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 468 毫秒
1.
Abstract

The kinetics and stability constants of l-tyrosine complexation with copper(II), cobalt(II) and nickel(II) have been studied in aqueous solution at 25° and ionic strength 0.1 M. The reactions are of the type M(HL)(3-n)+ n-1 + HL- ? M(HL)(2-n)+n(kn, forward rate constant; k-n, reverse rate constant); where M=Cu, Co or Ni, HL? refers to the anionic form of the ligand in which the hydroxyl group is protonated, and n=1 or 2. The stability constants (Kn=kn/k-n) of the mono and bis complexes of Cu2+, Co2+ and Ni2+ with l-tyrosine, determined by potentiometric pH titration are: Cu2+, log K1=7.90 ± 0.02, log K2=7.27 ± 0.03; Co2+, log K1=4.05 ± 0.02, log K2=3.78 ± 0.04; Ni2+, log K1=5.14 ± 0.02, log K2=4.41 ± 0.01. Kinetic measurements were made using the temperature-jump relaxation technique. The rate constants are: Cu2+, k1=(1.1 ± 0.1) × 109 M ?1 sec?1, k-1=(14 ± 3) sec?1, k2=(3.1 ± 0.6) × 108 M ?1 sec?1, k?2=(16 ± 4) sec?1; Co2+, k1=(1.3 ± 0.2) × 106 M ?1 sec?1, k-1=(1.1 ± 0.2) × 102 sec?1, k2=(1.5 ± 0.2) × 106 M ?1 sec?1, k-2=(2.5 ± 0.6) × 102 sec?1; Ni2+, k1=(1.4 ± 0.2) × 104 M ?1 sec?1, k-1=(0.10 ± 0.02) sec?1, k2=(2.4 ± 0.3) × 104 M ?1 sec?1, k-2=(0.94 ± 0.17) sec?1. It is concluded that l-tyrosine substitution reactions are normal. The presence of the phenyl hydroxyl group in l-tyrosine has no primary detectable influence on the forward rate constant, while its influence on the reverse rate constant is partially attributed to substituent effects on the basicity of the amine terminus.  相似文献   

2.
The self‐reactions of the linear pentylperoxy (C5H11O2) and decylperoxy (C10H21O2) radicals have been studied at room temperature. The technique of excimer laser flash photolysis was used to generate pentylperoxy radicals, while conventional flash photolysis was used for decylperoxy radicals. For the former, the recombination rate coefficients were estimated for the primary 1‐pentylperoxy isomer (n‐C5H11O2) and for the secondary 2‐ and 3‐pentylperoxy isomers combined (“sec‐C5H11O2”) by creating primary and secondary radicals in different ratios of initial concentrations and simulating experimental decay traces using a simplified chemical mechanism. The values obtained at 298 K were: k(n‐C5H11O2+n‐C5H11O2→Products)=(3.9±0.9)×10−13 cm3 molecule−1 s−1; k(sec‐C5H11O2+sec‐C5H11O2→Products)=(3.3±1.2)×10−14 cm3 molecule−1 s−1. Quoted errors are 1σ, whereas the total relative combined uncertainties correspond to an estimated uncertainty factor around 1.65. For decylperoxy radicals, the kinetics of all the types of secondary peroxy isomers reacting with each other were considered equivalent and grouped as sec‐C10H21O2 (as for sec‐C5H11O2). The UV absorption spectrum of these secondary radicals was measured, and the combined self‐reaction rate coefficients then derived as: k(sec‐C10H21O2+sec‐C10H21O2)=(9.4±1.3)×10−14 cm3 molecule−1 s−1 at 298 K. Again, quoted errors are 1σ and the total uncertainty factor corresponds to a value around 1.75. The sec‐dodecylperoxy radical was also investigated using the same procedure, but only an estimate of the rate coefficient could be obtained, due to aerosol formation in the reaction cell: k(sec‐C12H25O2+sec‐C12H25O2)≡1.4×10−13 cm3 molecule−1 s−1, with an uncertainty factor of about 2. Despite the fairly high uncertainty factors, a relationship has been identified between the room‐temperature rate coefficient for the self‐reaction and the number of carbon atoms, n, in the linear secondary radical, suggesting: log(k(sec‐RO2+sec‐RO2)/cm3 molecule−1 s−1)=−13.0–3.2×exp(−0.64×(n‐2.3)). Concerning primary linear alkylperoxy radicals, no real trend in the self‐reaction rate coefficient can be identified, and an average value of 3.5×10−13 cm3 molecule−1 s−1 is proposed for all radicals. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet: 31: 37–46, 1999  相似文献   

3.
An experimental study of the decomposition kinetics of chemically activated 2-methyl-l-butene and 3-methyl-l-butene produced from photolysis of diazomethane-isobutene-neopentane-oxygen mixtures is reported. The experimental rate constants for 3-methyl-l-butene decomposition were 1.74 ± 0.44 × 108 sec?1 and 1.01 ± 0.25 × 108 sec?1 at 3660 and 4358 Å, respectively. 2-Methyl-l-butene experimental decomposition rate constants were found to be 5.94 ± 0.59 × 107 sec?1 at 3660 Å and 3.42 ± 0.34 × 107 sec?1 at 4358 Å. Activated complex structures giving Arrhenius A-factors calculated from absolute rate theory of 1016.6 ± 0.5 sec?1 for 3-methyl-l-butene and 1016.2 ± 0.4 sec?1 for 2-methyl-l-butene, both calculated at 1000°K, were required to fit RRKM theory calculated rate constants to the experimental rate constants at reasonable E0 and E* values. Corrected calculations (adjusted E0 values) on previous results for 2-pentene decomposition gave an Arrhenius A-factor of 1016.45 ± 0.35 sec?1 at 1000°K. The predicted A-factors for these three alkene decompositions giving resonance-stabilized methylully radicals are in good internal agreement. The fact that these A-factors are only slightly less than those for related alkane decompositions indicates that methylallylic resonance in the decomposition products leads to only a small amount of tightening in the corresponding activated complexes. This tightening is a significantly smaller factor than the large reduction in the critical energy due to resonance stabilization.  相似文献   

4.
The decomposition rate of chemically activated ethyltrimethylgermane from the reaction 1CH2 + (CH3)4Ge, where 1CH2 was produced from diazomethane photolysis at 3660 Å, is 8.6 × 105 sec?1. This result combined with RRKM theory and critical energy estimates yields an Arrhenius A factor of log[A (sec?1)/methyl] = 14.7 ± 0.8 for methyl rupture from germanium. Log A values for methyl rupture from carbon, silicon, and germanium linearly correlate with the vibrational-rotational entropies of the corresponding tetramethyls. Extrapolation predicts log[A (sec?1)/methyl] = 14.4 and 14.3 for methyl rupture from tin and lead, respectively.  相似文献   

5.
Contributions to the Chemistry of Phosphorus. 119. Tri-isopropyl-cyclotriphosphane, (i-PrP)3, and Tri-sec-butyl-cyclotriphosphane, (s-BuP)3 The dehalogenation of isopropyldichlorophosphane and sec-butyldichlorophosphane with magnesium leads to the title compounds (i-PrP)3 ( 1 ) and (s-BuP)3 ( 2 ) respectively. The corresponding cyclotetraphosphanes, and in the case of 1 also (i-PrP)5, are formed as by-products. 1 and 2 are relatively stable triorganyl-cyclotriphosphanes. They were isolated in a pure state and have been fully characterized. Due to the chiral P-bonded carbon atoms 2 forms four diastereomers, which could be identified by NMR spectroscopy.  相似文献   

6.
In this study, a series of secondary‐ and tertiary‐amino‐substituted diaryl diselenides were synthesized and studied for their glutathione peroxidase (GPx) like antioxidant activities with H2O2, cumene hydroperoxide, or tBuOOH as substrates and with PhSH or glutathione (GSH) as thiol cosubstrates. This study reveals that replacement of the tert‐amino groups in benzylamine‐based diselenides by sec‐amino moieties drastically enhances the catalytic activities in both the aromatic thiol (PhSH) and GSH assay systems. Particularly, the N‐propyl‐ and N‐isopropylamino‐substituted diselenides are 8–18 times more active than the corresponding N,N‐dipropyl‐ and N,N‐diisopropylamine‐based compounds in all three peroxide systems when GSH is used as the thiol cosubstrate. Although the catalytic mechanism of sec‐amino‐substituted diselenides is similar to that of the tert‐amine‐based compounds, differences in the stability and reactivity of some of the key intermediates account for the differences in the GPx‐like activities. It is observed that the sec‐amino groups are better than the tert‐amino moieties for generating the catalytically active selenols. This is due to the absence of any significant thiol‐exchange reactions in the selenenyl sulfides derived from sec‐amine‐based diselenides. Furthermore, the seleninic acids (RSeO2H) derived from the sec‐amine‐based compounds are more stable toward further reactions with peroxides than their tert‐amine‐based analogues.  相似文献   

7.
Sec-butylbenzene has been investigated using resonant two-photon ionization (R2PI) and UV–UV hole-burning spectroscopy aided by ab initio calculations. All three conformers predicted from theory are observed in the spectrum, and are assigned by rotational band contour analysis. The most strongly populated conformer (G1) has a gauche arrangement of the side chain dihedral angle τ2(C1CαCβCγ). The populations of the anti (A) and the remaining gauche conformer (G2) are about 7% and 2%, respectively. The alpha methyl group is found to significantly affect the conformational preferences in sec-butylbenzene (sec-BB), compared to n-propylbenzene in which the anti conformer is favored.  相似文献   

8.
An absolute value of kr of ethyl radicals at 860 ± 17°K of 4.5 × 109 M?1·sec?1 was determined under VLPP conditions, where the value of kr/kr should be about 1/2. Thus kr(M?1·sec?1) ~ 1010 at 860°K. An error of as much as a factor of 2 in kr would be surprising, but possible. The value of 1010M?1·sec?1 seems to be a factor of from 2 to 5 too high to be compatible with extensive data on the reverse reaction and the accepted thermochemistry. Changes in the heat of formation and entropy of the ethyl radical can change the situation somewhat, but even these changes when applied to the work of Hiatt and Benson [3] indicate that ethyl combination should be ~ 109.3 M?1·sec?1. More work is necessary if a better value is desired.  相似文献   

9.
Quaternization of 2-aziridino-5-chlorobenzophenone (1) with methyl iodide resulted in formation of 2-(N-β-iodoethyl-N-methyl)aminobenzophenone ( 2 ), via an unstable quaternary compound. Rate constants for 1 → 2 conversion, as determined by an nmr method at 35 ± 0.1°, varied between 0.22 × 10?3 sec?1 in DMSO-d6, and 0.95 × 10?6 sec?1 in methanol-d4. Ammonolysis with hexamine, and subsequent cyclization afforded 7-chloro-l-methyl-5-phenyl-2,3-dihydro-lH-1,4-benzodiazepine (3, generic name medazepam) in 92% over-all yield.  相似文献   

10.
Two structurally closely related three‐arm star blocks were synthesized and characterized: tCum(PIB‐b‐PNBD)3 and tCum(PNBD‐b‐PIB)3 [where tCum (tricumyl) stands for the phenyl‐1,3,5‐tris(‐2‐propyl) fragment and PIB and PNBD are polyisobutylene and polynorbornadiene, respectively]. The syntheses were accomplished in two stages: (1) the preparation of the first (or inner) block fitted with appropriate chlorine termini capable of initiating the polymerization of the second (or outer) block with TiCl4 and (2) the mediation of the polymerization of the second block. Therefore, the synthesis of tCum(PIB‐b‐PNBD)3 was effected with tCum(PIB‐Clt)3 [where Clt is tert‐chlorine and number‐average molecular weight (Mn) = 102,000 g/mol] by the use of TiCl4 and 30/70 CH3Cl/CHCl3 solvent mixtures at ?35 °C. PNBD homopolymer contamination formed by chain transfer was removed by selective precipitation. According to gel permeation chromatography, the Mn's of the star blocks were 107,300–109,200 g/mol. NMR spectroscopy (750 MHz) was used to determine structures and molecular weights. Differential scanning calorimetry (DSC) indicated two glass‐transition temperatures (Tg's), one each for the PIB (?65 °C) and PNBD (232 °C) phases. Thermogravimetric analysis thermograms showed 5% weight losses at 293 °C in air and at 352 °C in N2. The synthesis of tCum(PNBD‐b‐PIB)3 was achieved by the initiation of isobutylene polymerization with tCum(PNBD‐Clsec)3 (where Clsec is sec‐chlorine and Mn = 2900 g/mol) by the use of TiCl4 in CH3Cl at ?60 °C. DSC for this star block (Mn = 14,200 g/mol) also showed two Tg's, that is, at ?67 and 228 °C for the PIB and PNBD segments, respectively. It is of interest that the Clsec terminus of PNBD, , readily initiated isobutylene polymerization. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 740–751, 2003  相似文献   

11.
This paper presents a dualistic behavior of 2‐substituted‐3‐hydroxyisoindolones in reactions with sec‐butyllithium (sec‐BuLi). 2‐tert‐Butyl‐3‐hydroxy‐2,3‐dihydro‐1H‐isoindol‐1‐one ( 1a ) treated with sec‐BuLi undergoes metalation at position 7. On the other hand, the reaction between 3‐hydroxy‐2‐phenyl‐2,3‐dihydroxyisoindol‐1‐one ( 1j ) and sec‐BuLi results in 3‐sec‐butyl‐2‐phenyl‐2,3‐dihydroisiondol‐1‐one ( 3j ).  相似文献   

12.
A series of nickel(II) catalysts containing phenyl and chiral sec‐phenethyl groups, {[(4‐R1‐2‐R2C6H2N?C)2Nap]NiBr2} (Nap: 1,8‐naphthdiyl, R1 = Me, R2 = Ph ( 3a ); R1 = Me, R2 = sec‐phenethyl ( 3b ); R1 = Cl, R2 = sec‐phenethyl ( 3c ); R1 = Me, R2 = Me ( 3d ) were synthesized and characterized. All organic compounds were fully characterized by FT‐IR and NMR spectroscopy and elemental analysis. The single crystal for X‐ray crystallography was isolated from 3a in CH2Cl2/n‐hexane under air; the crystal structure showed a binuclear complex 3a ′, in which each nickel atom was six‐coordinate. The two nickel atoms together with two bromine atoms form a planar four‐membered ring, with a bromine and H2O axial ligands. These complexes, activated by diethylaluminum chloride and chiral nickel pre‐catalysts rac‐ 3c , exhibited good activities (up to 2.85 × 106 g PE (mol Ni h bar)?1) for ethylene polymerization, and produced polyethylene products with a high degree of branching (up to 117 branched per 1000 carbons) at high temperature. The type and amount of branches of the polyethylenes obtained were determined by 1H and 13C NMR spectroscopy. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

13.
Trimethylene sulfone and 3? methyl sulfolane have been pyrolyzed using a modification of the toluene flow method and a comparative rate technique. The main decomposition reactions are where k1=1016.1±0.3 exp(?28,100±500/T) sec?1 and k2=1016.1±0.4 exp(?33,200±750/T) sec?1.  相似文献   

14.
By allowing dimethyl peroxide (10?4M) to decompose in the presence of nitric oxide (4.5 × 10?5M), nitrogen dioxide (6.5 × 10?5M) and carbon tetrafluoride (500 Torr), it has been shown that the ratio k2/k2′ = 2.03 ± 0.47: CH3O + NO → CH3ONO (reaction 2) and CH3O + NO2 → CH3ONO2 (reaction 2′). Deviations from this value in this and previous work is ascribed to the pressure dependence of both these reactions and heterogeneity in reaction (2). In contrast no heterogeneous effects were found for reaction (2′) making it an ideal reference reaction for studying other reactions of the methoxy radical. We conclude that the ratio k2/k2′ is independent of temperature and from k1 = 1010.2±0.4M?1 sec?1 we calculate that k2′ = 109.9±0.4M?1 sec?1. Both k2 and k2′ are pressure dependent but have reached their limiting high-pressure values in the presence of 500 Torr of carbon tetrafluoride. Preliminary results show that k4 = 10.9.0±0.6 10?4.5±1.1M?1 sec?1 (Θ = 2.303RT kcal mole?1) and by k4 = 108.6±0.6 10?2.4±1.1M?1 sec?1: CH3O + O2 → CH2O + HO2 (reaction 4) and CH3O + t-BuH → CH3OH + (t-Bu) (reaction 4′).  相似文献   

15.
The kinetics of the thermal decomposition of acetylacetone has been studied in a shock tube in the temperature range of 1120–1660 K. Detailed analyses of CO and H2O formation data indicate that H2O is formed by a four-center molecular channel, whereas CO is formed by the rapid dissociation of CH3CO produced by the C? C bond dissociation of acetylacetone. The Arrhenius equations for H2O and CH3CO formation channels are ??2 = 1014.24±0.21 exp(?60,800 ± 1,220/RT)sec?1 and ??3 = 1017.05±0.28 exp(?74,600 ± 1,680/RT) sec?1, respectively. The results of the study suggest that the six-center molecular channel for the production of acetone and ketene is not important under the condition used in this investigation.  相似文献   

16.
17.
Four different types of polydepsipeptide‐polyether block copolymers were synthesized via ring‐opening polymerization of 3(S)‐sec‐butylmorpholine‐2,5‐dione (BMD) in the presence of hydroxytelechelic poly(ethylene oxide) (PEO) with stannous octoate as a catalyst.The polymers were an AB block copolymer, an ABA block copolymer, an (A)2B star shaped copolymer and an (A)2B(A)2 copolymer, where A is a poly[3(S)‐sec‐butylmorpholine‐2,5‐dione] (PBMD) and B a poly(ethylene oxide) block. The molar ratio of BMD to PEO was varied to obtain copolymers with different weight fractions of PBMD blocks ranging from 59.8 to 96.7 wt.‐%. The crystallinity of the PEO phase in the copolymers decreases in the following order: AB > (A)2B > ABA > (A)2B(A)2 . The static contact angle θ decreases with increasing PEO content in the block copolymers.  相似文献   

18.
The pyrolysis of n-propyl nitrate and tert-butyl nitrite at very low pressures (VLPP technique) is reported. For the reaction the high-pressure rate expression at 300°K, log k1 (sec?1) = 16.5 ? 40.0 kcal/mole/2.3 RT, is derived. The reaction was studied and the high-pressure parameters at 300°K are log k2(sec?1) = 15.8 ? 39.3 kcal/mole/2.3 RT. From ΔS1,?10 and ΔS2,?20 and the assumption E?1 and E?2 ? 0, we derive log k?1(M?1·sec?1) (300°K) = 9.5 and log k?2 (M?1·sec?1) (300°K) = 9.8. In contrast, the pyrolysis of methyl nitrite and methyl d3 nitrite afford NO and HNO and DNO, respectively, in what appears to be a heterogeneous process. The values of k?1 and k?2 in conjunction with independent measurements imply a value at 300°K for of 3.5 × 105 M?1·sec?1, which is two orders of magnitude greater than currently accepted values. In the high-pressure static pyrolysis of dimethyl peroxide in the presence of NO2, the yield of methyl nitrate indicates that the combination of methoxy radicals with NO2 is in the high-pressure limit at atmospheric pressure.  相似文献   

19.
The processing of silica in an rf plasma was studied with a new experimental system, the process being carried out entirely in the vapor phase. SiO and Si were obtained by either thermal decomposition of SiO2 or reduction by H2. No other reducing material (which, like C, may become a contaminant) was used. The products were collected along the reactor length and analyzed afterward. Temperatures in the reaction zone were measured spectroscopically. The high reaction rates occurring in the gas phase did compensate for the low residence times. A diffusional model for the buildup of solid reaction products is discussed. Vapor phase plasma reduction of silica with hydrogen proved to be more effective than thermal decomposition and nuch more effective than heterogeneous processing. The efficiency evaluation was made on the laboratory-scale level.Nomenclature c SiO d SiO content, % by wt - c Si t Total Si content, % by wt - cSiO 2 e Experimental SiO2 content, % by wt - cSiO 2 t Theoretical SiO2 content, % by wt - D Species diffusivity, cm2 · sec–1 - F Rate of species deposition, mol · cm–2 · sec–1 - MSiO2 Molar flow of SiO2, mmol · sec–1 - n Species concentration, cm–3 - P Anode power, kW - r Reaction radius, cm - R * Radius of quartz tube, cm - T Temperature, K - T av Average temperature, K - USiO 2 t Total extent of SiO2 decomposition, % - V Ar a Molar flow of axial Ar, mmol · sec–1 - V Ar p Molar flow of peripheral Ar, mmol · sec–1 - VH2 Molar flow of H2, mmol · sec–1 - y Molar fraction - Z Position of a plane normal to the plasma axis down the lower turn of rf coil, cm Greek Letters Molar ratio of SiO over SiO2 in deposit  相似文献   

20.
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