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.
Kinetics for the acidic hydrolysis of several N6-substituted 2'-deoxyadenosines were studied in a wide pH-range. The proportions of the partial reactions proceeding via mono- and di-protonated substrates were estimated on the bases of the rate profiles obtained and the acidity constants determined spectrophotometrically for the monocations. The site of the initial protonation was established by the effects that trifluoroacetic acid exerted on the 15N NMR chemical shifts. The exceptional lability of the monocations of N6-acyl protected compounds is suggested to result from the preferred N7 protonation.  相似文献   

2.
The kinetics of excision of damaged purine bases from oxidatively damaged DNA by Escherichia coli Fpg protein were investigated. DNA substrates, prepared by treatment with H2O2/Fe(III)-EDTA or by gamma-irradiation under N2O or air, were incubated with Fpg protein, followed by precipitation of DNA. Precipitated DNA and supernatant fractions were analyzed by gas chromatography/isotope-dilution mass spectrometry. Kinetic studies revealed efficient excision of 8-hydroxyguanine (8-OH-Gua), 2,6-diamino-4-hydroxy-5-formamidopyrimidine (FapyGua) and 4, 6-diamino-5-formamidopyrimidine (FapyAde). Thirteen other modified bases in the oxidized DNA substrates, including 5-hydroxycytosine and 5-hydroxyuracil, were not excised. Excision was measured as a function of enzyme concentration, substrate concentration, time and temperature. The rate of release of modified purine bases from the three damaged DNA substrates varied significantly even though each DNA substrate contained similar levels of oxidative damage. Specificity constants (kcat/KM) for the excision reaction indicated similar preferences of Fpg protein for excision of 8-OH-Gua, FapyGua and FapyAde from each DNA substrate. These findings suggest that, in addition to 8-OH-Gua, FapyGua and FapyAde may be primary substrates for this enzyme in cells.  相似文献   

3.
The montmorillonite clay catalyzed condensation of activated monocleotides to oligomers of RNA is a possible first step in the formation of the proposed RNA world. The rate constants for the condensation of the phosphorimidazolide of adenosine were measured previously and these studies have been extended to the phosphorimidazolides of inosine and uridine in the present work to determine if substitution of neutral heterocycles for the basic adenine ring changes the reaction rate or regioselectivity. The oligomerization reactions of the 5'-phosphoromidazolides of uridine (ImpU) and inosine (ImpI) on montmorillonite yield oligo(U)s and oligo(I)s as long as heptamers. The rate constants for oligonucleotide formation were determined by measuring the rates of formation of the oligomers by HPLC. Both the apparent rate constants in the reaction mixture and the rate constants on the clay surface were calculated using the partition coefficients of the oligomers between the aqueous and clay phases. The rate constants for trimer formation are much greater than those dimer synthesis but there was little difference in the rate constants for the formation of trimers and higher oligomers. The overall rates of oligomerization of the phosphorimidazolides of purine and pyrimidine nucleosides in the presence of montmorillonite clay are the same suggesting that RNA formed on the primitive Earth could have contained a variety of heterocyclic bases. The rate constants for oligomerization of pyrimidine nucleotides on the clay surface are significantly higher than those of purine nucleotides since the pyrimidine nucleotides bind less strongly to the clay than do the purine nucleotides. The differences in the binding is probably due to Van der Waals interactions between the purine bases and the clay surface. Differences in the basicity of the heterocyclic ring in the nucleotide have little effect on the oligomerization process.  相似文献   

4.
The induced formation of uricase by the cultured cells of Streptomyces sp. and the effect of purine bases on the enzyme formation were studied. The microorganism was grown in media containing urate and/or purine bases (adenine, guanine, hypoxanthine or xanthine) and the development of the uricase activity of the cells were measured at intervals. The disappearance of urate and purine bases from the media was also determined. Without the purine bases, the production of uricase was significantly low even in the presence of urate and the disappearance of urate from the medium was in a slow rate. Upon the addition of hypoxanthine or xanthine in the presence of urate, a significant increase in the uricase activity of the cells and a concomitant rapid decrease of urate in the medium were observed. The purine bases added to the media were incorporated into the cells at a relatively early period of the culture and appeared to be converted into urate within the cells. The repression of uricase formation in the cultured cells and the derepression by the addition of the purine bases were discussed.  相似文献   

5.
To characterise the pH dependence of phosphodiester hydrolysis by the EcoRV endonuclease in the presence of Mn2+, single turnover reactions on a 12 bp DNA substrate were examined by stopped-flow and quench-flow methods between pH 6.0 and 8.5. At each pH value, the apparent rate constants for phosphodiester hydrolysis increased hyperbolically with the concentration of MnCl2, thus allowing values to be determined for the intrinsic rate constant at saturation with Mn2+ and the equilibrium dissociation constant for Mn2+. The equilibrium constants showed no systematic variation across the pH range tested, while the rate constants increased steeply with increasing pH up to an asymptote above pH 7.5. At low pH conditions, the gradient of a plot of log (rate constant) against pH approached a value of 2. DNA cleavage by EcoRV thus requires the de-protonation of two acidic groups. To determine whether aspartate 36 is one of the groups, mutants of EcoRV were made with other amino acid residues at position 36. Glutamate caused a partial loss of activity, while all other replacements gave near-zero activities. In contrast to wild-type EcoRV, the mutant with glutamate required the de-protonation of only one acidic group for DNA cleavage. A mechanism for EcoRV is proposed in which the water molecule that hydrolyses the phosphodiester bond is de-protonated by two Bronsted bases, probably the ionised forms of aspartate 36 and glutamate 45.  相似文献   

6.
Cells of a strain of Streptomyces sp. were incubated with an equivalent quantity of urate, xanthine, 6,8-dihydroxypurine or hypoxanthine in a medium deprived of other nitrogen source. The amount of uricase produced by these cells was shown to differ significantly, increasing in the following order of purine bases added to the medium: urate, xanthine, 6,8-dihydroxypurine and hypoxanthine. Of these was only urate indicated to be the inducer of uricase formation, and the difference in the quantity of uricase produced was found to be based on the duration of enzyme formation. The rate of uricase formation was essentially identical regardless of the purine bases supplied to cells.

Allantoin was accumulated in medium in remarkably different manners depending on the purine bases, which suggested the diversity in the mode of generation of urate in cells. Urate was generated at the slowest rate in the cells incubated with hypoxanthine, although the largest amount of uricase was produced, However, urate supplied to cells at the same rate but from medium failed to support the enzyme formation when the activity increased to a certain level. In order that the same amount of uricase was produced by the cells incubated with the different purine bases, the initial concentration of the purine bases should be raised so that they could remain in medium for the same incubation time.

Intracellular compartmentalization that might segregate endogenous and exogenous urate and might cause the difference in “effeciency” of these urate molecules as the inducer of uricase formation has been discussed.  相似文献   

7.
The MTH1 protein catalyzes hydrolysis of oxidatively damaged purine nucleotides including 8-hydroxy-dGTP to the monophosphates. The MTH1 protein seems to act as an important defense system against mutagenesis, carcinogenesis, and cell death induced by oxidized purine nucleotides. We previously reported that the functional groups at the 2- and 6-positions of the purine ring affect the recognition by the human MTH1 protein. 8-Hydroxy-dGTP and 8-hydroxy-dATP are substrates of MTH1, and both have the "7,8-dihydro-8-oxo structure." In this study, three nucleotide analogs containing this motif were examined. A synthetic purine analog containing the 7,8-dihydro-8-oxo structure and the 2-amino function (dJTP) was hydrolyzed to the monophosphate with high efficiency by MTH1. On the other hand, two analogs that lack the two-ring system of their bases [formamidopyrimidine-dGTP (FAPY-dGTP) and 2-OH-dYTP] were poor substrates. FAPY-dGTP is a mixture of conformers and was hydrolyzed more than ten-fold less efficiently than 8-hydroxy-dGTP. These results clarify the effects of the 2-amino group and the two-ring system of the purine base on the recognition by the human MTH1 protein.  相似文献   

8.
Theoretical considerations on the expected kinetics of the course of the Feulgen-Schiff reaction show that the leveling off of the first part of the Feulgen hydrolysis curve can be explained by the gradual conversion of deoxyribonucleic acid (DNA) to apurinic acid (APA). In addition, depolymerization of DNA caused by the acid used for hydrolysis can account for the decline after a maximum is reached in this curve. With the aid of polyacrylamide model films containing DNA, a detailed study was made both of the process of purine liberation which results in the formation of APA and of the depolymerization processes which cause losses of stainable material. The liberation of purine bases was analyzed by ultraviolet absorbance measurements and by gel chromatography of the neutralized hydrolysing acid. APA concentration was monitored by following the loss of ultraviolet absorbance associated with the purine losses. The depolymerization process was followed by phosphorus determinations. The experimental results were found to be in accordance with the kinetics expected from the theoretical model.  相似文献   

9.
The hydrolysis of 4-nitrophenyl acetate catalysed by cytoplasmic aldehyde dehydrogenase (EC 1.2.1.3) from sheep liver was studied by steady-state and transient kinetic techniques. NAD+ and NADH stimulated the steady-state rate of ester hydrolysis at concentrations expected on the basis of their Michaelis constants from the dehydrogenase reaction. At higher concentrations of the coenzymes, both NAD+ and NADH inhibited the reaction competitively with respect to 4-nitrophenyl acetate, with inhibition constants of 104 and 197 micron respectively. Propionaldehyde and chloral hydrate are competitive inhibitors of the esterase reaction. A burst in the production of 4-nitrophenoxide ion was observed, with a rate constant of 12 +/- 2s-1 and a burst amplitude that was 30% of that expected on the basis of the known NADH-binding site concentration. The rate-limiting step for the esterase reaction occurs after the formation of 4-nitrophenoxide ion. Arguments are presented for the existence of distinct ester- and aldehyde-binding sites.  相似文献   

10.
The rate constants and Km for the hydrolysis of the optically active nonglycosidic analogues of the CpA and C greater than p catalysed by RNase A and RNase BS-I were measured. The rate of hydrolysis of the model substrates in 10(5) and 10(3) slower that for the appropriate dinucleoside phosphate and nucleoside cyclophosphate. However, substitution of the relatively rigid ribofuranose ring with flexible alifatic chains is accompanied by little variation in binding constants. The analyses based on the single substrate system indicate that the observed difference in rate constants must be accounted for by a difference between the binding of the substrates in the transition state to the RNase active site. Consequently, the "rigidity" of the ribose rings in RNA leads to large decreases in the free energy of activation for the reactions catalysed by RNases.  相似文献   

11.
A series of purine bases and analogues were tested as substrates for trans-N-deoxyribosylase (EC 2.4.2.6). It was observed that the pyrimidine ring and its substituents on positions 1, 2 and 6, are of minor importance. On the other hand only a few modifications are tolerated on the imidazole moiety, as follows. 1. A tautomeric proton must be present on the imidazole ring. The "usual" shift is between position 9 and 7. 2. The position of the tautomeric proton governs the site of substitution. 3. For steric reasons no substituent is allowed on position 8.  相似文献   

12.
Crude extracts of Crithidia luciliae catalysed a deoxyribosyl transfer from purine deoxynucleosides to free purine bases. Fractionation of a 0-80% (NH4)2SO4 fraction from C. luciliae on DEAE-cellulose resulted in the separation of three nucleosidase activities. Two of these were ribonucleosidases, one specific for inosine, uridine and xanthosine and the other for inosine and guanosine, whereas the third activity was specific for purine deoxyribonucleosides. This pattern is similar to that found in Leishmania donovani. Significant deoxyribosyltransferase activity was, however, associated with the purine-2'-deoxyribonucleosidase from C. luciliae. The purine-2'-deoxyribonucleosidase was purified to homogeneity by a six-step procedure involving (NH4)2SO4 fractionation and chromatography on DEAE-cellulose, hydroxyapatite, Sephadex G-75, and a chromatofocusing resin. The purified enzyme migrated as a single band of 17 kDa on SDS/polyacrylamide gel electrophoresis. The enzyme catalysed the hydrolysis of deoxyinosine, deoxyguanosine and deoxyadenosine with Km values of 80 +/- 10.5 microM, 20.7 +/- 3.2 microM and 17.3 +/- 5.3 microM, respectively, and V values for these substrates in the ratio 1:0.5:0.39. The pH optimum for deoxyribosyl transfer from deoxyinosine to guanine was at pH 7.7, while deoxyinosine hydrolysis in the presence of guanine was optimal in the range pH 6-7. During the synthesis of deoxyinosine from hypoxanthine and deoxyadenosine two products were formed. One of these coeluted with deoxyinosine on HPLC, while the second was tentatively identified as the positional isomer, 7-(beta-D-2'-deoxyribofuranosyl)hypoxanthine.  相似文献   

13.
Interaction of the antitumour anthracyclines with mononucleotides and related compounds can be assessed through the perturbation of the spectral properties of the drugs. Purine-derived compounds induce spectral changes more efficiently than pyrimidine derivatives. No marked differences are observed when mono-, di- or triphosphate derivatives, deoxy forms, nucleosides or free nitrogen bases are used for the experiments. Visible absorbance data indicate the existence of a drug/purine nucleotide complex in solution. Assuming a simple equilibrium, this complex would be of low affinity (Keq 100 M-1). Circular dichroism spectra of daunomycin in the presence of ATP suggest that the resulting daunomycin/ATP complexes are not comparable to those formed by intercalation of the anthracycline into DNA. 31P-NMR of ATP in the presence of daunomycin does not support the notion that anthracycline/nucleotide complex formation involves interaction through the phosphate group(s) of the nucleotide. Analysis of the quenching of the drug's intrinsic fluorescence in the presence of nucleotides indicates a predominantly collisional, dynamic quenching mechanism. Values in the 2-6 mM and 85-100 mM range, respectively, are estimated for the reciprocal of the Stern-Volmer quenching constant for a variety of purine and pyrimidine derivatives. This indicates that purine derivatives are highly efficient quenchers of the fluorescence of anthracyclines, while pyrimidine derivatives are not. The fluorescence lifetime of daunomycin in the absence of quencher and the Stern-Volmer quenching constants obtained for different nucleotides are used to calculate the apparent bimolecular rate constants for collisions between fluorophore and quencher to occur. Values of (2-3) X 10(11) and 1 X 10(10) M-1 X s-1 are obtained, respectively, for purine and pyrimidine derivatives. This suggests a combination of static and dynamic quenching processes for purine compounds, which is consistent with the drug/purine nucleotide complex formation detected by visible absorbance. Because of the high intracellular concentration of certain nucleotides, particularly ATP, the above processes are predicted to be highly significant 'in vivo'.  相似文献   

14.
The phase transfer method was applied to perform the nucleophilic substitution of 2,6-dichloropurines by modified arylalkyl alcohol or phenols. Since under these conditions only the 6-halogen is exchanged, this method gives 2-chloro-6-aryloxy- and 2-chloro-6-arylalkoxy-purines. 2-Chloro-6-benzylthiopurine was synthesized by alkylation of 2-chloro-6-thiopurine with benzyl bromide. The stereoisomers of 2-chloro-6-(1-phenyl-1-ethoxy)purine were obtained from R- and S-enantiomers of sec.-phenylethylalcohol and 2,6-dichloropurine. All derivatives were tested for inhibition with purified hexameric E. coli purine nucleoside phosphorylase (PNP). For analogues showing IC50 < 10 microM, the type of inhibition and inhibition constants were determined. In all cases the experimental data were best described by the mixed-type inhibition model and the uncompetitive inhibition constant, Kiu, was found to be several-fold lower than the competitive inhibition constant, Kic. This effect seems to be due to the 6-aryloxy- or 6-arylalkoxy substituent, because a natural PNP substrate adenine, as well as 2-chloroadenine, show mixed type inhibition with almost the same inhibition constants Kiu and Kic. The most potent inhibition was observed for 6-benzylthio-2-chloro-, 6-benzyloxy-2-chloro-, 2-chloro-6-(2-phenyl-1-ethoxy), 2-chloro-6-(3-phenyl-1-propoxy)- and 2-chloro-6-ethoxypurines (Kiu = 0.4, 0.6, 1.4, 1.4 and 2.2 microM, respectively). The R-stereoisomer of 2-chloro-6-(1-pheny-1-ethoxy)purine has Kiu = 2.0 microM, whereas inhibition of its S counterpart is rather weak (IC50 > 12 microM). More rigid (e.g. phenoxy-), non-planar (cyclohexyloxy-), or more bulky (2,4,6-trimethylphenoxy-) substituents at position 6 of the purine base gave less potent inhibitors (IC50 = 26, 56 and > 100 microM, respectively). The derivatives are selective inhibitors of hexameric "high-molecular mass" PNPs because no inhibitory activity vs. trimeric Cellulomonas sp. PNP was detected. By establishing the ligand-dependent stabilization pattern of the E. coli PNP it was shown that the new derivatives, similarly as the natural purine bases, are able to form a dead-end ternary complex with the enzyme and orthophosphate. It was also shown that the derivatives are substrates in the reverse synthetic direction catalyzed by E. coli PNP.  相似文献   

15.
Nucleic acid bases such as adenine and uracil, and nitrobenzeneboronic acid substituted silicas were prepared by the reaction of chloromethylbenzene substituted silica with adenine sodium salt and trimethylsilylated uracil, and nitration of benzeneboronic acid substituted silica, respectively. From the results of HPLC of nucleosides and N-ethyl derivatives of nucleic acid bases using modified silicas, hydrophobic base stacking interaction, selective hydrogen bonding interaction between purine and pyrimidine bases, and reversible cyclic boronate ester formation between diols of nucleosides with boronic acid were effective for the separation of nucleic acid related compounds. Moreover, association constants for hydrogen bonding formation of nucleic acid bases were estimated.  相似文献   

16.
Geobacillus stearothermophilus T-6 encodes for a beta-xylosidase (XynB2) from family 52 of glycoside hydrolases that was previously shown to hydrolyze its substrate with net retention of the anomeric configuration. XynB2 significantly prefers substrates with xylose as the glycone moiety and exhibits a typical bell-shaped pH dependence curve. Binding properties of xylobiose and xylotriose to the active site were measured using isothermal titration calorimetry (ITC). Binding reactions were enthalpy driven with xylobiose binding more tightly than xylotriose to the active site. The kinetic constants of XynB2 were measured for the hydrolysis of a variety of aryl beta-D-xylopyranoside substrates bearing different leaving groups. The Br?nsted plot of log k(cat) versus the pK(a) value of the aglycon leaving group reveals a biphasic relationship, consistent with a double-displacement mechanism as expected for retaining glycoside hydrolases. Hydrolysis rates for substrates with poor leaving groups (pK(a) > 8) vary widely with the aglycon reactivity, indicating that, for these substrates, the bond cleavage is rate limiting. However, no such dependence is observed for more reactive substrates (pK(a) < 8), indicating that in this case hydrolysis of the xylosyl-enzyme intermediate is rate limiting. Secondary kinetic isotope effects suggest that the intermediate breakdown proceeds with modest oxocarbenium ion character at the transition state, and bond cleavage proceeds with even lower oxocarbenium ion character. Inhibition studies with several gluco analogue inhibitors could be measured since XynB2 has low, yet sufficient, activity toward 4-nitrophenyl beta-D-glucopyranose. As expected, inhibitors mimicking the proposed transition state structure, such as 1-deoxynojirimycin, bind with much higher affinity to XynB2 than ground state inhibitors.  相似文献   

17.
Hydrolysis of cyclic AMP and cyclic GMP analogues by a purified cGMP-stimulated phosphodiesterase from bovine adrenal tissue was investigated by reversed-phase HPLC. The results indicate that both a negative charge and an equatorial oxygen atom located at the cyclic phosphate residue are absolute requirements for the process of hydrolysis. Other substituents only gradually decreased the apparent hydrolytic activity. C-8-substituted derivatives were generally poor substrates due to the limited ability of these compounds to rotate freely around the glycosidic bond. While C-6- and 0-2'-substituted analogues carrying bulky substituents were also poorly hydrolysed, all other derivatives, including different C-2-, C-6-, 0-3'- and 0-5'-modified cyclic nucleotides, were good substrates. We consistently observed that cyclic GMP and cyclic GMP analogues were better hydrolysed than the corresponding cyclic AMP analogues. Hydrolysis was correlated with neither the hydrogen bond donor/acceptor abilities nor the hydrophobicity of selected cyclic nucleotide analogues. Based on quantum-chemical calculations of the size and direction of the dipole moments of different purine bases, we propose that the polarization of inducible amino acid side-chains within the binding site is involved in the differential binding of adenine-derived and guanine-derived nucleotides. However, the size of the dipole moment alone is not sufficient to explain the observed cGMP-preference. Rather, the direction of the polarization power relative to the other molecular structures involved in binding and hydrolysis seems to be the molecular mechanism by which the enzyme is able to discriminate between cAMP- and cGMP-like structures.  相似文献   

18.
Whole cells and isolated membranes of the marine bacterium MB22 converted nucleotides present in the external medium rapidly into nucleosides and then into bases. Nucleosides and purine bases formed were taken up by distinct transport systems. We found a high-affinity common transport system for adenine, guanine, and hypoxanthine, with a Km of 40 nM. This system was inhibited noncompetitively by purine nucleosides. In addition, two transport systems for nucleosides were present: one for guanosine with a Km of 0.8 microM and another one for inosine and adenosine with a Km of 1.4 microM. The nucleoside transport systems exhibited both mixed and noncompetitive inhibition by different nucleosides other than those translocated; purine and pyrimidine bases had no effect. The transport of nucleosides and purine bases was inhibited by dinitrophenol or azide, thus suggesting that transport is energy dependent. Inside the cell all of the substrates were converted mainly into guanosine, xanthine, and uric acid, but also anabolic products, such as nucleotides and nucleic acids, could be found.  相似文献   

19.
Association between calf spleen purine nucleoside phosphorylase and a series of phosphonylalkoxyalkyl derivatives of purine bases was studied by inhibition kinetics and fluorimetric titrations. Dissociation constants, determined by fluorimetric titration in phosphate-free conditions, were lower than inhibition constants in 1 mM phosphate, and inhibition was still weaker in 50 mM phosphate, in accord with the postulated bisubstrate analogue character of this class of inhibitors.  相似文献   

20.
The reaction of triethanolamine (TEA) with active substrates—p-nitrophenyl esters and cinnamoyl imidazole (CI)—is catalyzed by divalent heavy metal ions. With Hg2+, rate enhancements of 100–1000 (depending on the substrate) were observed, the overall rate constants of substrate decomposition thus exceeding those of spontaneous hydrolysis up to 100,000-fold. The predominant active species at low L:M ratio was found to be the Hg-(TEA)2 complex. The dependence of the reaction rate upon excess of amino alcohol—at constant Hg2+ concentration—is attributable to formation of another active complex—Hg-(TEA)3.The high reactivity of the system is due to the alcoholate group of metal-bound TEA, whose pK has been lowered by the proximity of the metal ion. This labile nucleophilic alcoholate attacks the substrate causing its alcoholysis and forming O-acyl-TEA. The lability of the metal-alcoholate bond can be enhanced by low concentrations of halide ions, thus causing up to 5-fold additional increase in alcoholysis rate. Higher halide ion concentrations cause inhibition, probably due to formation of inactive HgX2 molecules.Presumably an important role of the metal ion in metalloenzymes is to affect the decrease in the pK value of a reactive group so that it can exhibit activity under physiological conditions.  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