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.
Mitogenic stimulation of mouse lymphocytes by the enzyme galactose oxidase (GO) is inhibited if Con A is present during the enzymatic oxidation. The mechanism of this inhibition appears to involve steric hindrance of GO action at cell surface sites which bind Con A because (a) similar pulse exposure of unoxidized cells to Con A does not affect their subsequent ability to respond to GO stimulation; (b) Con A binding to fetuin interferes with GO oxidation of that glycoprotein substrate; and (c) sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis of cells labeled by GONaB3H4 in the presence of Con A shows a selective inhibition of labeling of some high-molecular-weight glycoproteins compared to controls labeled in the absence of Con A.  相似文献   

2.
An adaptation of the sodium periodate/sodium borotritide procedure for the identification of membrane sialoglycoproteins is described which eliminates interference from nonspecifically incorporated tritium. Synaptic membranes were labeled using the NaIO4NaB3H4 procedure and separated by polyacrylamide gel electrophoresis. Following electrophoresis the gels were fixed, sliced, and individual slices treated with neuraminidase. Treatment with neuraminidase selectively released [3H]sialyl derivatives from the fixed glycoproteins allowing the unambiguous identification of sialoglycoproteins. The sialoglycoprotein composition of synaptic membranes and synaptic junctions was compared.  相似文献   

3.
N-Phenylhydroxylamine is oxidized in aqueous phosphate buffer to nitrosobenzene, nitrobenzene, and azoxybenzene. Degradation is O2 dependent and shows general catalysis by H2PO4? (k1 = 2.3 M?2 sec?1) and PO4?3 (k2 = 2.3 × 105M?2 sec?1) or kinetically equivalent terms. Evidence is presented suggesting the intermediacy of a highly reactive species leading to these products.  相似文献   

4.
ADP and Pi-loaded membrane vesicles from l-malate-grown Bacillus alcalophilus synthesized ATP upon energization with ascorbateN,N,N′,N′-tetramethyl-p-phenylenediamine. ATP synthesis occurred over a range of external pH from 6.0 to 11.0, under conditions in which the total protonmotive force Δ\?gmH+ was as low as ?30 mV. The phosphate potentials (ΔGp) were calculated to be 11 and 12 kcal/mol at pH 10.5 and 9.0, respectively, whereas the Δ\?gmH+ values in vesicles at these two pH values were quite different (?40 ± 20 mV at pH 10.5 and ?125 ± 20 mV at pH 9.0). ATP synthesis was inhibited by KCN, gramicidin, and by N,N′-dicyclohexylcarbodiimide. Inward translocation of protons, concomitant with ATP synthesis, was demonstrated using direct pH monitoring and fluorescence methods. No dependence upon the presence of Na+ or K+ was found. Thus, ATP synthesis in B. alcalophilus appears to involve a proton-translocating ATPase which functions at low Δ\?gmH+.  相似文献   

5.
(1) H+/electron acceptor ratios have been determined with the oxidant pulse method for cells of denitrifying Paracoccus denitrificans oxidizing endogenous substrates during reduction of O2, NO?2 or N2O. Under optimal H+-translocation conditions, the ratios H+O, H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were 6.0–6.3, 4.02, 5.79 and 3.37, respectively. (2) With ascorbate/N,N,N′,N′-tetramethyl-p-phenylenediamine as exogenous substrate, addition of NO?2 or N2O to an anaerobic cell suspension resulted in rapid alkalinization of the outer bulk medium. H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were ?0.84, ?2.33 and ?1.90, respectively. (3) The H+oxidant ratios, mentioned in item 2, were not altered in the presence of valinomycinK+ and the triphenylmethylphosphonium cation. (4) A simplified scheme of electron transport to O2, NO?2 and N2O is presented which shows a periplasmic orientation of the nitrite reductase as well as the nitrous oxide reductase. Electrons destined for NO?2, N2O or O2 pass two H+-translocating sites. The H+electron acceptor ratios predicted by this scheme are in good agreement with the experimental values.  相似文献   

6.
The observed equilibrium constants (Kobs) for the l-phosphoserine phosphatase reaction [EC 3.1.3.3] have been determined under physiological conditions of temperature (38 °C) and ionic strength (0.25 m) and physiological ranges of pH and free [Mg2+]. Using Σ and square brackets to indicate total concentrations Kobs = Σ L-serine][Σ Pi]Σ L-phosphoserine]H2O], K = L-H · serine±]HPO42?][L-H · phosphoserine2?]H2O]. The value of Kobs has been found to be relatively sensitive to pH. At 38 °C, K+] = 0.2 m and free [Mg2+] = 0; Kobs = 80.6 m at pH 6.5, 52.7 m at pH 7.0 [ΔGobs0 = ?10.2 kJ/mol (?2.45 kcal/mol)], and 44.0 m at pH 8.0 ([H2O] = 1). The effect of the free [Mg2+] on Kobs was relatively slight; at pH 7.0 ([K+] = 0.2 m) Kobs = 52.0 m at free [Mg2+] = 10?3, m and 47.8 m at free [Mg2+] = 10?2, m. Kobs was insignificantly affected by variations in ionic strength (0.12–1.0 m) or temperature (4–43 °C) at pH 7.0. The value of K at 38 °C and I = 0.25 m has been calculated to be 34.2 ± 0.5 m [ΔGobs0 = ?9.12 kJ/mol (?2.18 kcal/ mol)]([H2O] = 1). The K for the phosphoserine phosphatase reaction has been combined with the K for the reaction of inorganic pyrophosphatase [EC 3.6.1.1] previously estimated under the same physiological conditions to calculate a value of 2.04 × 104, m [ΔGobs0 = ?28.0 kJ/mol (?6.69 kcal/mol)] for the K of the pyrophosphate:l-serine phosphotransferase [EC 2.7.1.80] reaction. Kobs = [Σ L-serine][Σ Pi][Σ L-phosphoserine][H2O], K = [L-H · serine±]HPO42?][L-H · phosphoserine2?]H2O. Values of Kobs for this reaction at 38 °C, pH 7.0, and I = 0.25 m are very sensitive to the free [Mg2+], being calculated to be 668 [ΔGobs0 = ?16.8 kJ/mol (?4.02 kcal/mol)] at free [Mg2+] = 0; 111 [ΔGobs0 = ?12.2 kJ/mol (?2.91 kcal/mol)] at free [Mg2+] = 10?3, m; and 9.1 [ΔGobs0 = ?5.7 kJ/mol (?1.4 kcal/mol) at free [Mg2+] = 10?2, m). Kobs for this reaction is also sensitive to pH. At pH 8.0 the corresponding values of Kobs are 4000 [ΔGobs0 = ?21.4 kJ/mol (?5.12 kcal/mol)] at free [Mg2+] = 0; and 97.4 [ΔGobs0 = ?11.8 kJ/ mol (?2.83 kcal/mol)] at free [Mg2+] = 10?3, m. Combining Kobs for the l-phosphoserine phosphatase reaction with Kobs for the reactions of d-3-phosphoglycerate dehydrogenase [EC 1.1.1.95] and l-phosphoserine aminotransferase [EC 2.6.1.52] previously determined under the same physiological conditions has allowed the calculation of Kobs for the overall biosynthesis of l-serine from d-3-phosphoglycerate. Kobs = [Σ L-serine][Σ NADH][Σ Pi][Σ α-ketoglutarate][Σ d-3-phosphoglycerate][Σ NAD+][Σ L-glutamat0] The value of Kobs for these combined reactions at 38 °C, pH 7.0, and I = 0.25 m (K+ as the monovalent cation) is 1.34 × 10?2, m at free [Mg2+] = 0 and 1.27 × 10?2, m at free [Mg2+] = 10?3, m.  相似文献   

7.
The relative effectiveness of oxidizing (.OH, H2O2), ambivalent (O2?) and reducing free radicals (e? and CO2?) in causing damage to membranes and membrane-bound glyceraldehyde-3-phosphate dehydrogenase of resealed erythrocyte ghosts has been determined. The rates of damage to membranebound glyceraldehyde-3-phosphate dehydrogenase (R(enz)) were measured and the rates of damage to membranes (R(mb)) were assessed by measuring changes in permeability of the resealed ghosts to the relatively low molecular weight substrates of glyceraldehyde-3-phosphate dehydrogenase. Each radical was selectively isolated from the mixture produced during gamma-irradiation, using appropriate mixtures of scavengers such as catalase, superoxide dismutase and formate. .OH, O2? and H2 O2 were approximately equally effective in inactivating membrane-bound glyceraldehyde-3-phosphate dehydrogenase, while e? and CO2? were the least effective. R(enz) values of O2? and H2O2 were 10-times and of .OH 15-times that of e?. R(mb) values were quite similar for e? and H2O2 (about twice that of O2?), while that of .OH was 3-times that of O2?. Hence, with respect to R(mb): .OH >e? = H2O2 >O2? , and with respect to R(enz): .OH >O2? = H2O2 >e?. The difference between the effectiveness of the most damaging and the least damaging free radicals was more than 10-fold greater in damage to the enzyme than to the membranes. Comparison between H2O2 added as a chemical reagent and H2O2 formed by irradiation showed that membranes and membrane-bound glyceraldehyde-3-phosphate dehydrogenase were relatively inert to reagent H2O2 but markedly susceptible to the latter.  相似文献   

8.
The effects of absolute temperature (T), ionic strength (μ), and pH on the polymerization of tobacco mosaic virus protein from the 4 S form (A) to the 20 S form (D) were investigated by the method of sedimentation velocity. The loading concentration in grams per liter (C) was determined at which a just-detectable concentration (β) of 20 S material appeared. It was demonstrated experimentally that under the conditions employed herein, an equilibrium concentration of 20 S material was achieved in 3 h at the temperature of the experiment and that 20 S material dissociated again in 4 h or less to 4 S material either upon lowering the temperature or upon dilution. Thus, the use of thermodynamic equations for equilibrium processes was shown to be valid. The equation used to interpret the results, log (C?β) = constant + (ΔH12.3RT) + (ΔW1el2.3RT) ? K′ + ζpH, was derived from three separate models of the process, the only difference being in the anatomy of the constant; thus, the method of analysis is essentially independent of the model. ΔH1 and ΔW1el are the enthalpy and the change in electrical work per mole of A protein (the trimer of the polypeptide chain), Ks is the salting-out constant on the ionic strength basis, ζ is the number of moles of hydrogen ion bound per mole of A protein in the polymerization, and R is the gas constant. The three models leading to this equation are: a simple 11th-order equilibrium between A1 (the trimer of the polypeptide chain) and D, either the double disk or the double spiral of approximately the same molecular weight, designated model A; a second model, designated B, in which A1 was assumed to be in equilibrium with D at the same time that it is in equilibrium with A2, A3, etc., dimers and trimers, etc., of A1 in an isodesmic system; and a phase-separation model, designated model C, in which A protein is treated as a soluble material in equilibrium with D, considered as an insoluble phase. From electrical work theory, ΔWel1/T was shown to be essentially independent of T; therefore, in experiments at constant μ and constant pH the equation of log (C ? β) versus 1/T is linear with a slope of ΔH1/2.3R. The results fit such an equation over nearly a 20 °C-temperature range with a single value of ΔH1 of +32 kcal/mol A1. Results obtained when T and pH were held constant but μ was varied did not fit a straight line, which shows that more than simple salting-out is involved. When the effect of ionic strength on the electrical work contribution was considered in addition to salting-out, the data were interpreted to indicate a value of ΔW1el of 1.22 kcal/mol A1 at pH 6.7 and a value of 4.93 for Ks. When μ and T were held constant but pH was varied, and when allowance was made for the effect of pH changes on the electrical work contribution, a value of 1.1 was found for ζ. This means that something like 1.1 mol of hydrogen ion must be bound per mole of A1 protein in the formation of D. When this is added to the small amount of hydrogen ion bound per A1 before polymerization, at the pH values used, it turned out that for D to be formed, 1.5 H+ ions must be bound per A1 or 0.5 per protein polypeptide chain. This amounts to 1 H+ ion per polypeptide chain for half of the protein units, presumably those in one but not the other layer of the double disk or turn of the double spiral. When polymerization goes beyond the D stage, as shown by previously published data, additional H+ ions are bound. Simultaneous osmotic pressure studies and sedimentation studies were carried out, in both cases as a function of loading concentration C. These results were in complete disagreement with models A and C but agreed reasonably well with model B. The sedimentation studies permitted evaluation of the constant, β, to be 0.33 g/liter.  相似文献   

9.
Kinetic constants for SO42? transport by upper and lower rat ileum in vitro have been determined by computer fitting of rate vs concentration data obtained using the everted sac technique. MoO42? inhibition of this transport is competitive, and kinetic constants for the inhibition were similarly determined. Transport is also inhibited by the anions WO42?, S2O32? and SeO42?, in the order S2O32? > SeO42? ≧ MoO42? > WO42?. These anions have no effect on the transport of l-valine. Low SO42? transport rates were observed in sacs from animals fed a high-molybdenum diet. The significance of the results with respect to the problem of molybdate toxicity in animals is discussed, and related to the known protective effect of SO42?.  相似文献   

10.
Delocalized chemiosmotic coupling of oxidative phosphorylation requires that a single-value correlation exists between the extent of Δ\?gmH+ and the kinetic parameters of respiration and ATP synthesis. This expectation was tested experimentally in nigericin-treated plant mitochondria in single combined experiments, in which simultaneously respiration (in State 3 and in State 4) was measured polarographically, FΔψ (which under these conditions was equivalent to Δ\?gmH+) was evaluated potentiometrically from the uptake of tetraphenylphosphonium+ and the rate of phosphorylation was estimated from the transient depolarization of mitochondria during State 4-State 3-State 4 transitions. The steady-state rates of the different biochemical reactions were progressively inhibited by specific inhibitors active with different modalities on various steps of the energy-transducing process: succinate respiration was inhibited competitively with malonate or noncompetitively with antimycin A, or by limiting the rate of transport into the mitochondria of the respiratory substrate with phenylsuccinate; Δ\?gmH+ was dissipated by uncoupling with increasing concentrations of valinomycin; ADP phosphorylation was limited with oligomycin. The results indicate generally that when the rate of respiratory electron flow is decreased, a parallel inhibition of the rate of phosphorylation is also observed, while very limited effects can be detected on the extent of Δ\?gmH+. This behavior is in marked contrast to the effect of uncoupling where the decreased rate of ATP synthesis is clearly due to energy limitation. Extending previous observations in bacterial photosynthesis and in respiration by animal mitochondria and submitochondrial particles the results indicate, therefore, that respiration tightly controls the rate of ATP synthesis, with a mechanism largely independent of Δ\?gmH+. These data cannot be reconciled with a delocalized chemiosmotic coupling model.  相似文献   

11.
5-hydroxylysine, an analogue of glutamate and lysine, causes NH4+ production by N2-fixing A. cylindrica; it also reversibly inhibits GS activity in vitro but has no effect on alanine dehydrogenase or GOGAT. On adding 5-hydroxylysine intracellular pools of glutamine, glutamate and aspartate decrease; those of alanine and serine increase. 5-hydroxylysine alleviates the inhibitory effect of NH4+ on heterocyst production and C2H2 reduction and in NH4+-grown cultures results in heterocyst synthesis and in C2H2 reduction. The data suggest that the GS-GOGAT pathway is the sole route of importance in primary NH4+ assimilation in A. cylindrica, that NH4+ alone does not inhibit nitrogenase and heterocyst production, and that GS and/or a product is involved in regulating the production of both.  相似文献   

12.
Respiration-driven proton translocation has been studied with the oxidant pulse method for cells of denitrifying Paracoccus denitrificans oxidizing H2 during reduction of O2, NO?3, NO?2 or N2O. A simplified scheme of anaerobic electron transport and associated proton translocation is shown that is consistent with the measured H+oxidant ratios. Furthermore, the kinetics and energetics of NO?3 uptake in whole cells of P. denitrificans were studied. For this purpose, we measured H2 consumption or N2O production after addition of NO?3 to a cell suspension, which indirectly gave information about uptake (and reduction) of NO?3. It was found that a lag phase in H2 consumption or N2O production appeared whenever the membrane potential was dissipated by addition of thiocyanate, carbonyl cyanide m-chlorophenylhydrazone or triphenyl-methylphosphonium bromide. However, these lag phases were not observed when NO?2 was present at the moment of introduction of NO?3. On the basis of these findings we conclude that there are two uptake systems for NO?3. One system is dependent on the proton-motive force and is probably used for initiation of NO?3 uptake. The other is an NO?3NO?2 antiport and its function is to take over NO?3 uptake from the first system.  相似文献   

13.
G. Peters  M.A.J. Rodgers 《BBA》1981,637(1):43-52
Laser flash photolysis techniques have yielded rate constants for physical and reactive quenching modes of O2(1Δg) by nicotine, nicotinamide adenine dinucleotide (oxidized and reduced forms) and the reduced forms of nicotinamide mononucleotide, nicotinamide adenine dinucleotide phosphate and nicotinamide hypoxanthine dinucleotide. In the case of the last four named compounds, kinetic spectroscopy furnished evidence for one-electron transfers to O2(1Δg). Specifically, production of O?2 was demonstrated unequivocally by reaction with 1,4-benzoquinone. Quantitative determinations revealed the extent of reactive quenching to be near 60% in each case.  相似文献   

14.
A plasma membrane-enriched vesicle fraction has been prepared from Trypanosoma brucei by sonication and differential centrifugation on sucrose gradients. This fraction is enriched 5-fold in the plasma membrane marker enzymes adenyl cyclase (EC 4.6.1.1) and ouabain-inhibitable, (Na+ + K+)-dependent adenosine triphosphatase (EC 3.6.1.3). It is also enriched up to 14-fold in iodinated surface proteins, and up to 4-fold in [3H]mannose-labeled glycoproteins, of which the major variable surface coat glycoprotein is the main constituent. Proteins of the plasma membrane fraction and other subcellular fractions have been identified by electrophoretic analysis in sodium dodecyl sulfate-polyacrylamide gradient slab gels. Several high molecular weight surface glycopeptides have been selectively investigated and partially characterized by a combination of metabolic labeling with [3H]mannose, lactoperoxidase-catalyzed surface iodination, and affinity chromatography on Con A-Sepharose. In addition to the major variable surface coat glycoprotein (estimated Mr = 58 000), there are several minor surface glycopeptides (Mr = 76 000, 86 000 and 92 000–100 000) which are apparent extrinsic membrane components, and two surface glycopeptides (Mr = 42 000 and 130 000) which are intrinsic membrane components.  相似文献   

15.
The cell-free preparations from autotrophieally grown Pseudomonas saccharophila catalyzed the process of electron transport from H2 or various other organic electron donors to either O2 or NO3? with concomitant ATP generation. The respective PO ratios with H2 and NADH were 0.63 and 0.73, the respective PNO3? ratios were 0.57 and 0.54. In contrast, the PO and PNO3? ratios with succinate were 0.18 and 0.11, respectively. ATP formation coupled to the oxidation of ascorbate, in the absence or presence of added N,N,N′,N′-tetramethyl-p-phenylenediamine or cytochrome c, could not be detected. Various uncouplers inhibited phosphorylation with either O2 or NO3? as terminal electron acceptors without affecting the oxidation of H2 or other substrates. The NADH oxidation at the expense of O2 or NO3? reduction as well as the associated phosphorylation were inhibited by rotenone and amytal. The aerobic and anaerobic H2 oxidation and coupled ATP synthesis, on the other hand, was unaffected by the flavoprotein inhibitors as well as by the NADH trapping system. The NADH, H2, and succinate-linked electron transport to O2 or NO3? and the associated phosphorylations were sensitive, however, to antimycin A or 2-n-nonyl-4-hydroxyquino-line-N-oxide, and cyanide or azide. The data indicated that although the phosphorylation sites 1 and II were associated with NADH oxidation by O2 or NO3?, the energy conservation coupled to H2 oxidation under aerobic or anaerobic conditions appeared to involve site II only.  相似文献   

16.
The reactivities of anionic nitroalkanes with 2-nitropropane dioxygenase of Hansenula mrakii, glucose oxidase of Aspergillus niger, and mammalian d-amino acid oxidase have been compared kinetically. 2-Nitropropane dioxygenase is 1200 and 4800 times more active with anionic 2-nitropropane than d-amino acid oxidase and glucose oxidase, respectively. The apparent Km values for anionic 2-nitropropane are as follows: 2-nitropropane dioxygenase, 1.61 mm; glucose oxidase, 16.7 mm; and d-amino acid oxidase, 11.1 mm. Anionic 2-nitropropane undergoes an oxygenase reaction with 2-nitropropane dioxygenase and glucose oxidase, and an oxidase reaction with d-amino acid oxidase. In contrast, anionic nitroethane is oxidized through an oxygenase reaction by 2-nitropropane dioxygenase, and through an oxidase reaction by glucose oxidase. All nitroalkane oxidations by these three flavoenzymes are inhibited by Cu and Zn-superoxide dismutase of bovine blood, Mn-superoxide dismutases of bacilli, Fe-superoxide dismutase of Serratia marcescens, and other O2? scavengers such as cytochrome c and NADH, but are not affected by hydroxyl radical scavengers such as mannitol. None of the O2? scavengers tested affected the inherent substrate oxidation by glucose oxidase and d-amino acid oxidase. Furthermore, the generation of O2? in the oxidation of anionic 2-nitropropane by 2-nitropropane dioxygenase was revealed by ESR spectroscoy. The ESR spectrum of anionic 2-nitropropane plus 2-nitropropane dioxygenase shows signals at g1 = 2.007 and g11 = 2.051, which are characteristic of O2?. The O2? generated is a catalytically essential intermediate in the oxidation of anionic nitroalkanes by the enzymes.  相似文献   

17.
Proton inventory investigations of the hydrolysis N-acetylbenzotriazole at pH 3.0 (or the equivalent point on the pD rate profile) have been conducted at two different temperatures and at ionic strengths ranging from 0 to 3.0 M. The solvent deuterium isotope effects and proton inventories are remarkably similar over this wide range of conditions. The proton inventories suggest a cyclic transition state involving four protons contributing to the solvent deuterium isotope effect for the water-catalyzed hydrolysis. The hydrolysis data are described by the equation kn = ko (1 ? n + nπa1)4 with πa1 ~ 0.74, where ko is the observed first-order rate constant in protium oxide, n is the atom fraction of deuterium in the solvent, kn is the rate constant in a protium oxide-deuterium oxide mixture, and πa1 is the isotopic fractionation factor.  相似文献   

18.
Activity levels of sulfotransferases, requisite for the sulfation of chondroitin sulfate proteoglycan, were measured in cell-free homogenates prepared from neonatal epiphyseal cartilage of normal C57B1/6J or homozygous brachymorphic mice. In the presence of [35S]-PAPS only or [35S]-PAPS plus an exogenous sulfate acceptor, comparable amounts of 35SO42? were incorporated into chondroitin sulfate by the normal and mutant types of cartilage. In contrast, the mutant cartilage catalyzed the conversion of only 30% of the 35SO42? into chondroitin sulfate as compared to normal mouse cartilage when synthesis was initiated from ATP and H235SO4. These results suggest that the production of an undersulfated proteoglycan which has previously been reported in brachymorphic mice (Orkin, R.W. etal. (1976) Devel. Biol. 50, 82–94) may result from a defect in the synthesis of the sulfate donor PAPS.  相似文献   

19.
The kinetics of bisulfite addition to 5-fluorouracil were studied as a function of increasing concentrations of potential general acids. Values of kobsd[SO3=] measured at 25°C and ionic strength 1.0 M increased linearly and then became invariant with increasing concentrations of either HSO3? or (OHCH2CH2)2N+C(CH2OH)3 HCl (BisTris+HCl). A small kinetic hydrogen-deuterium isotope effect (kHSkDS = 1.10) was observed for the general acid catalysed portion of the addition reaction. The kinetics of bisulfite elimination from 5-fluoro-5,6-dihydrouracil-6-sulfonate were studied in ethanolamine buffers. As previously observed with 1,3-dimethyl-5,6-dihydrouracil-6-sulfonate, this reaction is subject to general base catalysis and exhibits a large kinetic hydrogen-deuterium isotope effect (k2H2Ok2D2O = 3.8). The kinetic results for the addition reaction are consistent with a multistep reaction pathway involving the initial formation of an oxyanion sulfite addition intermediate (II) which subsequently adds a proton and undergoes tautomerization to yield the final 5-fluoro-5,6-dihydrouracil-6-sulfonate product. Thus the elimination of bisulfite from 5-fluoro-5,6-dihydrouracil-6-sulfonate probably proceeds by an ElcB mechanism which involves, at relatively low concentrations of general base, rate determining general base catalyzed proton abstraction from carbon 5 to yield intermediate II followed by the rapid elimination of sulfite to yield 5-fluorouracil. These results may be related to both the enzymatically catalyzed dehalogenation of bromoand iodouracil and the methylation of deoxyuridylate by thymidylate synthetase.  相似文献   

20.
The pH dependence of the reaction of tris(hydroxymethyl)aminomethane (Tris) with the activated carbonyl compound 4-trans-benzylidene-2-phenyloxazolin-5-one (I) is given by the equation k′2 = kbKa(Ka + [H+]) + ka[OH?]Ka(Ka + [H+]), where Ka is the dissociation constant of TrisH+. Spectrophotometric experiments show that the Tris ester of α-benzamido-trans-cinnamic acid is formed quantitatively over a range of pH values, regardless of the relative contribution of kb and ka terms to k2. Hence, both terms refer to alcoholysis. While the mechanism of the reaction is not determined unequivocally in the present work, the magnitude of the kb term, together with its dependence on the basic form of Tris, suggests that ester formation is occurring by nucleophilic attack of a Tris hydroxyl group on the carbonyl carbon of the oxazolinone, with intramolecular catalysis by the Tris amino group. The rate enhancement due to this group is at least 102 and possibly of the order 106. This system is compared with other model systems for the acylation step of catalysis by serine esterases and proteinases.  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