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.
The effects of alpha-, beta- and gamma-cyclodextrins on the amylose and maltopentaose hydrolysis catalysed by porcine pancreatic alpha-amylase (PPA) were investigated. The results of the statistical analysis performed on the kinetic data using the general initial velocity equation of a one-substrate reaction in the presence of one inhibitor indicate that the type of inhibition involved depends on the substrate used: the inhibition of amylose hydrolysis by alpha-, beta- and gamma-cyclodextrin is of the competitive type, while the inhibition of maltopentaose hydrolysis is of the mixed noncompetitive type. Consistently, the Lineweaver-Burk plots intersect on the vertical axis when amylose is used as the substrate, while in the case of maltopentaose, the intersection occurs at a point located in the second quadrant. The inhibition of the hydrolysis therefore involves only one abortive complex, PPA-cyclodextrin, when amylose is used as the substrate, while two abortive complexes, PPA-cyclodextrin and PPA-maltopentaose-cyclodextrin, are involved with maltopentaose. The mixed noncompetitive inhibition thus shows the existence of one accessory binding site. In any case, only one molecule of inhibitor binds to PPA. In line with these findings, the difference spectra of PPA produced by alpha-, beta- and gamma-cyclodextrin indicate that binding occurs at a tryptophan and a tyrosine residue. The corresponding dissociation constants and the inhibition constants obtained using the kinetic approach are in the same range (1.2-7 mM). The results obtained here on the inhibition of maltopentaose hydrolysis by cyclodextrin are similar to those previously obtained with acarbose as the inhibitor [Alkazaz, M., Desseaux, V., Marchis-Mouren, G., Prodanov, E. & Santimone, M. (1998) Eur. J. Biochem. 252, 100-107], but differ from those obtained with amylose as the substrate and acarbose as inhibitor [Alkazaz, M., Desseaux, V., Marchis-Mouren, G., Payan, F., Forest, E. & Santimone, M. (1996) Eur. J. Biochem. 241, 787-796]. It is concluded that the hydrolysis of both long and short chain substrates requires at least one secondary binding site, including a tryptophan residue.  相似文献   

2.
1. A large fluorescence enhancement of 2-p-toluidinylnaphthalene-6-sulfonate (TNS) observed in the presence of amylose was utilized to monitor quantitatively the time course of porcine pancreatic alpha-amylase [EC 3.2.1.1] (PPA)-catalyzed hydrolysis of amylose with a number-average degree of polymerization of 16.8. 2. The slope of the plot of decrease in the relative fluorescence intensity of the TNS-amylose system (termed as the fluorescence value) versus the number of linkages hydrolyzed (reducing value) (Kondo, H. et al. (1977) Agric. Biol. Chem. 41, 631-634) in the course of PPA-catalyzed hydrolysis was shown to be useful to describe the degree of "multiple attack," which is defined by the number of reattacks on a long chain substrate molecular per one encounter of the enzyme and the substrate. A parameter gamma was defined as the ratio of the reciprocal of the slopes obtained at each pH to that at pH 10.5, where the multiple attack is not operating. 3. The gamma versus pH profile gave an apparent pK value of about 9, indicating that some ionizable groups participate in the multiple attack mechanism. 4. Based on a reaction scheme involving a "sliding" of the substrate molecule on the enzyme, which may contribute to the multiple attack mechanism, besides binding, dissociation, and cleavage steps of the substrate, and on the assumption of the steady state for the enzyme-substrate complex, rate equations were obtained to describe the time course of hydrolysis of a linear substrate. The product distribution with the progress of the reaction can be calculated theoretically, and is dependent on the number of multiple attack and the mode of sliding. The number of multiple attack can be estimated from this distribution, and the fluorescence value can be calculated theoretically by combining the product distribution with the relative efficiency of fluorescence intensity of each maltooligosaccharide (Nakatani, H. et. al. (1977) Biopolymers 16, 2363-2370). By comparing the experimental data with the theoretical ones, it was suggested that the multiple attack occurs through the sliding by maltose unit of the retained fragment on the enzyme, which is one of the fragments produced by the initial cleavage of the substrate molecule. 5. It was found that anions (chloride, bromide, and nitrate ions) which critically affect the enzyme activity have no effect on the degree of multiple attack.  相似文献   

3.
A modified amylose containing 10% of tritiated D-allose residues has been hydrolyzed by porcine pancreatic alpha amylase (PPA). This reaction produced a number of radioactive oligosaccharides of low molecular weight, including modified mono-, di-, and tri-saccharides, as well as larger products. Analysis of these products by chemical and enzymic methods identified D-allose, two isomers of modified maltose, and isomers of modified maltotriose. These results may be interpreted in terms of current PPA models to indicate that D-allose residues may be productively bound at all five subsites of the active site of the enzyme. The distribution of modified residues in these products, however, further suggests that productive binding of D-allose at the subsite where catalytic attack occurs (subsite 3) is less favorable than binding of D-glucose. These results are compared with results of a series of PPA substrates having modifications at C-3 and at other positions. Trends observed in enzyme hydrolysis of these modified substrates reflect factors that contribute to PPA catalysis, with respect to steric, electronic, and hydrogen-bonding interactions between enzyme and substrate.  相似文献   

4.
Amylopectin fine structures were studied following limited hydrolysis of gelatinised waxy maize starch by amylases with a different level of inner chain attack (LICA). This was done by size exclusion chromatography as well as by debranching the (partially hydrolysed) amylopectin samples and studying the size distributions of the released chains. α-Amylases from Bacillus amyloliquefaciens and Aspergillus oryzae, with a relatively high LICA, drastically altered amylopectin chain length distribution and reduced the amylopectin molecular size (MS) significantly even at a low to moderate degree of hydrolysis (DH). Porcine pancreatic α-amylase (PPA), with a rather low LICA but a high multiple attack action on amylose, reduced the amylopectin MS much slower. Following hydrolysis by PPA to a DH of 10% and enzymic debranching of the amylopectin residue, several subpopulations of chains consisting of 2-12 glucose units were detected, indicating a multiple attack action on the amylopectin side chains. During the early stages of hydrolysis, the maltogenic Bacillus stearothermophilus α-amylase (BStA) preferentially hydrolysed the exterior chains of amylopectin. However, during the later phases, BStA also hydrolysed inner chains, presumably with a high multiple attack action. The present results clearly show that different enzymes can be used for (limited) conversion of amylopectin into structures differing in molecular weight and chain length distributions.  相似文献   

5.
Two inhibitors, acarbose and cyclodextrins (CD), were used to investigate the active site structure and function of barley alpha-amylase isozymes, AMY1 and AMY2. The hydrolysis of DP 4900-amylose, reduced (r) DP18-maltodextrin and maltoheptaose (catalysed by AMY1 and AMY2) was followed in the absence and in the presence of inhibitor. Without inhibitor, the highest activity was obtained with amylose, kcat/Km decreased 103-fold using rDP18-maltodextrin and 10(5) to 10(6)-fold using maltoheptaose as substrate. Acarbose is an uncompetitive inhibitor with inhibition constant (L1i) for amylose and maltodextrin in the micromolar range. Acarbose did not bind to the active site of the enzyme, but to a secondary site to give an abortive ESI complex. Only AMY2 has a second secondary binding site corresponding to an ESI2 complex. In contrast, acarbose is a mixed noncompetitive inhibitor of maltoheptaose hydrolysis. Consequently, in the presence of this oligosaccharide substrate, acarbose bound both to the active site and to a secondary binding site. alpha-CD inhibited the AMY1 and AMY2 catalysed hydrolysis of amylose, but was a very weak inhibitor compared to acarbose.beta- and gamma-CD are not inhibitors. These results are different from those obtained previously with PPA. However in AMY1, as already shown for amylases of animal and bacterial origin, in addition to the active site, one secondary carbohydrate binding site (s1) was necessary for activity whereas two secondary sites (s1 and s2) were required for the AMY2 activity. The first secondary site in both AMY1 and AMY2 was only functional when substrate was bound in the active site. This appears to be a general feature of the alpha-amylase family.  相似文献   

6.
K Ishikawa  I Matsui  K Honda  H Nakatani 《Biochemistry》1990,29(30):7119-7123
Porcine pancreatic alpha-amylase (EC 3.2.1.1, abbreviated as PPA) hydrolyzes alpha-D-(1,4) glucosidic bonds in starch and amylose at random, and the optimum pH for the substrates is 6.9. The optimum pH, however, shifted to 5.2 for the hydrolytic reaction of low molecular weight oligosaccharide substrates such as p-nitrophenyl alpha-D-maltoside, gamma-cyclodextrin, maltotetaitol, and maltopentaitol. The optimum pH for the oligosaccharides consisting of more than five glucose residues, such as maltopentaose and maltohexaitol, was 6.9. From the analysis of the hydrolysates, it was clear that the shift of the optimum pH occurred only when the fifth subsite of PPA in the productive binding modes was occupied by a glucosyl residue of the substrates. The value of Km was independent of pH between 4 and 10 but that of kcat was dependent on pH. The pH profiles of kcat for the above substrates did not fit a simple bell-shaped curve predicted by a two-catalytic-group mechanism. Instead, they were well analyzed theoretically by three pK values and two intrinsic kcat values. Enthalpy changes for the three pK's (4.90, 5.35, and 8.55 at 30 degrees C) were determined from the temperature dependence of pH profiles for maltopentaitol and maltohexaitol to be 0.0, 2.87, and 7.33 kcal/mol, respectively. These results indicate that productive binding modes of the substrates directly affect the catalytic function of the enzyme. From the present thermodynamic analysis and reported three dimensional structure at the active site of PPA [Buisson, G. (1987) EMBO J. 6, 3909-3916], one can assume that a histidyl residue (101, 201, or 299) acts as a proton donor and two carboxyl groups (Asp 197, Glu 233, or Asp 300) act as proton donors or acceptors, and the productive binding mode covering the fifth subsite changes configurations between the catalytic residues and the glucosidic bond hydrolyzed and modulates kinetic parameters depending on pH.  相似文献   

7.
Conversion of iron(II) verdoheme to iron biliverdin in the presence of OH(-) was investigated using B3LYP method. Both 3-21G and 6-31G* basis sets were employed for geometry optimization calculation as well as energy stabilization estimation. Calculation at 6-31G* level was found necessary for a correct spin state estimation of the iron complexes. Two possible pathways for the conversion of iron verdoheme to iron biliverdin were considered. In one path the iron was six-coordinate while in the other it was considered to be five-coordinate. In the six-coordinated pathway, the ground state of bis imidazole iron verdoheme is singlet while that for open chain iron biliverdin it is triplet state with 4.86 kcal/mol more stable than the singlet state. The potential energy surface suggests that a spin inversion take place during the course of reaction after TS. The ring opening process in the six-coordinated pathway is in overall -2.26 kcal/mol exothermic with a kinetic barrier of 9.76 kcal/mol. In the five-coordinated pathway the reactant and product are in the ground triplet state. In this path, hydroxyl ion attacks the iron center to produce a complex, which is only 1.59 kcal/mol more stable than when OH(-) directly attacks the macrocycle. The activation barrier for the conversion of iron hydroxy species to the iron biliverdin complex by a rebound mechanism is estimated to be 32.68 kcal/mol. Large barrier for rebound mechanism, small barrier of 4.18 kcal/mol for ring opening process of the hydroxylated macrocycle, and relatively same stabilities for complexes resulted by the attack of nucleophile to the iron and macrocycle indicate that five-coordinated pathway with direct attack of nucleophile to the 5-oxo position of macrocycle might be the path for the conversion of verdoheme to biliverdin.  相似文献   

8.
The action pattern of several amylases was studied at 35, 50, and 70 degrees C using potato amylose, a soluble (Red Starch) and insoluble (cross-linked amylose) chromophoric substrate. With potato amylose as substrate, Bacillus stearothermophilus alpha-amylase (BStA) and porcine pancreatic alpha-amylase displayed a high degree of multiple attack (DMA, i.e., the number of bonds broken during the lifetime of an enzyme-substrate complex minus one), the fungal alpha-amylase from Aspergillus oryzae a low DMA, and the alpha-amylases from B. licheniformis, Thermoactinomyces vulgaris, B. amyloliquifaciens, and B. subtilis an intermediate DMA. These data are discussed in relation to structural properties of the enzymes. The level of multiple attack (LMA), based on the relation between the drop in iodine binding of amylose and the increase in total reducing value, proved to be a good alternative for DMA measurements. The LMA of the endo-amylases increased with temperature to a degree depending on the amylase. In contrast, BStA showed a decreased LMA when temperature was raised. Furthermore, different enzymes had different activities on Red Starch and cross-linked amylose. Hence, next to the temperature, the action pattern of alpha-amylases is influenced by structural parameters of the starch substrate.  相似文献   

9.
Porcine pancreatic α-amylase (PPA) binds to N-linked glycans of glycoproteins (Matsushita, H., Takenaka, M., and Ogawa, H. (2002) J. Biol Chem., 277, 4680-4686). Immunostaining revealed that PPA is located at the brush-border membrane (BBM) of enterocytes in the duodenum and that the binding is inhibited by mannan but not galactan, indicating that PPA binds carbohydrate-specifically to BBM. The ligands for PPA in BBM were identified as glycoprotein N-glycans that are significantly involved in the assimilation of glucose, including sucrase-isomaltase (SI) and Na(+)/Glc cotransporter 1 (SGLT1). Binding of SI and SGLT1 in BBM to PPA was dose-dependent and inhibited by mannan. Using BBM vesicles, we found functional changes in PPA and its ligands in BBM due to the N-glycan-specific interaction. The starch-degrading activity of PPA and maltose-degrading activity of SI were enhanced to 240 and 175%, respectively, while Glc uptake by SGLT1 was markedly inhibited by PPA at high but physiologically possible concentrations, and the binding was attenuated by the addition of mannose-specific lectins, especially from Galanthus nivalis. Additionally, recombinant human pancreatic α-amylases expressed in yeast and purified by single-step affinity chromatography exhibited the same carbohydrate binding specificity as PPA in binding assays with sugar-biotinyl polymer probes. The results indicate that mammalian pancreatic α-amylases share a common carbohydrate binding activity and specifically bind to the intestinal BBM. Interaction with N-glycans in the BBM activated PPA and SI to produce much Glc on the one hand and to inhibit Glc absorption by enterocytes via SGLT1 in order to prevent a rapid increase in blood sugar on the other.  相似文献   

10.
Mutagenesis of Bacteroides thetaiotaomicron with the transposon Tn4351 produced five classes of mutants that were not able to grow on amylose or amylopectin. These classes of mutants differed in their ability to grow on maltoheptaose (G7) and in the level of starch-degrading enzymes produced when bacteria were grown on maltose. All of the mutants were deficient in starch binding. Since one class of mutants retained normal levels of starch-degrading enzymes, this indicates that binding of the starch molecule by a cell surface receptor is necessary for starch utilization by B. thetaiotaomicron. Analysis of a starch-negative mutant that grew on G7 indicated that B. thetaiotaomicron possessed two starch-binding components or sites. One component (site A), apparently missing in this mutant, had an absolute preference for larger starch oligomers, whereas the other component (site M) also had a high affinity for maltodextrins (G4 through G7). Mutants not able to grow on maltodextrins (greater than G4) probably lacked both of these binding components. Only one class of mutants did not grow normally on maltose, but instead had a 4- to 5-h lag on maltose and a slower growth rate than the wild type. This class of mutants did not produce any of the starch-degrading enzymes or bind starch, even when growing on maltose. Such a phenotype probably resulted from transposon inactivation of a central regulatory gene or a gene encoding an enzyme that produces the inducer. The fact that both the degradative enzymes and the starch-binding activity were affected in this mutant indicates that genes encoding the cell surface starch-binding site are under the same regulatory control as genes encoding the enzymes.  相似文献   

11.
Isolation of a cDNA encoding a protease from Perinereis aibuhitensis Grube   总被引:2,自引:0,他引:2  
The cDNA encoding a protease of Perinereis aibuhitensis Grube (PPA) was cloned. The deduced amino acid sequence analysis showed that the protein had 49% identity to the C-terminal amino acid 169-246 of serine protease of Heterodera glycines. Northern blotting analysis indicated that the cDNA could hybridize with mRNA of approximately 260 bases isolated from the marine earthworm. The cDNA was amplified by polymerase chain reaction and cloned into pMAL-p2 to construct expression vector pMAL-PPA. pMAL-PPA was introduced into Escherichia coli BL21(DE3) and overexpression of PPA fused with maltose binding protein was achieved by isopropyl-β-D-thiogalactopyranoside induction. The fusion protein was purified by affinity chromatography on an amylose resin column and ion-exchange chromatography on a diethylaminoethyl-Sepharose 4B column. Rabbits were immunized with the purified protein and antiserum was prepared. The antibody could react with a protein of approximately 9 kDa extracted from the marine earthworm as shown by Western blotting analysis. The activity analysis of the recombinant PPA suggested that it was probably a plasminogen activator.  相似文献   

12.
Porcine pancreatic alpha-amylase (EC 3.2.1.1; abbreviated PPA), which hydrolyzes alpha-D-(1,4) glucosidic bonds in starch and amylose, displays an optimum at pH 6.9 for the majority of substrates. The optimum pH, however, shifted to 5.2 for the hydrolysis of some low molecular substrates (Ishikawa, K., et al., 1990, Biochemistry 29, 7119-7123). Details of the substrate-dependent shift of the optimum pH in PPA were studied by use of a series of maltooligosaccharides with 14C-labeled reducing end glucose as substrates. The optimum pH for maltotriose was 5.2, whereas that for maltopentaose and maltohexaose was unchanged at pH 6.9. The pH profile for the intermediate size substrate maltotetraose showed abnormality; the apparent optimum pH was broadened between 5.5 and 6.5 and the bond cleavage pattern depended on pH, unlike that for the other substrates examined. These results were independent of either buffer systems or substrate concentration. Analyses of the hydrolysates of the maltooligosaccharides revealed that the shift of the optimum pH to the neutral region occurred only when the fifth subsite of PPA in the productive binding modes was occupied by a glucosyl residue of a substrate. The three-catalytic residue model of PPA deduced from the analysis of the hydrolysis of some modified maltooligosaccharides (p-nitrophenyl-alpha-D-maltoside, gamma-cyclodextrin, maltopentaitol, and maltohexaitol) (Ishikawa, K., et al., 1990, Biochemistry 29, 7119-7123) was successfully adapted to the linear maltooligosaccharides used in this work. These results indicate that the different productive binding modes of the linear oligosaccharide substrates affect directly the catalytic power and the optimum pH of PPA.  相似文献   

13.
An atomistic biophysical model permitting the calculation of initial attacks to a 38-bp representation of B-DNA base moieties by water radicals is presented. This model is based on a previous radiation damage model developed by Aydogan et al. (Radiat. Res. 157, 38-44, 2002). Absolute efficiencies for radical attack to the 38-bp DNA molecule are calculated to be 41, 0.8 and 15% for hydroxyl radical ((.)OH), hydrogen radical (H(.)), and hydrated electron (e(aq))(,) respectively. Among the nucleobases, guanine is found to have the highest percentage (.)OH attack probability at 36%. Adenine, cytosine and thymine moieties have initial attack probabilities of 24, 18 and 22%, respectively. A systematic study is performed to investigate (.)OH attack probabilities at each specified attack site in four molecular models including free bases, single nucleotides, single base pairs, and the central eight base pairs of the 38-bp DNA molecule. Cytosine is the free base moiety for which the closest agreement is observed between the model prediction and the experimental data. The initial (.)OH attack probabilities for cytosine as the free base are calculated to be 72 and 28%, while experimental data are reported at 87 and 13% for the C5 and C6 positions on the base, respectively. In this study, we incorporated atomic charges to scale the site-specific (.)OH reaction rates at the individual atomic positions on the pyrimidine and purine bases. Future updates to the RIDNA model will include the use of electron densities to scale the reaction rates. With respect to reactions of the aqueous electron with DNA, a comparison of the initial distribution of electron attack sites calculated in this study and experimental results suggests an extremely rapid and extensive redistribution of the e(-)(aq) after their initial reactions with DNA.  相似文献   

14.
The inhibitory effect of 0.19 alpha-amylase inhibitor (0.19 AI) from wheat kernel on the porcine pancreas alpha-amylase (PPA)-catalyzed hydrolysis of p-nitrophenyl-alpha-D-maltoside (pNP-G2) was examined. 0.19 AI is a homodimer of 26.6 kDa with 13.3-kDa subunits under the conditions used. The elution behaviors in gel filtration HPLC of PPA and 0.19 AI indicated that a PPA molecule bound with a 0.19 AI molecule (homodimer) at a molar ratio of 1:1. 0.19 AI inhibited PPA activity in a competitive manner with an inhibitor constant, K(i), of 57.3 nM at pH 6.9, 30 degrees C, and the binding between them was found to be endothermic and entropy-driven. The activation energy for the thermal inactivation of 0.19 AI was determined to be 87.0 kJ/mol, and the temperature, T(50), giving 50% inactivation in a 30-min incubation at pH 6.9 was 88.1 degrees C. The high inhibitory activity of 0.19 AI against PPA and its high thermal stability suggest its potential for use in the prevention and therapy of obesity and diabetes.  相似文献   

15.
Total starch, amylose content and amylose-included lipid phosphorus and lysophosphatidylcholine (LPC) were measured in normal Glacier (G) and Hi Amylose Glacier (HA) barley varieties during germination. From days three to six, alkaline and acidic lysophospholipase (LPL) activities in the starchy endosperm were measured and the distribution of these activities between a soluble and particulate form determined. During germination the amylose content of the starches increases as the total starch levels decline. The starch-bound LPC and lipid phosphorus disappear at the same rate between days three and six in both barley varieties, indicating no discrimination among the different lipid-included amylose population for degradation. However, both lipid phosphorus and LPC disappear more rapidly in the G than in the HA variety. This is presumably due to the slightly larger content of LPC per mg amylose of the G than of the HA variety, equivalent to 134 and 150 anhydroglucose residues per lipid molecule in G and HA, respectively. There is no increase in starch-bound lipid phosphorus or LPC expressed as nmol of phosphorus or LPC per mg amylose as amylose content declines, indicating no selective resistance of lipid-included amylose to degradation. The alkaline and acidic LPC activities in each variety increase 2–4-fold between days four and five. In both varieties ca 30% of the acidic LPL and ca 50–60% of the alkaline LPL is particulate from days three to six. No correlation can be made between the content of amylose or amylose-included lipid and particulate LPL activity. However, the possibility that particulate LPL activity is associated with specific populations of residual amylose-included lipid molecules cannot be excluded.  相似文献   

16.
The G values for single-strand breaks G(ssb) in polyuridylic acid (poly U) have been measured by low-angle laser light scattering in aqueous solutions under various conditions (e.g. in the presence of N2O, Ar and t-butanol). In N2O-saturated solutions at room temperature and pH 5.6, the G(ssb) is 2.3. The efficiency of ssb formation was found to be 41 per cent for OH radicals, 19 per cent for H atoms and congruent to zero for e-aq. On the basis of 20 per cent and less than 5 per cent attack on the sugar moiety by OH radicals and H atoms, respectively, the large G(ssb) values obtained cannot be explained solely as resulting from radicals produced by reaction of OH radicals and H atoms on the sugar moiety. It is therefore proposed that base radicals produced by the reaction of OH radicals or H atoms with the uracil moiety can also lead to chain break formation in poly U via radical transfer to the sugar moiety.  相似文献   

17.
Little computational or experimental information is available on site-specific hydroxyl attack probabilities to DNA. In this study, an atomistic stochastic model of OH radical reactions with DNA was developed to compute relative OH attack probabilities at individual deoxyribose hydrogen atoms. A model of the self-complementary decamer duplex d(CCAACGTTGG) was created including Na(+) counter ions and the water molecules of the first hydration layer. Additionally, a method for accounting for steric hindrance from nonreacting atoms was implemented. The model was then used to calculate OH attack probabilities at the various C-H sites of the sugar moiety. Results from this computational model show that OH radicals exhibit preferential attack at different deoxyribose hydrogens, as suggested by their corresponding percentage solvent-accessible surface areas. The percentage OH attack probabilities for the deoxyribose hydrogens [1H(5')+2H(5'), H(4'), H(3'), 1H(2')+2H(2'), H(1')] were calculated as approximately 54.6%, 20.6%, 15.0%, 8.5% and 1.3%, respectively, averaged across the sequence. These results are in good agreement with the latest experimental site-specific DNA strand break data of Balasubramanian et al. [Proc. Natl. Acad. Sci. USA 95, 9738-9742 (1998)]. The data from this stochastic model suggest that steric hindrance from nonreacting atoms significantly influences site-specific hydroxyl radical attack probabilities in DNA. A number of previous DNA damage models have been based on the assumption that C(4') is the preferred site, or perhaps the only site, for OH-mediated DNA damage. However, the results of the present study are in good agreement the experimental results of Balasubramanian et al. in which OH radicals exhibit preferential initial attack at sugar hydrogen atoms in the order 1H(5')+2H(5') > H(4') > H(3') > 1H(2')+2H(2') > H(1').  相似文献   

18.
19.
Translational initiation factor 3 (IF3) is an RNA helix destabilizing protein which interacts with strongly conserved sequences in 16S rRNA, one at the 3' terminus and one in the central domain. It was therefore of interest to identify particular residues whose exposure changes upon IF3 binding. Chemical and enzymatic probing of central domain nucleotides of 16S rRNA in 30S ribosomal subunits was carried out in the presence and absence of IF3. Bases were probed with dimethyl sulfate (DMS), at A(N-1), C(N-3), and G(N-7), and with N-cyclohexyl-N'-[2-(N-methyl-4-morpholinio)ethyl] carbodiimide p-toluenesulfonate (CMCT), at G(N-1) and U(N-3). RNase T1 and nuclease S1 were used to probe unpaired nucleotides, and RNase V1 was used to monitor base-paired or stacked nucleotides. 30S subunits in physiological buffers were probed in the presence and absence of IF3. The sites of cleavage and modification were detected by primer extension. IF3 binding to 30S subunits was found to reduce the chemical reactivity and enzymatic accessibility of some sites and to enhance attack at other sites in the conserved central domain of 16S rRNA, residues 690-850. IF3 decreased CMCT attack at U701 and U793 and V1 attack at G722, G737, and C764; IF3 enhanced DMS attack at A814 and V1 attack at U697, G833, G847, and G849. Many of these central domain sites are strongly conserved and with the conserved 3'-terminal site define a binding domain for IF3 which correlates with a predicted cleft in two independent models of the 30S ribosomal subunit.  相似文献   

20.
Hydroxymethyl radicals .CH2OH, generated by the radiolysis of methanol (0.5 mol dm-3) in N2O-saturated aqueous solutions, were reacted with 1,3-dimethyluracil or 1,3-dimethylthymine (10(-3) mol dm-3). The products were identified and their G values determined. It has been concluded that in 1,3-dimethyluracil .CH2OH attack occurs only at C(6) while in 1,3-dimethylthymine there is partitioning between addition (two-thirds) and H-abstraction from the C(5)-methyl group (one-third). A rate constant for CH2OH addition to 1,3-dimethyluracil of about 10(4) dm3 mol-1 s-1 is estimated. Complexities that may arise in the radiolysis of pyrimidines such as 1,3-dimethylthymine, apparently as a consequence of the formation of 5-alkylidenepyrimidines, are discussed. A value of 0.15 has been estimated for the disproportionation/combination ratio for the hydroxymethyl radical self-termination reaction.  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