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.
Condensation of triethoxymethane and aniline with dimedone gives 2-anilinomethylene dimedone as the main product. An1H-NMR-spectroscopic investigation of the kinetics in chloroform-d 1 and methanol-d 4 shows different rate determining steps in these solvents. There are two predominant rate determining steps in a complicated multistep reaction sequence.The initial one involves condensation of aniline with triethoxymethane to form diphenyl formamidine via ethyl phenyl formimidate. The second step involves reaction of the intermediate diphenyl formamidine with dimedone. The rates are strongly dependent upon the nature of the solvent and the concentration of catalytic acid. In methanol the reaction of dimedone with the intermediate diphenyl formamidine is rate determining. For preparative purposes the isolation of the intermediate diphenyl formamidine and the subsequent use of less polar solvents offer an advantage, because the second step is found to be accelerated.
6. Mitt.:Wolfbeis, O. S., Mh. Chem.112, 369 (1981).  相似文献   

2.
The degradation of 4-chlorophenol (4-CP) by using gamma rays generated by a 60Co source in the presence of O3 was investigated. The radiolysis of 4-CP and the kinetics of 4-CP mineralization were analyzed based on the determination of total organic carbon (TOC). The influence of initial 4-CP concentration and the free radicals scavengers (such as NaHCO3 and t-butanol) on the 4-CP degradation was also studied. The results showed that when the radiation rate was 336 Gy·min−1, 4-chlorophenol at concentration of 10 mg·L−1 could be completely degraded at the radiation dose of 2 kGy. The degradation of 4-chlorophenol could be described by a first-order reaction model, the rate constant of 4-CP degradation by combined ozonation and radiation was 0.1016 min−1, which was 2.4 times higher than the sum of radiation (0.0294 min−1) and ozonation (0.0137 min−1). It revealed that the combination of radiation and ozonation resulted in synergistic effect, which can remarkably increase the degradation efficiency of 4-CP.  相似文献   

3.
Polymers having 2,5‐diphenyl‐1,3,4‐oxadiazole (BCO) or anthracene (BCA) as an electron transport unit and N,N′‐diphenyl‐N,N′‐bis(4‐butylphenyl)‐benzidine (BTPD) as a hole transport unit were prepared by condensation polymerization using Friedel–Crafts reaction. It was found that BCO was less reactive than BCA. The low reactivity of the BCO monomer can be explained by the oxygen atom in the oxadiazole unit, which acts as a Lewis base and reduces the activity of the catalyst. The redox behavior measured by cyclic voltammetry showed for both BTPD‐BCO and BTPD‐BCA almost the same oxidation potential. In addition, the BTPD‐BCO also exhibited a reduction peak. Hole and electron drifts mobility of the polymers were measured by the time‐of‐flight method. The hole drift mobility of both BTPD‐BCO and BTPD‐BCA was 7.4 × 10?5 cm2 V?1 s?1. The electron drift mobilities of BTPD‐BCO and BTPD‐BCA were 6.5 × 10?5 cm2 V?1 s?1 and 5.2 × 10?6 cm2 V?1 s?1, respectively. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 3083–3089, 2007  相似文献   

4.
Experiments were performed in an attempt to identify the reactive intermediate(s) involved in the degradation of a polyamide, Nylon 66, in chlorinated water. According to previous studies, N‐chlorination is certainly one reaction that ultimately contributes to polyamide degradation. In this case, the intermediates involved could either be Cl2 or HClO. Available information also indicates that, for many polymers, singlet molecular oxygen (a1Δg), chemically generated from HClO, could likewise be involved as an intermediate in a degradation reaction. Thus, tests were undertaken to specifically address this latter issue with respect to polyamide degradation. The degradation of Nylon 66 was monitored under a variety of conditions by FT‐IR spectroscopy. The rate of degradation was pH‐dependent, and degradation was most pronounced at pH<5. Characteristic tests for the intermediacy of singlet oxygen, however, were negative. Rather, the data strongly pointed to Cl2 as the key intermediate in the degradation. The presence of compounds capable of being oxidized by Cl2 arrested the degradation reaction. These results should be pertinent in an attempt to stabilize polymers exposed, e.g., to water that has been chlorinated to kill bacteria (i.e., drinking water).  相似文献   

5.
Two antibacterial and antifungal agents, chloroxylenol (4‐chloro‐3,5‐dimethyl‐phenol) and triclosan (5‐chloro‐2‐(2’,4’‐dichlorophenoxy)‐phenol), were studied experimentally in solid state with an X‐ray, 35Cl‐nuclear quadrupole resonance (NQR) and 17O‐nuclear quadrupole double resonance (NQDR) spectroscopies and, theoretically, with the density functional theory/quantum theory of atoms in molecules (DFT/QTAIM). The crystallographic structure of triclosan, which crystallises in space group P31 with one molecule in the asymmetric unit [a = 12.64100(10), b = 12.64100(10), c = 6.71630(10) Å], was solved with an X‐ray and refined to a final R‐factor of 2.81% at room temperature. The NQR frequencies of 35Cl and 17O were detected with the help of the density functional theory (DFT) assigned to particular chlorine and oxygen sites in the molecules of both compounds. The NQR frequencies at 35Cl sites in chloroxylenol and triclosan were found to be more differentiated than frequencies at the 17O site. The former better describes the substituent withdrawing effects connected to π‐electron delocalization within the benzene rings and the influence of temperature; whereas, those at the 17O site provide more information on O‐H bond and intermolecular interactions pattern. The conformation adopted by diphenyl ether of triclosan in solid state was found to be typical of diphenyl ethers, but the opposite to those adopted when it was bound to different inhibitors. According to an X‐ray study, temperature had no effect on the conformation of the diphenyl ring of triclosan, which was the same at 90 K and at room temperature (RT). The scattering of NQR frequencies reproduced by the DFT under assumption of the X‐ray data at 90 K and RT is found to be a good indicator of the quality of resolution of the crystallographic structure. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

6.
The unperturbed mean-square end-to-end distance 〈R02〉 and its temperature variation d In 〈R02〉/dT for isotactic polypropylene have been estimated from intrinsic viscosity data in three theta solvents, i.e., diphenyl, diphenyl ether, and dibenzyl ether, measured at their θ temperatures as determined by precipitation temperature measurements. The characteristic ratios, 〈R02〉/nl2, where n is the number of bonds of length l in the main chain, evaluated by assuming Φ = 2.87 × 1021, are 5.80 in diphenyl (at θ = 125.1°C.), 5.41 in diphenyl ether (at θ = 142.8°C.), and 4.56 in dibenzyl ether (at θ = 183.2°C.). These values lead to the temperature coefficient d In 〈R02〉/dT = ?4.09 × 10?3 deg.?1 Results are compared with the data previously reported on polyethylene.  相似文献   

7.
Synthesis of a new Class of Chiral β—Mercaptoalcohols from Amino Acids   总被引:1,自引:0,他引:1  
The syntheses of three new optically active β-mercaptoalcohols,(R)-1,1-diphenyl-2-mercapto-3-methyl-1-butanol,(R)-1,1-diphenyl-2-mercapto-4-methyl-1-pentanol,and (R)-1,1-diphenyl-2-mercapto-1-benzenepropanol from the corresponding amino acids are described.The enantiomeric excesses of these β-mercaptoalcohols were determined by ^1H NMR as their (S)-mandeloyl derivatives.  相似文献   

8.
We report the synthesis and characterisation of new examples of meso‐hydroxynickel(II) porphyrins with 5,15‐diphenyl and 10‐phenyl‐5,15‐diphenyl/diaryl substitution. The OH group was introduced by using carbonate or hydroxide as nucleophile by using palladium/phosphine catalysis. The NiPor?OHs exist in solution in equilibrium with the corresponding oxy radicals NiPor?O.. The 15‐phenyl group stabilises the radicals, so that the 1H NMR spectra of {NiPor?OH} are extremely broad due to chemical exchange with the paramagnetic species. The radical concentration for the diphenylporphyrin analogue is only 1 %, and its NMR line‐broadening was able to be studied by variable‐temperature NMR spectroscopy. The EPR signals of NiPor?O. are consistent with somewhat delocalised porphyrinyloxy radicals, and the spin distributions calculated by using density functional theory match the EPR and NMR spectroscopic observations. Nickel(II) meso‐hydroxy‐10,20‐diphenylporphyrin was oxidatively coupled to a dioxo‐terminated porphodimethene dyad, the strongly red‐shifted electronic spectrum of which was successfully modelled by using time‐dependent DFT calculations.  相似文献   

9.
The effect of cetylpyridinium bromide (CPB) and sodium dodecyl sulphate (SDS) on the direction and the rate of the reaction ofo-dimethylaminomethylphenol (MP) withp-nitrophenyl diphenyl phosphate (1) has been studied by31P NMR and spectrophotometry. It was shown that the reaction of MP with1 proceeds in two steps both with and without the surfactant. The product of transesterification is formed in the first step. The second step is hydrolysis catalyzed by the aminomethyl group yielding equal amounts of diphenyl phosphate ando-dimethylaminomethyl phenyl phosphate. The reaction of MP with1 is catalyzed by CPB and inhibited by SDS. The ratio between the rates of the first and the second stages changes in the presence of surfactant. The parameters of the reaction of MP with1 inhibited by micellar SDS were calculated.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 242–245, February, 1994.  相似文献   

10.
Abstract

Three acetanilide herbicides—propanil, propachlor and pronamide—were studied for alkaline degradation using the HPLC technique and found to be relatively stable for chemical degradation with the highest hydrolysis rate for pronamide (k b = 0.4966 M?1 hr?1), followed by propachlor (kb = 0.1466 M?1 hr?1) and propanil kb = 0.0154 M?1 hr?1). Simultaneously, the Ames mutagenicity test for non-degraded and half-degraded acetanilides have been performed. Testing two types of samples (primary chemicals and mixtures of half-degraded primary chemicals with their degradation products, both in 10%-acetonitrile buffer; the ranges of concentrations were 2–400 mg l?1 for propanil and propachlor and 2–50 mg l?1 (for pronamide) have shown the increase in mutagenicity for propanil, the decrease for propachlor and almost the same dose-response curve for pronamide during their degradation. The preliminary results of the research indicate that some environmental problems related to the impact of pesticide degradation products can appear, not only to a pesticide itself.  相似文献   

11.
Molecular depth profiling of polymers by secondary ion mass spectrometry (SIMS) has focused on the use of polyatomic primary ions due to their low penetration depth and high damage removal rates in some polymers. This study is the third in a series of systematic characterizations of the effect of polymer chemistry on degradation under polyatomic primary ion bombardment. In this study, time‐of‐flight SIMS (ToF‐SIMS) was used to assess 5 keV SF5+‐induced damage of ~90 nm thick spin‐cast poly(2‐hydroxyethyl methacrylate) (PHEMA) and ~130 nm thick trifluoroacetic anhydride‐derivatized PHEMA (TFAA‐PHEMA) films. The degradation of these polymers under extended SF5+ bombardment (~2 × 1014 ions cm?2) was compared to determine the effect of the pendant group chemistry on their degradation. The sputter rate and ion‐induced damage accumulation rate of PHEMA were similar to a poly(n‐alkyl methacrylate) of similar pendant group length, suggesting that the addition of a terminal hydroxyl group to the alkyl pendant group does not markedly change the stability of poly(n‐alkyl methacrylates) under SF5+ bombardment. The sputter rate and ion‐induced damage accumulation rate of TFAA‐PHEMA were much higher than a poly(n‐alkyl methacrylate) of similar pendant group length, suggesting that derivatization of the terminal hydroxyl group can significantly reduce degradation of the polymer under SF5+ bombardment. This result is in good agreement with the literature on the thermal and radiation‐induced degradation of fluorinated poly(alkyl methacrylates), which suggests that the electron‐withdrawing fluorinated pendant group increases the probability of depolymerization. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

12.
355 nm光照下利用瞬态吸收光谱技术进行了有氧、无氧条件下二苯醚与亚硝酸体系的反应机理研究, 考察了其中瞬态物种的衰减行为, 并对其光解产物进行了GC-MS分析. 研究表明, HNO2在355 nm紫外光的照射下产生的OH自由基和二苯醚反应生成C12H10O-OH 加合物, N2条件下C12H10O-OH衰减的速率常数为(1.86±0.14)×105 s-1, 在有氧条件下, C12H10O-OH可转化为C12H10O-OHO2, 衰减的速率常数为(6.6±0.4)×106 s-1. N2条件下最终产物为苯酚、2-羟基二苯醚、4-羟基二苯醚、4-硝基二苯醚.  相似文献   

13.
研究了在光照下,乙二胺四乙酸(EDTA)的添加促进石墨相氮化碳(g-C3N4)光催化降解甲基橙(MO).研究了H+和羧酸根负离子对光降解MO的影响.紫外-可见漫反射光谱(DRS)研究表明,EDTA的加入并没有改变g-C3N4的电子结构和光电特性.EDTA的加入捕获了空穴(h+),促进了光生e-/h+对的分离,从而使光降解活性提高.证明了·O2-是光催化降解过程中的主要活性物种.基于上述研究结果,我们提出了一种可能的EDTA促进g-C3N4光催化降解MO的机理.这些结果为提高g-C3N4光催化降解水体中有机污染物的性能提供了一种新方法.  相似文献   

14.
The thermal degradation of poly(2,6-dimethyl-1,4-phenylene ether) has been investigated to 1000°C in an inert atmosphere. X-ray diffraction, thermogravimetric analysis, and differential scanning calorimetry were employed to study the physical changes in the polymer, and vapor-phase chromatography, infrared spectroscopy, and mass spectrometric thermal analysis were used to elucidate the chemical aspects of the degradation process. It was found that degradation occurs in two steps: (1) a rapid exothermic process occurs between 430 and 500°C, leading to the evolution of phenolic products, water, and a black, highly crosslinked residue, and (2) a slower, char-forming process occurs above 500°C, characterized by the evolution of methane, carbon monoxide, and hydrogen. The chars formed in process 2 were found by x-ray analysis to be amorphous. The infrared spectrum of a sample heated to 510°C is nearly identical with that of the starting polymer, indicating that oxidative reactions are not important in the first process. The data for the low-temperature process are consistent with a thermal degradation scheme based on the radical-redistribution reaction of polyphenylene ethers and/or the degradation of o-benzylphenols formed by the thermal rearrangement of o-methyl diphenyl ethers. The char-forming process is best explained by simultaneous operation of the Szwarc mechanism of toluene pyrolysis, producing hydrogen and methane and reactions that cleave the aromatic rings and produce carbon monoxide.  相似文献   

15.
The kinetics of the thermal rearrangement 4‐ethyl‐3,5‐diphenyl‐4H‐1,2,4‐triazoles, 1 , to the corresponding 1‐ethyl‐3,5‐diphenyl‐1‐alkyl‐1H‐1,2,4‐triazoles, 2 , was studied in 15‐Crown‐5 and octadecane at 330 °C. The reaction was very slow in octadecane but proceed well in 15‐Crown‐5. The reaction order for the reaction was not constant but changed from an initial second order rate law towards a first order rate law as the reaction progressed. This was confirmed by the concentration dependent reaction order, nc, which was larger than the time dependent rate law, nt. The rationale for the observation was, that at high substrate concentrations the reaction order was second order while at lower concentrations a competing solvent assisted reaction plays an increasing important role. The data were in agreement with a mechanism in which the neutral 4‐alkyl‐triazoles in an intermolecular nucleophilic displacement reaction form a triazolium triazolate, which in a subsequent nucleophilic reaction gives the observed product.  相似文献   

16.
The objective of this investigation was to verify the degradation of endrin by gamma irradiation. 60Co was used as radiation source for irradiation of 50 mg L?1 endrin with a varied dose of 1–6 kGy. High performance liquid chromatography (HPLC) coupled with diode array detector was used as an analytical technique to monitor the degradation rate along with numbers of degradation products formed. At dose rate of 6 kGy ≥99% of endrin was degraded. It is proposed that utilization of ionization radiations can be an effective and efficient tool for the removal of halogenated pesticides.  相似文献   

17.
Song  Zhen  Chen  Xiaohui  Zhang  Di  Ren  Lei  Fang  Lina  Cheng  Weiming  Gong  Ping  Bi  Kaishun 《Chromatographia》2009,70(11):1575-1580

A selective and validated stability-indicating LC method was developed for the kinetic study of the degradation of PAC-1, which was carried out in aqueous solutions at 37, 60, 80 and 100 °C with pH 1.5–9.0. Separation was performed on a Kromasil C18 column with acetonitrile–water–fomic acid (30:70:0.1, v/v/v) as mobile phase with a flow rate of 1.0 mL min−1 at 281 nm. The degradation rate obtained indicated a first-order reaction law and the activation energy (E a) was calculated. The results showed that temperature and pH values were significant factors affecting the degradation of PAC-1. An unknown degradation product in alkaline condition was isolated using a reverse-phase semi-preparative LC system. The structure of the degradation product is identified as 2-hydroxy-3-(2-propenyl)-[[2-hydroxy-3-(2-propenyl)phenyl]methylene]hydrazone utilizing the 1H NMR, 13C NMR, IR and Q-TOF-MS techniques.

  相似文献   

18.
Summary. The degradation of imazapyr, an imidazolinone herbicide, in aqueous solution has been investigated with TiO2 slurry as photocatalyst at 30°C under UV radiation. The depletion of imazapyr concentration in an aqueous suspension followed 1st order kinetic behavior. The influence of pH and the charge densities of imazapyr geometries were calculated at the semi-empirical AM1 level, and the effect of temperature was investigated. The addition of electron acceptors such as potassium persulfate and hydrogen peroxide showed that the rate constant doubled at least. At higher persulfate concentrations the herbicide degradation was more efficient in direct photolysis than TiO2-photocatalysis. The degradation rate constant increased by 38% upon variation of the temperature between 20.0 and 50.0°C and displayed non-Arrhenius behavior.  相似文献   

19.
The degradation of imazapyr, an imidazolinone herbicide, in aqueous solution has been investigated with TiO2 slurry as photocatalyst at 30°C under UV radiation. The depletion of imazapyr concentration in an aqueous suspension followed 1st order kinetic behavior. The influence of pH and the charge densities of imazapyr geometries were calculated at the semi-empirical AM1 level, and the effect of temperature was investigated. The addition of electron acceptors such as potassium persulfate and hydrogen peroxide showed that the rate constant doubled at least. At higher persulfate concentrations the herbicide degradation was more efficient in direct photolysis than TiO2-photocatalysis. The degradation rate constant increased by 38% upon variation of the temperature between 20.0 and 50.0°C and displayed non-Arrhenius behavior.  相似文献   

20.
A comprehensive study using liquid chromatography electrospray ionization multistage mass spectrometry (LC-ESI MS n ) was performed to get detailed structural information on poly(butylene adipate-co-butylene terephthalate) co-polyester and its product of partial degradation. LC-MS and LC-MSn identified the existence of cyclic structures in the original samples that disappear completely during the degradation. The occurrence of methanol transesterification in the degradation process was confirmed. MS2 on the first 13C isotope peak helped to determine the elemental composition of the fragments and facilitated end group determination. The method can be used to provide an alternative for high mass accuracy MS2 experiments. Sequence information was also revealed for certain copolymeric structures.  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