首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 734 毫秒
1.
将一种改进的目标因子分析法(MTFA)用于复杂反应的速差动力学分析体系,采用此法处理实验数据时,引入了一个非零截距因子,采用模拟纯谱运算。可使方法的准确度大大提高。据此建立了一种停流技术进样、速差动力学分析同时测定、因子分析处理实验数据的新方法;对于10-6mol·L-1级的肾上腺素-去甲基肾上腺素混合物,比例范围在1:5~5:1之间,测定相对误差小于±10%,用于合成样品中肾上腺素和去甲基肾上腺素的同时测定,获得了满意结果。  相似文献   

2.
连串反应速差动力学分析中的数据处理方法的研究吴新国,蔡汝秀,林智信,程介克(武汉大学化学系,武汉,430072)关键词连串反应,速差动力学分析,比例方程法,肾上腺素,去甲肾上腺素动力学分析在分析化学中的作用和重要性已广为人知[1],速差动力学分析是其...  相似文献   

3.
去甲肾上腺素电极过程的圆二色谱电化学研究   总被引:3,自引:0,他引:3  
现场圆二色薄层光谱电化学研究去甲肾上腺素的电化学氧化还原过程 .研究表明去甲肾上腺素 ( pH =7.0磷酸缓冲溶液中 )在玻碳电极上经历了不可逆的电化学氧化 ,且遵从后行化学反应 (EC)机理 ,去甲肾上腺素醌和去甲肾上腺素红的再还原遵从简单电子转移 (E)机理 .由双对数法获得去甲肾上腺素电化学氧化的式电位为E10’=0 .2 0V ,电子转移系数和电子转移数之积为αn =0 .38,标准复相电极反应常数k10 =1 .2× 1 0 -4 cm·s-1.去甲肾上腺素醌和去甲肾上腺素红的电化学还原反应参数分别为E2 0’=0 .2 5V ,αn =0 .37,k2 0 =4.4× 1 0 -5 cm·s-1和E3 0’=- 0 .2 5V ,αn =0 .33,k3 0 =1 .1× 1 0 -4 cm·s-1.  相似文献   

4.
循环伏安法测定去甲肾上腺素   总被引:1,自引:0,他引:1  
制备了聚L-半胱氨酸修饰电极,研究了去甲肾上腺素在聚合物薄膜上的电化学行为,试验结果表明:在磷酸盐缓冲溶液中,聚L-半胱氨酸薄膜对去甲肾上腺素的电化学氧化具有明显的催化作用,对应用此修饰电极的循环伏安法测定去甲肾上腺素的条件进行试验并优化.结果发现:在pH 7.0的磷酸盐缓冲介质中测定时可排除肾上腺素与抗坏血酸干扰.还原峰电流的测定值与去甲肾上腺素浓度在5.0×10-7~1.2×10-4mol·L-1范围内呈线性关系,其检出限为2.0×10-8mol·L-1,已用于针剂样品分析,测得回收率在97.6%~102.1%之间.  相似文献   

5.
吕元琦  邬春华  袁倬斌 《分析化学》2004,32(8):1016-1018
制备了聚 2 ,3 吡啶二羧酸修饰玻碳电极 ,研究了去甲肾上腺素在聚合物薄膜修饰电极上的电化学行为 ,实验结果表明 :在 pH 7.4的 0 .1mol/LPBS底液中 ,聚 2 ,3 吡啶二羧酸薄膜对去甲肾上腺素的电化学氧化具有明显的催化作用 ,并可以清除去甲肾上腺素测定中的抗坏血酸干扰。去甲肾上腺素检测的线性范围是 2 .0× 1 0 -6~ 5 .0× 1 0 -5mol/L ;检出限为 7.2× 1 0 -8mol/L ,用于分析针剂样品的结果和标示值一致  相似文献   

6.
研究发现富含多酚氧化酶的蘑菇组织可催化溶解氧氧化去甲肾上腺素和左旋多巴在碱性条件下生成具有强荧光的三羟基和二羟基吲哚类物质 ,建立了以蘑菇组织柱为酶反应器的去甲肾上腺素和左旋多巴的流通式荧光分析法 .该法对去甲肾上腺素和左旋多巴响应的线性范围分别为 6× 10 -8~ 1× 10 -5g·mL-1和 3× 10 -8~ 1× 10 -5g·mL-1,检测限分别为 2×10 -8g·mL-1和 1× 10 -8g·mL-1( 3S/k) .该组织反应器可连续使用 14d ,对 1× 10 -7g·mL-1去甲肾上腺素和 1× 10 -7g·mL-1左旋多巴测定的相对标准偏差 (R .S .D .)均小于 3 % (n =11) .详细研究了常见离子和抗氧化剂对本体系的干扰情况 .将该体系用于药物制剂中去甲肾上腺素和左旋多巴含量的测定 ,结果与药典标准方法测得值一致 .实验结果证明了方法的可行性和可靠性  相似文献   

7.
基于肾上腺素对血红蛋白酶催化体系的抑制作用, 建立了酶催化动力学光度法测定肾上腺素的新方法. 实验研究了体系的最佳条件及动力学行为, 测定的线性范围为4.5×10-7~1.4×10-5 mol/L, 方法检出限为5.2×10-8 mol/L. 对浓度为9.0×10-6 mol/L的肾上腺素进行11次平行测定的相对标准偏差为3.5%. 此方法可用于药剂中肾上腺素含量的测定.  相似文献   

8.
用循环伏安法制备了聚L-丝氨酸修饰玻碳电极,研究了去甲肾上腺素在聚L-丝氨酸修饰玻碳电极上的电化学行为,建立了测定去甲肾上腺素的新方法。实验结果表明:在pH 6.0的磷酸盐缓冲溶液中,聚L-丝氨酸薄膜对去甲肾上腺素的电化学氧化具有明显的催化作用,利用循环伏安法测定其还原峰电流可排除抗坏血酸干扰。去甲肾上腺素检测线性范围为4.0×10-7~1.5×10-4mol/L;检出限为1.0×10-8mol/L。该修饰电极具有良好的灵敏度、选择性和稳定性,已用于针剂样品分析。  相似文献   

9.
张宏  金葆康 《分析化学》2002,30(11):1285-1288
利用电沉积的方法制得纳米金修饰玻碳电极 ,该修饰电极对去甲肾上腺素 (NE)氧化反应有催化作用。去甲肾上腺素在纳米金修饰玻碳电极上有很强的吸附作用。研究了磷酸缓冲溶液的pH值和浓度对NE的电化学行为的影响。从去甲肾上腺素和抗坏血酸在纳米金修饰电极的循环伏安图上可观察到两个明显分开的氧化峰 ,峰电位差达到 1 3 1mV ,因此 ,可利用该修饰电极在抗坏血酸存在下选择性测定去甲肾上腺素 ,线性范围为 1× 1 0 - 4 ~5× 1 0 - 6 mol L。  相似文献   

10.
试验表明:在pH 9.5氨水-氯化铵缓冲介质中,牛血红蛋白对过氧化氢与酸性络蓝K的氧化还原反应的催化作用因异丙肾上腺素的存在而被抑制,而且其抑制率与异丙肾上腺素的浓度在8.1×10~(-8)~1.2×10~(-5)mol·L~(-1)范围内呈线性关系,其检出限(3S/N)为2.2×10~(-9)mol·L~(-1)。据此,提出了一种简单而灵敏的测定异丙肾上腺素的酶催化光度法。此方法已用于盐酸异丙肾上腺素注射液中异丙肾上腺素含量的测定,并以此样品为基体用标准加入法作回收试验,测得回收率在92.8%~105.4%之间,测定值的相对标准偏差(n=5)在2.6%~5.8%之间。  相似文献   

11.
We calculate, down to low temperature and for different isotopes, the reaction rate constants for the hydrogen abstraction reaction H + H(3)COH → H(2) + CH(2)OH/CH(3)O. These explain the known abundances of deuterated forms of methanol in interstellar clouds, where CH(2)DOH can be almost as abundant as CH(3)OH. For abstraction from both the C- and the O-end of methanol, the barrier-crossing motion involves the movement of light hydrogen atoms. Consequently, tunneling plays a dominant role already at relatively high temperature. Our implementation of harmonic quantum transition state theory with on the fly calculation of forces and energies accounts for these tunneling effects. The results are in good agreement with previous semiclassical and quantum dynamics calculations (down to 200 K) and experimental studies (down to 295 K). Here we extend the rate calculations down to lower temperature: 30 K for abstraction from the C-end of methanol and 80 K for abstraction from the OH-group. At all temperatures, abstraction from the C-end is preferred over abstraction from the O-end, more strongly so at lower temperature. Furthermore, the tunneling behavior strongly affects the kinetic isotope effects (KIEs). D + H(3)COH → HD + CH(2)OH has a lower vibrationally adiabatic barrier than H + H(3)COH → H(2) + CH(2)OH, giving rise to an inverse KIE (k(H)/k(D) < 1) at high temperature, in accordance with previous experiments and calculations. However, since tunneling is more facile for the light H atom, abstraction by H is favored over abstraction by D below ~135 K, with a KIE k(H)/k(D) of 11.2 at 30 K. The H + D(3)COD → HD + CD(2)OD reaction is calculated to be much slower than the D + H(3)COH → HD + CH(2)OH, in agreement with low-temperature solid-state experiments, which suggests the preference for H (as opposed to D) abstraction from the C-end of methanol to be the mechanism by which interstellar methanol is deuterium-enriched.  相似文献   

12.
The beta-diketone Hamac = 3-(N-acetylamido)pentane-2,4-dione was characterized by potentiometric, spectrophotometric, and kinetic methods. In water, Hamac is very soluble (2.45 M) and strongly enolized, with [enol]/[ketone] = 2.4 +/- 0.1. The pK(a) of Hamac is 7.01 +/- 0.07, and the rate constants for enolization, k(e), and ketonization, k(k), at 298 K are 0.0172 +/- 0.0004 s(-1) and 0.0074 +/- 0.0015 s(-1), respectively. An X-ray structure analysis of the copper(II) complex Cu(amac)(2).toluene (=C(21)H(28)CuN(2)O(6); monoclinic, C2/c; a = 20.434(6), b = 11.674(4), c = 19.278(6) ?; beta = 100.75(1) degrees; Z = 8; R(w) = 0.0596) was carried out. The bidentate anions amac(-) coordinate the copper via the two diketo oxygen atoms to form a slightly distorted planar CuO(4) coordination core. Rapid-scan stopped-flow spectrophotometry was used to study the kinetics of the reaction of divalent metal ions M(2+) (M = Ni,Co,Cu) with Hamac in buffered aqueous solution at variable pH and I = 0.5 M (NaClO(4)) under pseudo-first-order conditions ([M(2+)](0) > [Hamac](0)) to form the mono complex M(amac)(+). For all three metals the reaction is biphasic. The absorbance/time data can be fitted to the sum of two exponentials, which leads to first-order rate constants k(f) (fast initial step) and k(s) (slower second step). The temperature dependence of k(f) and k(s) was measured. It follows from the kinetic data that (i) the keto tautomer of Hamac, HK, does not react with the metal ions M(2+), (ii) the rate constant k(f) increases linearly with [M(2+)](0) according to k(f) = k(0) + k(2)[M(2+)](0), and (iii) the rate constant k(s) does not depend on [M(2+)](0) and describes the enolization of the unreactive keto tautomer HK. The pH dependence of the second-order rate constant k(2) reveals that both the enol tautomer of Hamac, HE, and the enolate, E(-), react with M(2+) in a second-order reaction to form the species M(amac)(+). At 298 K rate constants k(HE) are 18 +/- 6 (Ni), 180 +/- 350 (Co), and (9 +/- 5) x 10(4) (Cu) M(-1) s(-1) and rate constants k(E) are 924 +/- 6 (Ni), (7.4 +/- 0.6) x 10(4) (Co), and (8.4 +/- 0.2) x 10(8) (Cu) M(-1) s(-1). The acid dissociation of the species M(amac)(+) is triphasic. Very rapid protonation (first step) leads to M(Hamac)(2+), which is followed by dissociation of M(Hamac)(2+) and M(amac)(+), respectively (second step). The liberated enol Hamac ketonizes (third step). The mechanistic implications of the metal dependence of rate constants k(HE), k(E), k(-HE), and k(-E) are discussed.  相似文献   

13.
M Carmona  M Silva  D Pérez-Bendito 《The Analyst》1991,116(10):1075-1079
A stopped-flow method for the simultaneous determination of epinephrine and norepinephrine, two catecholamines with a wide spectrum of biological activity, is proposed. The method is based on the oxidation of these compounds with 1,10-phenanthroline-iron(III) complex, which is monitored by measuring the initial rate of change of the absorbance of the ferroin formed at 510 nm. The difference in kinetic behaviour between the two species was exploited by applying a modified version of the proportional-equation method for the resolution of epinephrine-norepinephrine mixtures at the micrograms ml-1 level over the ratio range 1:10-10:1, with an error of less than 5% and an average precision of about 2.5%. The method was successfully applied to the determination of the two catecholamines in pharmaceuticals.  相似文献   

14.
The kinetics for the gas-phase reaction of phenyl radical with propyne has been measured by cavity ring-down spectrometry (CRDS), and the mechanism and initial product branching have been elucidated with the help of quantum chemical calculations. Absolute rate constants measured by the CRDS technique can be expressed by the following Arrhenius equation: (k/cm(3) mol(-1) s(-1)): k(propyne)(T=301-428 K)=(3.68+/-0.92) x 10(11)exp[-(1685+/-80)/T]. The experiment is unable to distinguish between the possible reactive channels, but theory indicates that phenyl radicals preferably add to the unsaturated terminal carbon atom in propyne under our experimental conditions. Theoretical kinetic calculations, employing high-level G2M(RCC, RMP2) and G3 energetic and IRCMax(RCCSD(T)//B3LYP-DFT) molecular parameters, reproduce the total experimental rate constants within a factor of three. Calculated total and branching rate constants are provided for high-T kinetic modeling. Addition reactions of phenyl to C3H4 are estimated to be less important molecular-growth pathways in high-T conditions (T>1000 K) in comparison to the C6H5 + C2H2 reaction.  相似文献   

15.
A simultaneous kinetic method based on the measurement of two rates at two points during the course of the successive reactions is proposed. The successive kinetics of the reaction of tris(1,10-phenanthroline)iron(III) with epinephrine (EP) and norepinephrine (NE) (10−5 to 10−4 M) is investigated and exploited for the determination of binary mixtures of the two catecholamines using the two-rate method. The method is compared with classical and extension proportional equation methods and a smaller error is obtained for the two-rate method. The reason why the two-rate method based on successive reactions works well is explained by studies on a series of simulated kinetic curves.  相似文献   

16.
A kinetic study of the reactions of ground state V, Fe, and Co with SO2 is reported. V, Fe, and Co were produced by the 248 nm photodissociation of VCl4, ferrocene, and Co(C5H5)(CO)2, respectively, and were detected by laser-induced fluorescence. V + SO2 proceeds by an abstraction reaction with rate constants given by k=(2.33 +/- 0.57)x 10(-10) exp[-(1.14 +/- 0.19) kcal mol(-1)/RT] cm3 molecule(-1) s(-1) over the temperature range 296-571 K. Fe + SO2 was studied in the N2 buffer range of 10-185 Torr between 294 and 498 K. The limiting, low-pressure third-order rate constants are given by k(0)=(3.45 +/- 1.19)x 10(-30) exp[-(2.81 +/- 0.24) kcal mol(-1)/RT] cm6 molecule(-2) s(-1). Co + SO2 was studied in the CO2 buffer range of 5-40 Torr between 294 and 498 K. This reaction is independent of temperature over the indicated range and has a third-order rate constant of k0=(5.23 +/- 0.28)x 10(-31) cm6 molecule(-2) s(-1). Results of this work are compared to previous work on the Sc, Ti, Cr, Mn, and Ni + SO2 systems. The reaction efficiencies for the abstraction reactions depend on the ionization energies of the transition metal atoms and on the reaction exothermicities, and the reaction efficiencies of the association reactions are strongly dependent on the energies needed to promote an electron from a 4s2 configuration to a 4s1 configuration.  相似文献   

17.
The reaction of AlMe(3) with (t-Bu(3)PN)(2)TiMe(2) 1 proceeds via competitive reactions of metathesis and C-H activation leading ultimately to two Ti complexes: [(mu(2)-t-Bu(3)PN)Ti(mu-Me)(mu(4)-C)(AlMe(2))(2)](2) 2, [(t-Bu(3)PN)Ti(mu(2)-t-Bu(3)PN)(mu(3)-CH(2))(2)(AlMe(2))(2)(AlMe(3))] 3, and the byproduct (Me(2)Al)(2)(mu-CH(3))(mu-NP(t-Bu(3))) 4. X-ray structural data for 2 and 3 are reported. Compound 3 undergoes thermolysis to generate a new species [Ti(mu(2)-t-Bu(3)PN)(2)(mu(3)-CH(2))(mu(3)-CH)(AlMe(2))(3)] 5. Monitoring of the reaction of 1 with AlMe(3) by (31)P[(1)H] NMR spectroscopy revealed intermediates including (t-Bu(3)PN)TiMe(3) 6. Compound 6 was shown to react with AlMe(3) to give 2 exclusively. Kinetic studies revealed that the sequence of reactions from 6 to 2 involves an initial C-H activation that is a second-order reaction, dependent on the concentration of Ti and Al. The second-order rate constant k(1) was 3.9(5) x 10(-4) M(-1) s(-1) (DeltaH(#) = 63(2) kJ/mol, DeltaS(#) = -80(6) J/mol x K). The rate constants for the subsequent C-H activations leading to 2 were determined to be k(2) = 1.4(2) x 10(-3) s(-1) and k(3) = 7(1) x 10(-3) s(-1). Returning to the more complex reaction of 1, the rate constant for the ligand metathesis affording 4 and 6 was k(met) = 6.1(5) x 10(-5) s(-1) (DeltaH(#) = 37(3) kJ/mol, DeltaS(#) = -203(9) J/mol x K). The concurrent reaction of 1 leading to 3 was found to proceed with a rate constant of k(obs) of 6(1) x 10(-5) s(-1) (DeltaH(#) = 62(5) kJ/mol, DeltaS(#)= -118(17) J/mol x K). Using these kinetic data for these reactions, a stochastic kinetic model was used to compute the concentration profiles of the products and several intermediates with time for reactions using between 10 and 27 equivalents of AlMe(3). These models support the view that equilibrium between 1 x AlMe(3) and 1 x (AlMe(3))(2) accounts for varying product ratios with the concentration of AlMe(3). In a similar vein, similar equilibria account for the transient concentrations of 6 and an intermediate en route to 3. The implications of these reactions and kinetic and thermodynamic data for both C-H bond activation and deactivation pathways for Ti-phosphinimide olefin polymerization catalysts are considered and discussed.  相似文献   

18.
The high-temperature rate constants of the reactions NCN + NO and NCN + NO(2) have been directly measured behind shock waves under pseudo-first-order conditions. NCN has been generated by the pyrolysis of cyanogen azide (NCN(3)) and quantitatively detected by sensitive difference amplification laser absorption spectroscopy at a wavelength of 329.1302 nm. The NCN(3) decomposition initially yields electronically excited (1)NCN radicals, which are subsequently transformed to the triplet ground state by collision-induced intersystem crossing (CIISC). CIISC efficiencies were found to increase in the order of Ar < NO(2) < NO as the collision gases. The rate constants of the NCN + NO/NO(2) reactions can be expressed as k(NCN+NO)/(cm(3) mol(-1)s(-1)) = 1.9 × 10(12) exp[-26.3 (kJ/mol)/RT] (±7%,ΔE(a) = ± 1.6 kJ/mol, 764 K < T < 1944 K) and k(NCN+NO(2))/(cm(3) mol(-1)s(-1)) = 4.7 × 10(12) exp[-38.0(kJ/mol)/RT] (±19%,ΔE(a) = ± 3.8 kJ/mol, 704 K < T < 1659 K). In striking contrast to reported low-temperature measurements, which are dominated by recombination processes, both reaction rates show a positive temperature dependence and are independent of the total density (1.7 × 10(-6) mol/cm(3) < ρ < 7.6 × 10(-6) mol/cm(3)). For both reactions, the minima of the total rate constants occur at temperatures below 700 K, showing that, at combustion-relevant temperatures, the overall reactions are dominated by direct or indirect abstraction pathways according to NCN + NO → CN + N(2)O and NCN + NO(2) → NCNO + NO.  相似文献   

19.
The thermal decomposition of propane has been studied using both shock tube experiments and ab initio transition state theory-based master equation calculations. Dissociation rate constants for propane have been measured at high temperatures behind reflected shock waves using high-sensitivity H-ARAS detection and CH(3) optical absorption. The two major dissociation channels at high temperature are C(3)H(8) → CH(3) + C(2)H(5) (eq 1a) and C(3)H(8) → CH(4) + C(2)H(4) (eq 1b). Ultra high-sensitivity ARAS detection of H-atoms produced from the decomposition of the product, C(2)H(5), in (1a), allowed measurements of both the total decomposition rate constants, k(total), and the branching to radical products, k(1a)/k(total). Theoretical analyses indicate that the molecular products are formed exclusively through the roaming radical mechanism and that radical products are formed exclusively through channel 1a. The experiments were performed over the temperature range 1417-1819 K and gave a minor contribution of (10 ± 8%) due to roaming. A multipass CH(3) absorption diagnostic using a Zn resonance lamp was also developed and characterized in this work using the thermal decomposition of CH(3)I as a reference reaction. The measured rate constants for CH(3)I decomposition agreed with earlier determinations from this laboratory that were based on I-atom ARAS measurements. This CH(3) diagnostic was then used to detect radicals from channel 1a allowing lower temperature (1202-1543 K) measurements of k(1a) to be determined. Variable reaction coordinate-transition state theory was used to predict the high pressure limits for channel (1a) and other bond fission reactions in C(3)H(8). Conventional transition state theory calculations were also used to estimate rate constants for other tight transition state processes. These calculations predict a negligible contribution (<1%) from all other bond fission and tight transition state processes, indicating that the bond fission channel (1a) and the roaming channel (1b) are indeed the only active channels at the temperature and pressure ranges of the present experiments. The predicted reaction exo- and endothermicities are in excellent agreement with the current version of the Active Thermochemical Tables. Master equation calculations incorporating these transition state theory results yield predictions for the temperature and pressure dependence of the dissociation rate constants for channel 1a. The final theoretical results reliably reproduce the measured dissociation rate constants that are reported here and in the literature. The experimental data are well reproduced over the 500-2500 K and 1 × 10(-4) to 100 bar range (errors of ~15% or less) by the following Troe parameters for Ar as the bath gas: k(∞) = 1.55 × 10(24)T(-2.034) exp(-45?490/T) s(-1), k(0) = 7.92 × 10(53)T(-16.67) exp(-50?380/T) cm(3) s(-1), and F(c) = 0.190 exp(-T/3091) + 0.810 exp(-T/128) + exp(-8829/T).  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