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.
A method has been developed for calculating rate constants for dehydration of aldehydes that induce ATPase reactions by kinases and where 18O is transferred from the aldehyde or its hydrate to inorganic phosphate during the reaction. The method involves measurement of the fraction of 18O in phosphate by 31P NMR after the ATPase reaction has proceeded for several minutes with zero-order kinetics. The reaction is started by addition of the aldehyde in a small volume of H2 18O, and the speed of washout of 18O by reversible dehydration relative to the rate of the ATPase reaction allows calculation of the rate constants if the hydration equilibrium constant is known from the proton NMR spectrum of the aldehyde. Dehydration rate constants (s-1 at pH 8-8.5, 0.1 M buffer, 25 degrees C) for the following aldehydes (all over 95% hydrated) and kinases used are as follows: D-glyceraldehyde with glycerokinase, 0.03; 2,5-anhydro-D-mannose 6-phosphate with fructose-6-phosphate kinase, 0.025; 2,5-anhydro-D-mannose or 2,5-anhydro-D-talose with fructokinase, 0.029 and 0.017, respectively; D-gluco-hexodialdose with hexokinase, 0.068. With betaine aldehyde and choline kinase or glyoxylate and pyruvate kinase, no 18O was transferred to phosphate during the ATPase reactions. However, the dehydration rate constant for glyoxylate (0.007 s-1 at pH 7 extrapolated to zero buffer concentration and up to 0.11 s-1 at pH 9.0 with 0.3 M buffer) was determined by extrapolating the initial rate of reduction of the free aldehyde catalyzed by lactate dehydrogenase to infinite enzyme levels.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
A R Rendina  W W Cleland 《Biochemistry》1984,23(22):5157-5168
Aldehyde analogues of the normal alcohol substrates induce ATPase activities by glycerokinase (D-glyceraldehyde), fructose-6-phosphate kinase (2,5-anhydromannose 6-phosphate), fructokinase (2,5-anhydromannose or 2,5-anhydrotalose), hexokinase (D-gluco-hexodialdose), choline kinase (betaine aldehyde), and pyruvate kinase (glyoxylate). Since purified deuterated aldehydes give V and V/K isotope effects near 1.0 for glycerokinase, fructokinase with 2,5-anhydro[1-2H]talose, hexokinase, choline kinase, and pyruvate kinase, the hydrates of these almost fully hydrated aldehydes are the activators of the ATPase reactions. Fructose-6-phosphate kinase and fructokinase with 2,5-anhydro[1-2H]mannose show V/K deuterium isotope effects of 1.10 and 1.22, respectively, suggesting either that both hydrate and free aldehyde may be activators (predicted values are 1.37 if only the free aldehyde activates the ATPase) or, more likely, that the phosphorylated hydrate breaks down in a rate-limiting step on the enzyme while MgADP is still present and the back-reaction to yield free hydrate in solution is still possible. 18O was transferred from the aldehyde hydrate to phosphate during the ATPase reactions of glycerokinase, fructose-6-phosphate kinase, fructokinase, and hexokinase but not with choline kinase or pyruvate kinase. Thus, direct phosphorylation of the hydrates by the first four enzymes gives the phosphate adduct of the aldehyde, which decomposes nonenzymatically, while with choline kinase and pyruvate kinase the hydrates induce transfer to water (metal-bound hydroxide or water with pyruvate kinase on the basis of pH profiles). Observation of a lag in the release of phosphate from the glycerokinase ATPase reaction at 15 degrees C supports the existence of a phosphorylated hydrate intermediate with a rate constant for breakdown of 0.035-0.043 s-1 at this temperature. Kinases that phosphorylate creatine, 3-phosphoglycerate, and acetate did not exhibit ATPase activities in the presence of keto or aldehyde analogues (N-methylhydantoic acid, D-glyceraldehyde 3-phosphate, and acetaldehyde, respectively), possibly because of the absence of an acid-base catalytic group in the latter two cases. These analogues were competitive inhibitors vs. the normal substrates, and in the latter case, the hydrate of acetaldehyde was shown to be the inhibitory species on the basis of the deuterium isotope effect on the inhibition constant.  相似文献   

3.
Phosphoenolpyruvate carboxykinase from chicken liver mitochondria and rat liver cytosol catalyzes the phosphorylation of alpha-substituted carboxylic acids such as glycolate, thioglycolate, and DL-beta-chlorolactate in reactions with absolute requirements for divalent cation activators. 31P NMR analysis of the reaction products indicates that phosphorylation occurs at the alpha-position to generate the corresponding O- or S-bridged phosphate monoesters. In addition, the enzymes catalyze the bicarbonate-dependent phosphorylation of hydroxylamine. The chicken liver enzyme also catalyze the bicarbonate-dependent phosphorylation of hydroxylamine. The chicken liver enzyme also catalyzes the bicarbonate-dependent phosphorylation of fluoride ion. The kappa cat values for these substrates are 20-1000-fold slower than the kappa cat for oxaloacetate. Pyruvate and beta-hydroxypyruvate are not phosphorylated, since the enzyme does not catalyze the enolization of these compounds. Oxalate, a structural analogue of the enolate of pyruvate, is a competitive inhibitor of phosphoenolpyruvate carboxykinase (Ki of 5 microM) in the direction of phosphoenolpyruvate formation. Oxalate is also an inhibitor of the chicken liver enzyme in the direction of oxaloacetate formation and in the decarboxylation of oxaloacetate. The chicken liver enzyme is inhibited by beta-sulfopyruvate, an isoelectronic analogue of oxaloacetate. The extensive homologies between the reactions catalyzed by phosphoenolpyruvate carboxykinase and pyruvate kinase suggest that the divalent cation activators in these reactions may have similar functions. The substrate specificity indicates that phosphoenolpyruvate carboxykinase decarboxylates oxaloacetate to form the enolate of pyruvate which is then phosphorylated by MgGTP on the enzyme.  相似文献   

4.
Benzyloxycarbonyl (Z)-Ala-Pro-Phe-glyoxal and Z-Ala-Ala-Phe-glyoxal have both been shown to be inhibitors of alpha-chymotrypsin with minimal Ki values of 19 and 344 nM, respectively, at neutral pH. These Ki values increased at low and high pH with pKa values of approximately 4.0 and approximately 10.5, respectively. By using surface plasmon resonance, we show that the apparent association rate constant for Z-Ala-Pro-Phe-glyoxal is much lower than the value expected for a diffusion-controlled reaction. 13C NMR has been used to show that at low pH the glyoxal keto carbon is sp3-hybridized with a chemical shift of approximately 100.7 ppm and that the aldehyde carbon is hydrated with a chemical shift of approximately 91.6 ppm. The signal at approximately 100.7 ppm is assigned to the hemiketal formed between the hydroxy group of serine 195 and the keto carbon of the glyoxal. In a slow exchange process controlled by a pKa of approximately 4.5, the aldehyde carbon dehydrates to give a signal at approximately 205.5 ppm and the hemiketal forms an oxyanion at approximately 107.0 ppm. At higher pH, the re-hydration of the glyoxal aldehyde carbon leads to the signal at 107 ppm being replaced by a signal at 104 ppm (pKa approximately 9.2). On binding either Z-Ala-Pro-Phe-glyoxal or Z-Ala-Ala-Phe-glyoxal to alpha-chymotrypsin at 4 and 25 degrees C, 1H NMR is used to show that the binding of these glyoxal inhibitors raises the pKa value of the imidazolium ion of histidine 57 to a value of >11 at both 4 and 25 degrees C. We discuss the mechanistic significance of these results, and we propose that it is ligand binding that raises the pKa value of the imidazolium ring of histidine 57 allowing it to enhance the nucleophilicity of the hydroxy group of the active site serine 195 and lower the pKa value of the oxyanion forming a zwitterionic tetrahedral intermediate during catalysis.  相似文献   

5.
The aldehyde (RS)-2-benzyl-4-oxobutanoic acid, which is 25% hydrated at pH 7.5, has recently been shown to be a strong reversible competitive inhibitor of carboxypeptidase A [Ki = 0.48 nM; Galardy, R. E., & Kortylewicz, Z. P. (1984) Biochemistry 23, 2083-2087]. The ketone analogue of this aldehyde (RS)-2-benzyl-4-oxopentanoic acid (IV) is not detectably hydrated under the same conditions and is 1500-fold less potent (Ki = 730 microM). The ketone homologue (RS)-2-benzyl-5-oxohexanoic acid (XIII) is also a weak inhibitor (Ki = 1.3 mM). The alpha-monobrominated derivatives of these two ketones are, however, strong competitive inhibitors with Ki's of 0.57 microM and 1.3 microM, respectively. Oximes derived from the aldehyde, the ketones IV and XIII, and a homologue of the aldehyde are weak inhibitors with Ki's ranging from 480 to 7900 microM. The inhibition of carboxypeptidase A by the alpha-monobrominated ketones is reversible and independent of the time (up to 6 h) of incubation of enzyme and inhibitor together. Bromoacetone at a concentration of 30 mM does not inhibit carboxypeptidase A. Incubation of an equimolar mixture of 2-benzyl-4-bromo-5-oxohexanoic acid (XV) and enzyme for 1 h led to the recovery of 82% of XV, demonstrating that it is the major species reversibly bound during assay of inhibition. Taken together, these results indicate that tight binding of carbonyl inhibitors to carboxypeptidase A requires specific binding of inhibitor functional groups such as benzyl and an electrophilic carbonyl carbon such as that of an alpha-bromo ketone or aliphatic aldehyde.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
SCH-23390 is a high-affinity antagonist selective for D1 dopamine receptors (Ki = 2.5 nM). It does not contain a functional group that can be conveniently coupled to commercially available resins for affinity chromatography or to prepare photolabels for photoaffinity labeling of receptors. To construct an affinity resin for purification of dopamine D1 receptors, an aldehyde analogue of SCH-23390, (+/-)-7-chloro-8-hydroxy-1-(4'-formylphenyl)-3-methyl-2,3,4,5-tetrahydro -1H- 3-benzazepine (ASCH), was synthesized. 8-Methoxy-1-(4'-bromophenyl)-SCH-23390 was lithiated, formylated, and O-demethylated to form the aldehyde. NMR and IR analyses were performed to characterize the product. Assays were performed with the radioligand [125I]SCH-23982 to define the biological activity of the aldehyde. ASCH displaced [125I]SCH-23982 binding from caudate membranes with a Ki value of 7.1 nM. ASCH has been coupled through the aldehyde group on the phenyl ring to diaminodipropylamine-agarose for affinity chromatography. After solubilization of caudate membranes in 1% digitonin, the affinity resin retained binding sites for [125I]SCH-23982 that were eluted with 10 mM SCH-23390. The aldehyde was also covalently coupled to biotin hydrazide for fluorescence labeling of dopamine D1 receptors. The biotin-conjugated aldehyde of SCH-23390 displaced [125I]SCH-23982 binding from caudate membranes with a Ki value of 9.3 nM.  相似文献   

7.
1. Pig heart pyruvate dehydrogenase complex is inactivated by phosphorylation (MgATP2-) of an alpha-chain of the decarboxylase component. Three serine residues may be phosphorylated, one of which (site 1) is the major inactivating site. 2. The relative rates of phosphorylation are site 1 greater than 2 greater than site 3. 3. The kinetics of the inactivating phosphorylation were investigated by measuring inactivation of the complex with MgATP2-. The apparent Km for the Mg complex of ATP was 25.5 microM; ADP was a competitive inhibitor (Ki 69.8 microM) and sodium pyruvate an uncompetitive inhibitor (Ki 2.8 microM). Inactivation was accelerated by increasing concentration ratios of NADH/NAD+ and of acetyl-CoA/CoA. 4. The kinetics of additional phosphorylations (predominantly site 2 under these conditions) were investigated by measurement of 32P incorporation into non-radioactive pyruvate dehydrogenase phosphate containing 3-6% of active complex, and assumed from parrallel experiments with 32P labelling to contain 91% of protein-bound phosphate in site 1 and 9% in site 2. 5. The apparent Km for the Mg complex of ATP was 10.1 microM; ADP was a competitive inhibitor (Ki 31.5 microM) and sodium pyruvate an uncompetitive inhibitor (Ki 1.1 mM). 6. Incorporation was accelerated by increasing concentration ratios of NADH/NAD+ and of acetyl-CoA/CoA, although it was less marked at the highest ratios.  相似文献   

8.
Enolases (2-phospho-d-glycerate hydrolase, EC 4.2.1.11) were purified from both pig liver and muscle. Graphs of InC vs.r 2 from sedimentation equilibrium experiments are linear, which suggests homogeneous preparations of liver and muscle enolases. From these data the molecular weight of liver enolase is calculated to be approximately 92,000 D and that of muscle enolase to be approximately 85,000 D. SDS-PAGE experiments give a molecular weight value of 46,000 D for liver enolase and a value of 44,000 D for muscle enolase. These molecular weight values for liver and muscle enzymes are within the range for other enolases and show that both of these pig enolases are dimers. Amino acid composition data support the sedimentation equilibrium data and also give a smaller molecule weight (84,968 D) for muscle enolase compared to that of the liver enzyme (89,021 D). The two enzymes differ in their content of lysine [liver enolase (L)=94 residues, muscle enolase (M)=68 residues], histidine (L=13, M=21), serine (L=53, M=36), proline (L=52, M=34), and cysteine (L=4, M=21). Partial specific volumes of 0.737 ml/g for liver enolase and 0.735 ml/g for muscle enolase were calculated from the amino acid composition data. Pig liver and muscle enolases differ radically in their isoelectric points (pI=6.4–6.5 for liver enolase, and pI=8.8–9.0 for muscle enolase), and in their degree of inactivation by 750 mM LiCI (liver enolase is inactivated to a greater degree than the muscle enolase). Despite these physical and chemical differences, the kinetic constantsK M values for Mg2+, 2-phosphoglyceric acid, and phospho(enol)pyruvate appear not to be significantly different for these two forms of enolase. The physical, chemical, and kinetic data for pig liver and muscle enolases are compared to similar data for pig kidney enolase.  相似文献   

9.
Inhibition of CA V decreases glucose synthesis from pyruvate   总被引:1,自引:0,他引:1  
The carbonic anhydrase inhibitor acetazolamide reduces citrulline synthesis by intact guinea pig liver mitochondria and also inhibits mitochondrial carbonic anhydrase (CA V) and the more lipophilic carbonic anhydrase inhibitor ethoxzolamide reduces urea synthesis by intact guinea pig hepatocytes in parallel with its inhibition of total hepatocytic carbonic anhydrase activity. Intact hepatocytes from 48-h starved male guinea pig livers were incubated at 37 degrees C in Krebs-Henseleit with 95% O2/5% CO2 at pH 7.1 with 5 mM pyruvate, 5 mM lactate, 3 mM ornithine, 10 mM NH4Cl, 1 mM oleate; with these inclusions both urea and glucose synthesis start with HCO3- -requiring enzymes, carbamyl phosphate synthetase I and pyruvate carboxylase, respectively. Urea and glucose synthesis were inhibited in parallel by increasing concentrations of ethoxzolamide, estimated Ki for each approximately 0.1 mM. In other experiments hepatocytes were incubated at 37 degrees C in Krebs-Henseleit with 95% O2/5% CO2 at pH 7.1 with 10 mM glutamine, 1 mM oleate; with these inclusions glucose synthesis no longer starts with a HCO3- -requiring enzyme. Urea synthesis was inhibited by ethoxzolamide with an estimated Ki of 0.1 mM, but glucose synthesis was unaffected. Intact mitochondria were prepared from 48-h starved male guinea pig livers. Pyruvate carboxylase activity of intact mitochondria was determined in isotonic KCl-Hepes buffer, pH 7.4, 25 degrees C, with 7.5 mM pyruvate, 3 mM ATP, and 10 mM NaHCO3. Inclusion of ethoxzolamide resulted in reduction in the rate of pyruvate carboxylation in intact mitochondria, but not in disrupted mitochondria. It is concluded that carbonic anhydrase is functionally important for gluconeogenesis in the male guinea pig liver when there is a requirement for bicarbonate as substrate.  相似文献   

10.
A fluorescent method for monitoring the activity of the secreted Candida carboxyl (aspartic) proteinase (EC 3.4.23.6) was developed using a fluorogenic substrate based on resonance energy transfer. The fluorescent assay was used to monitor proteinase production, purification, and inhibition. The Km for the fluorogenic substrate, 4-(4-dimethylaminophenylazo)benzoyl-gamma-aminobutyryl-Ile-His-Pro - Phe-His-Leu-Val-Ile-His-Thr- [5-(2-aminoethyl)amino]naphthalene-1-sulfonic acid, was found to be 4.3 microM at the optimum pH of 4.5. Reaction products were separated by reverse-phase high-performance liquid chromatography and identified by amino acid analysis or by 252Cf plasma desorption mass spectrometry. Cleavage of the fluorogenic substrate was between the histidine-threonine residues, releasing the fluorescent product, threonine-[5-(2-aminoethyl)amino]naphthalene-1-sulfonic acid. Proteolytic activity was expressed as nanomoles of fluorescent product released at 22 degrees C/60 min, pH 4.5, and the release of 0.9 nmol product was equivalent to one hemoglobin proteolytic unit (O.D.A700 increase of 0.100) produced at 37 degrees C/60 min, pH 3.5. The aspartic proteinase inhibitor pepstatin had an IC50 of 27 nM when tested in a dose-response study with the purified enzyme. The apparent Ki for pepstatis was 2.9 nM. Several synthetic inhibitors of the enzymes were identified with IC50's in the nanomolar range. The most potent compound, A70450, was characterized as a fast, tight-binding inhibitor having an IC50 of 1.3 nM and apparent Ki of 0.17 nM.  相似文献   

11.
Glutaminol adenylate 5 is a competitive inhibitor of glutaminyl-tRNA synthetase with respect to glutamine (Ki = 280 nM) and to ATP (Ki = 860 nM). The corresponding methyl phosphate ester 4 is a weaker inhibitor (Ki approximately 10 microM) with respect to glutamine.  相似文献   

12.
In contrast to the pyruvate dehydrogenase complex (PDC) from animal mitochondria, our in situ and in vitro studies indicate that the ATP:ADP ratio has little or no effect in regulating the mitochondrial pyruvate dehydrogenase complex from green pea seedlings. Pyruvate was a competitive inhibitor of ATP-dependent inactivation (Ki = 59 microM), while the PDC had a Km for pyruvate of microM. Thiamine pyrophosphate, the coenzyme for the pyruvate dehydrogenase (PDH) component of the complex, did not inhibit ATP-dependent inactivation when used alone but it enhanced inhibition by pyruvate. As such, thiamine pyrophosphate was a competitive inhibitor (Ki = 130 nM) of ATP-dependent inactivation. A model is proposed for the pyruvate plus thiamine pyrophosphate inhibition of ATP-dependent inactivation of the pyruvate dehydrogenase complex in which pyruvate exerts its inhibition of inactivation by altering or protecting the protein substrate from phosphorylation and not by directly inhibiting PDH kinase.  相似文献   

13.
1-Hydroxycyclopropane carboxylic acid phosphate has been synthesized from diethyl succinate by acyloin condensation followed by ring contraction and phosphorylation. This compound is a potent competitive inhibitor of enzymes utilizing phosphoenolpyruvate. For phosphoenolpyruvate from maize, Ki = 7.3 μM at pH 8.0 in the presence of Mg2+. For pyruvate kinase, Ki = 2.0 mM at pH 7.0. For enolase, Ki = 8.0 μM at pH 8.0. In each case, this compound is a substantially better inhibitor than the commonly used phosphoenolpyruvate analogs phosphoglycolate and phospholactate, presumably because of the similarity in geometric and electronic structure between the cyclopropane compound and phosphoenolpyruvate.  相似文献   

14.
The phosphorylation of pig liver pyruvate kinase by cyclic adenosine 3':5'-monophosphate-dependent protein kinase has been studied. For comparison, mixed histone and a synthetic heptapeptide were also used as substrates. Protein kinase was purified by chromatography on DEAE-cellulose, hydroxyapatite, and Sephadex G-200. The enzyme was stimulated by cyclic AMP with apparent Ka values of 2.5 and 0.8 x 10-7 M for pyruvate kinase and histone substrates, respectively. Divalent cations were essential for the activity of the protein kinase. Variation of the concentration of ATP resulted in approximately straight lines in Lineweaver-Burk plots for the phosphorylation of both pyruvate kinase and mixed histone. The apparent Km values for ATP were 21 and 11 muM, respectively. The phosphorylation rate increased with the concentration of pyruvate kinase even at a concentration of 2 muM pyruvate kinase. At a high ionic strength, the phosphorylation rate of both pyruvate kinase and histone decreased. The phosphorylation rate varied markedly with pH in imidazole/HC1 and Tris/HC1 buffers. At slightly alkaline pH values, pyruvate kinase was phosphorylated at a much higher rate than pH7, but this was not the case for histone. At pH 8.5, the phosphorylation rate of pyruvate kinase was 3.5 times the rate at pH 7, while the corresponding increase for the histone phosphorylation was 50 per cent. In potassium phosphate buffers, the phosphorylation rate of both substrates did not change significantly over the pH range studied. Arrhenius' plots of the protein kinase reaction resulted in a break at about 10 degrees when pyruvate kinase was used as substrate, whereas a straight line was obtained when using histone. The negative allosteric effectors of pyruvate kinase, alanine, and phenylalanine, increased the phosphorylation rate of pyruvate kinase at pH 8 by 50 and 120 per cent, respectively. The same effectors did not influence the phosphorylation rate of mixed histone or a synthetic heptapeptide. It is concluded that the conformations adopted by pyruvate kinase in the presence of allosteric inhibitors make it a better substrate for the protein kinase.  相似文献   

15.
S Uribe  P Rangel  J P Pardo 《Cell calcium》1992,13(4):211-217
The interactions of Ca2+ with mitochondria from Saccharomyces cerevisiae were explored. Mitochondria were loaded with the metallochromic dye Fluo-3 to measure the concentration of free calcium in the matrix. Addition of EGTA or Ca2+ led to fluctuations in mitochondrial free calcium between 120 and 400 nM. Ca2+ variations were slower at 4 degrees C than at 25 degrees C or in the presence of phosphate instead of acetate. The net uptake of 45Ca2+ was higher with phosphate than with acetate. The optimum pH for Ca2+ uptake was 6.8. Ruthenium red did not affect the uptake of Ca2+. Addition of antimycin-A or uncouplers led to a small and transient release of Ca2+. Addition of EGTA or the monovalent cations Na+ or K+ resulted in higher release of Ca2+. Site I but not site II dependent O2 consumption was partially inhibited by EGTA. The effect of Ca2+ on NADH oxidation is similar to results reported with enzymes from mammalian sources which use NADH, such as the pyruvate, isocitrate and oxoglutarate dehydrogenases.  相似文献   

16.
We determined the kinetics of the reaction of human neuronal enolase and yeast enolase 1 with the slowly-reacting chromophoric substrate d-tartronate semialdehyde phosphate (TSP), each in tris (tris (hydroxymethyl) aminomethane) and another buffer at several Mg2+ concentrations, 50 or 100 μM, 1 mM and 30 mM. All data were biphasic, and could be satisfactorily fit, assuming either two successive first-order reactions or two independent first-order reactions. Higher Mg2+ concentrations reduce the relative magnitude of the slower reaction. The results are interpreted in terms of a catalytically significant interaction between the two subunits of these enzymes.  相似文献   

17.
Steady-state kinetics of carboxypeptidase Y, a proteinase from yeast, were studied by using the reaction of 4-nitrophenyl trimethylacetate as a probe. The pH profile of kcat. is sigmoidal in H2O-based buffers for the carboxypeptidase Y-catalysed hydrolysis of this ester (kcat. referring to the rate of deacylation of trimethylacetyl-carboxypeptidase Y). The corresponding pD profile in 2H2O is doubly sigmoidal, with inflexions at pD approximately 3.8 and approximately 6.8. The ionization of pKDapp. approximately 3.8 is caused by a rapid inactivation in 2H2O media by a process that is only slowly reversed on transfer to pH 7.00 phosphate buffer in H2O. The corresponding inactivation in H2O-based buffers of low pH is considerably slower (approximately 30-fold), follows a first-order rate-dependence and is very strongly pH-dependent, indicating some form of co-operative change in enzyme tertiary structure.  相似文献   

18.
Elementary reactions have been studied quantitatively in the complex overall process catalyzed by horseradish peroxidase whereby isobutyraldehyde and molecular oxygen react to form triplet state acetone and formic acid. The rate constant for the reaction of the enol form of isobutyraldehyde with compound I of peroxidase is (8 +/- 1) X 10(6) M-1 s-1 and with compound II (1.3 +/- 0.3) X 10(6) M-1 s-1. Neither the enolate anion nor the keto form is reactive. The reactivity of enols with peroxidase parallels that of unionized phenols and a common mechanism is proposed. The overall catalyzed reaction of isobutyraldehyde and oxygen consists of an initial burst followed by a steady state phase. The burst is caused by the following sequence: 1) an initial high yield of compound I is formed from reaction of native enzyme with the autoxidation product of isobutyraldehyde, a peracid and 2) compound I rapidly depletes the equilibrium pool of enol which is present. After this burst a steady state phase is observed in which the rate-limiting step is the conversion of the keto to the enol form of the aldehyde catalyzed by phosphate buffer. The rate constant for the keto form reacting with phosphate is (8.7 +/- 0.6) X 10(-5) M-1 s-1. All constants were measured in dilute aqueous ethanol at 35 degrees C, pH 7.4, and ionic strength 0.67 M. Both the initial burst of light and the steady state emission from triplet acetone can be observed with the naked eye. Since the magnitude of the burst is a measure of the equilibrium amount of enol, the keto-enol equilibrium constant is readily calculated and hence also the rate constant for conversion of enol to keto. The keto-enol equilibrium constant is unaffected by phosphate which therefore acts as a true catalyst.  相似文献   

19.
G Cevc  A Watts  D Marsh 《Biochemistry》1981,20(17):4955-4965
The dependence of the gel-to-fluid phase transition temperature of dimyristoyl- and dipalmitoylphosphatidylserine bilayers on pH, NaCl concentration, and degree of hydration has been studied with differential scanning calorimetry and with spin-labels. On protonation of the carboxyl group (pK2app = 5.5), the transition temperature increases from 36 to 44 degrees C in the fully hydrated state of dimyristoylphosphatidylserine (from 54 to 62 degrees C for dipalmitoylphosphatidylserine), at ionic strength J = 0.1. In addition, at least two less hydrated states, differing progressively by 1 H2O/PS, are observed at low pH with transition temperatures of 48 and 52 degrees C for dimyristoyl- and 65 and 68.5 degrees C for dipalmitoylphosphatidylserine. On deprotonation of the amino group (pK3app = 11.55) the transition temperature decreases to approximately 15 degrees C for dimyristoyl- and 32 degrees C for dipalmitoylphosphatidylserine, and a pretransition is observed at approximately 6 degrees C (dimyristoylphosphatidylserine) and 21.5 degrees C (dipalmitoylphosphatidylserine), at J = 0.1. No titration of the transition is observed for the fully hydrated phosphate group down to pH less than or equal to 0.5, but it affinity for water binding decreases steeply at pH greater than or equal to 2.6. Increasing the NaCl concentration from 0.1 to 2.0 M increases the transition temperature of dimyristoyphosphatidylserine by approximately 8 degrees C at pH 7, by approximately 5 degrees at pH 13, and by approximately 0 degrees C at pH 1. These increases are attributed to the screening of the electrostatic titration-induced shifts in transition temperature. On a further increase of the NaCl concentration to 5.5 M, the transition temperature increases by an additional 9 degree C at pH 7, 13 degree C at pH 13, approximately 7 degree C in the fully hydrated state at pH 1, and approximately 4 and approximately 0 degree C in the two less hydrated states. These shifts are attributed to displacement of water of hydration by ion binding. From the salt dependence it is deduced that the transition temperature shift at the carboxyl titration can be accounted for completely by the surface charge and change in hydration of approximately 1 H2O/lipid, whereas that of the amino group titration arises mostly from other sources, probably hydrogen bonding. The shifts in pK (delta pK2 = 2.85, delta pK3 = 1.56) are consistent with a reduced polarity in the head-group region, corresponding to an effective dielectric constant epsilon approximately or equal to 30, together with surface potentials of psi congruent to -100 and -150 mV at the carboxyl and amino group pKs, respectively. The transition temperature of dimyristoylphosphatidylserine-water mixtures decreases by approximately 4 degree C each water/lipid molecule added, reaching a limiting value at a water content of approximately 9-10 H2O/lipid molecule.  相似文献   

20.
Cosgrove S  Rogers L  Hewage CM  Malthouse JP 《Biochemistry》2007,46(39):11205-11215
Z-Ala-Ala-Phe-glyoxal (where Z is benzyloxycarbonyl) has been shown to be a competitive inhibitor of pepsin with a Ki = 89 +/- 24 nM at pH 2.0 and 25 degrees C. Both the ketone carbon (R13COCHO) and the aldehyde carbon (RCO13CHO) of the glyoxal group of Z-Ala-Ala-Phe-glyoxal have been 13C-enriched. Using 13C NMR, it has been shown that when the inhibitor is bound to pepsin, the glyoxal keto and aldehyde carbons give signals at 98.8 and 90.9 ppm, respectively. This demonstrates that pepsin binds and preferentially stabilizes the fully hydrated form of the glyoxal inhibitor Z-Ala-Ala-Phe-glyoxal. From 13C NMR pH studies with glyoxal inhibitor, we obtain no evidence for its hemiketal or hemiacetal hydroxyl groups ionizing to give oxyanions. We conclude that if an oxyanion is formed its pKa must be >8.0. Using 1H NMR, we observe four hydrogen bonds in free pepsin and in pepsin/Z-Ala-Ala-Phe-glyoxal complexes. In the pepsin/pepstatin complex an additional hydrogen bond is formed. We examine the effect of pH on hydrogen bond formation, but we do not find any evidence for low-barrier hydrogen bond formation in the inhibitor complexes. We conclude that the primary role of hydrogen bonding to catalytic tetrahedral intermediates in the aspartyl proteases is to correctly orientate the tetrahedral intermediate for catalysis.  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