid in three normal alkanols (hexanol-1, octanol-1, and decanol-1). Enthalpies of activation for both substrates in the various solvents are found to be a linear function of the number of carbon atoms or methylene groups in the hydrocarbon chain of the solvent. For both reaction series the isokinetic temperature is found to be equal to the melting point of the substrate. The free energy of activation at the isokinetic temperature in kcal/mole is 29.0 for n-octadecylmalonic acid and 29.4 for methylmalonic acid. Based on the results of the present investigation as well as on previously reported data in the case of malonic acid and n-butylmalonic acid, an empirical method of calculating the rate of reaction for the decarboxylation of malonic acid and its n-alkyl derivatives in normal alkanols is proposed. As a further test of the method of calculation the decarboxylation of n-dodecylmalonic acid in heptanol-1 at 110.30°C was studied. The calculated value of the pseudo-first-order specific reaction velocity constant of the reaction agreed with the experimental value to within about 0.1 percent.  相似文献   

18.
采用浸渍法制备了Ni掺杂的Cu-ZnO催化剂,采用多种物理化学手段研究了其化学物理性质及甘油加氢制取1,2-丙二醇反应催化性能。结果发现,金属Ni助剂的引入可以进一步优化Ni-Cu-ZnO催化剂的甘油加氢生成1, 2-丙二醇的反应活性。少量金属Ni的加入,Ni-Cu-ZnO催化剂的甘油转化率变化不大,生成1, 2-丙二醇的选择性明显增加。而进一步增加Ni含量到nNi/nCu=0.5,Ni含量过高会导致Ni-Cu-ZnO催化剂中实际Cu原子的量减少,从而导致甘油转化率下降。Ni掺杂的Cu-ZnO催化剂甘油加氢性能稳定性较好,在反应102 h后没有明显变化。  相似文献   

19.
用NMR谱研究了n-(对位取代苯基)氮杂-15-冠与Na(Ⅰ)的配位反应,测定了Na(Ⅰ)配合物的稳定常数.发现这些稳定常数与其极限化学位移之间存在着线性关系.  相似文献   

20.
A simple and high-throughput method for the identification of disulfide-containing peptides utilizing peptide-matrix adducts is described. Some commonly used matrices in MALDI mass spectrometry were found to specifically react with sulfhydryl groups within peptide, thus allowing the observation of the peptide-matrix adduct ion [M+n+n′ matrix+H]+ or [M+n+n′ matrix+Na]+ (n = the number of cysteine residues, n′=1, 2,…, n) in MALDI mass spectra after chemical reduction of disulfide-linked peptides. Among several matrices tested, α-cyano-4-hydroxycinnamic acid (CHCA, molecular mass 189 Da) and α-cyano-3-hydroxycinnamic acid (3-HCCA) were found to be more effective for MALDI analysis of disulfide-containing peptides/proteins. Two reduced cysteines involved in a disulfide bridge resulted in a mass shift of 189 Da per cysteine, so the number of disulfide bonds could then be determined, while for the other matrices (sinapinic acid, ferulic acid, and caffeic acid), a similar addition reaction could not occur unless the reaction was carried out under alkaline conditions. The underlying mechanism of the reaction of the matrix addition at sulfhydryl groups is proposed, and several factors that might affect the formation of the peptide-matrix adducts were investigated. In general, this method is fast, effective, and robust to identify disulfide bonds in proteins/peptides.  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