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 soluble form of penicillin-binding protein 3 (PBP 3) from Neisseria gonorrhoeae was expressed and purified from Escherichia coli and characterized for its interaction with beta-lactam antibiotics, its catalytic properties with peptide and peptidoglycan substrates, and its role in cell viability and morphology. PBP 3 had an unusually high k(2)/K' value relative to other PBPs for acylation with penicillin (7.7 x 10(5) M(-1) s(-1)) at pH 8.5 at 25 degrees C and hydrolyzed bound antibiotic very slowly (k(3) < 4.6 x 10(-5) s(-1), t(1/2) > 230 min). PBP 3 also demonstrated exceptionally high carboxypeptidase activity with a k(cat) of 580 s(-1) and a k(cat)/K(m) of 1.8 x 10(5) M(-1) s(-1) with the substrate N(alpha)-Boc-N(epsilon)-Cbz-L-Lys-D-Ala-D-Ala. This is the highest k(cat) value yet reported for a PBP or other serine peptidases. Activity against a approximately D-Ala-D-Lac peptide substrate was approximately 2-fold lower than against the analogous approximately D-Ala-D-Ala peptide substrate, indicating that deacylation is rate determining for both amide and ester hydrolysis. The pH dependence profiles of both carboxypeptidase activity and beta-lactam acylation were bell-shaped with maximal activity at pH 8.0-8.5. PBP 3 displayed weak transpeptidase activity in a model transpeptidase reaction but was active as an endopeptidase, cleaving dimeric peptide cross-links. Deletion of PBP 3 alone had little effect on viability, growth rate, and morphology of N. gonorrhoeae, although deletion of both PBP 3 and PBP 4, the other low-molecular-mass PBP in N. gonorrhoeae, resulted in a decreased growth rate and marked morphological abnormalities.  相似文献   

2.
A water soluble carbodiimide, 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDC), has been used to crosslink horse heart cytochrome c and trypsin-solubilized bovine liver microsomal cytochrome b5. The reaction was conducted under a variety of solution conditions, and the products were purified by a combination of gel filtration and ion-exchange chromatography. Under all conditions of pH, ionic strength, EDC/protein ratio and reaction time that were studied, multiple 1:1 crosslinked complexes were observed with no evidence of a single, dominant species. Acetate, which is often used as a quencher of such reactions, was found to increase the complexity of the reaction products, presumably through EDC-promoted coupling to cytochrome c. Hydroxylamine treatment of the crosslinked complexes, a procedure frequently used to reverse EDC modification of tyrosyl residues, did not reduce the number of crosslinked components observed. The cytochrome b5 heme group was readily extracted from each of the 1:1 crosslinked complexes by standard techniques, so the crosslinking of heme propionate 7 with Lys79 of cytochrome c that might have been anticipated on the basis of molecular graphics modeling [Salemme, F.R. (1976) J. Mol. Biol. 102, 563-568] was not evident from this analysis. Analysis of HPLC tryptic peptide maps produced from crosslinked complexes revealed reduced specificity of trypsin in hydrolysis of EDC-crosslinked protein-protein complexes and unsatisfactory resolution of crosslinked or branched peptides. Nevertheless, it was possible to demonstrate that residues 52-72 of cytochrome b5, a region predicted to be critical to interaction with cytochrome b5 [Salemme, F.R. (1976) J. Mol. Biol. 102, 563-568] was absent from all peptide maps of 1:1 cytochrome c.cytochrome b5 complexes. Based on these results and a review of the literature involving EDC crosslinking of electron transfer proteins, we conclude that the techniques available for specific protein hydrolysis and separation of crosslinked peptides are not adequate to permit routine unambiguous identification of crosslinking sites in carbodiimide-crosslinked complexes.  相似文献   

3.
The aim of this study was the development of a sensitive and specific substrate for protease A (PrtA), a serralysin-like metzincin from the entomopathogenic microorganism, Photorhabdus. First, cleavage of three biological peptides, the A and B chains of insulin and beta-lipotropin, and of 15 synthetic peptides, was investigated. In the biological peptides, a preference for the hydrophobic residues Ala, Leu and Val was observed at three substrate positions, P2, P1' and P2'. At these positions in the synthetic peptides the preferred residues were Val, Ala and Val, respectively. They contributed to the efficiency of hydrolysis in the order P1' > P2 > P2'. Six amino acids of the synthetic peptides were sufficient to reach the maximum rate of hydrolysis, in accordance with the ability of PrtA to cleave three amino acids from both the N- and the C-terminus of some fragments of biological peptides. Using the best synthetic peptide, a fluorescence-quenched substrate, N-(4-[4'(dimethylamino)phenylazo]benzoyl-EVYAVES-5-[(2-aminoethyl)amino]naphthalene-1-sulfonic acid, was prepared. The approximately 4 x 10(6) M(-1) x s(-1) specificity constant of PrtA (at K(m) approximately 5 x 10(-5) M and k(cat) approximately 2 x 10(2) s(-1)) on this substrate was the highest activity for a serralysin-type enzyme, allowing precise measurement of the effects of several inhibitors and pH on PrtA activity. These showed the characteristics of a metalloenzyme and a wide range of optimum pH, similar to other serralysins. PrtA activity could be measured in biological samples (Photorhabdus-infected insect larvae) without interference from other enzymes, which indicates that substrate selectivity is high towards PrtA. The substrate sensitivity allowed early (14 h post infection) detection of PrtA, which might indicate PrtA's participation in the establishment of infection and not only, as it has been supposed, in bioconversion.  相似文献   

4.
alpha 2-Macroglobulin and the complement components C3 and C4 each contain a metastable binding site that is essential for covalent attachment. Two cyclic peptides are useful models of these unusual protein sites. Five-membered lactam 1 (CH3CO-Gly-Cys-Gly-Glu-Glp-Asn-NH2) contains an internal residue of pyroglutamic acid (Glp). Fifteen-membered thiolactone 2 (CH3CO-Gly-Cys-Gly-Glu-Glu-Asn-NH2 15-thiolactone) contains a thiol ester bond between Cys-2 and Glu-5. These isomeric hexapeptides are spontaneously interconverted in water. Competing with the two isomerization reactions are three reactions involving hydrolysis of 1 and 2. These five processes were found to occur simultaneously under physiologic conditions (phosphate-buffered saline, pH 7.3, 37 degrees C). Best estimates of the five rate constants for these apparent first-order reactions were obtained by comparing the observed molar percentages of peptides 1-4 with those calculated from a set of exponential equations. Both isomerization reactions (ring expansion of 1 to 2, k1 = 6.4 X 10(-5) s-1; ring contraction of 2 to 1, k-1 = 69 X 10(-5) s-1) proceeded faster than any of the hydrolysis reactions: alpha-cleavage of 1 with fragmentation to form dipeptide 3 (k2 = 3.3 X 10(-5) s-1), gamma-cleavage of 1 with ring opening to yield mercapto acid 4 (k3 = 0.35 X 10(-5) s-1), and hydrolysis of 2 with ring opening to give 4 (k4 = 1.9 X 10(-5) s-1). The isomerization rate ratio (k1/k-1 = 10.9) agreed with the isomer ratio at equilibrium (1:2 = 11 starting from 1 and 10 starting from 2). The alpha/gamma regioselectivity ratio (k2/k3 = 9.7) for hydrolysis of the internal Glp residue of 1 was consistent with results for model tripeptides. Part of the chemistry of the protein metastable binding sites can be explained by similar isomerization and hydrolysis reactions.  相似文献   

5.
LexA repressor of Escherichia coli and phage lambda repressor are inactivated in vivo and in vitro by specific cleavage of an Ala-Gly peptide bond in reactions requiring RecA protein. At mildly alkaline pH, the in vitro cleavage reaction also proceeds spontaneously, suggesting that peptide bond hydrolysis is an activity of the repressors rather than of RecA. The spontaneous cleavage reaction, termed "autodigestion", has been characterized for the LexA and lambda repressors. The results show that the reaction is intramolecular. The rate of LexA autodigestion was studied over the pH range 7.15-11.77 and over the temperature range 4-46 degrees C. The logarithm of the rate constant increased linearly with pH and reached a plateau value (2.5 X 10(-3) s-1 at 37 degrees C) at pH above 10. The data closely followed a model in which a single residue side chain (apparent pK = 9.8 at 37 degrees C) must be deprotonated for the protein to show activity. Analysis of the temperature dependence gave the heat of proton dissociation as 19.9 kcal/mol and the heat of activation for hydrolysis as 15.3 kcal/mol at 25 degrees C. Autodigestion of lambda repressor, studied over the pH range 8.65-10.70 at 37 degrees C, was similar to the LexA reaction in its pH dependence, yielding a pK of 9.8. The maximum rate at 37 degrees C for lambda repressor, 6.1 X 10(-5) s-1, was 40 times slower than for LexA, a difference similar to that previously observed in vivo and in vitro for RecA-dependent cleavage reactions. There was no significant solvent deuterium isotope effect on the autodigestion of LexA. Changes in buffer composition, including high concentrations of glycine for lambda repressor and of imidazole or hydroxylamine for LexA, indicated that solvent components other than water do not participate in the rate-determining step. Removal or addition of metal ions did not significantly affect LexA autodigestion. These and other observations suggest that the deprotonated form of an amino acid side chain plays a central role in the chemistry of the cleavage reaction. The above observations establish repressor autodigestion as a member of an emerging set of biologically important self-processing reactions.  相似文献   

6.
H Yamada  R Kuroki  M Hirata  T Imoto 《Biochemistry》1983,22(19):4551-4556
The salt bridge between Lys-13 (epsilon-NH3+) and Leu-129 (alpha-COO-) in lysozyme was converted to an amide bond by 1-ethyl-3-[3-(dimethylamino)propyl]carbodiimide hydrochloride (EDC) reaction in the presence of imidazole (0.3-1 M) at pH 5 and room temperature, followed by dialysis at pH 10. Absence of imidazole under a similar condition did not give this intramolecularly cross-linked lysozyme derivative (CL-lysozyme) but resulted in the formation of intermolecularly cross-linked lysozyme oligomers. From the mechanistic studies on the formation of CL-lysozyme, imidazole was suggested to play the following three roles. (1) Some carboxyl groups activated by EDC in lysozyme were converted to acylimidazole groups which protected them from the reaction with amino groups in other lysozyme molecules at pH 5. These could be hydrolyzed at pH 10 to regenerate free carboxyls. (2) High concentrations of imidazole (pH 5) increased the ionic strength of the solution which weakened the salt bridge in lysozyme and facilitated the activation of the alpha-carboxyl group by EDC. (3) The alpha-carboxyl group activated by EDC was converted to an acylimidazole group which could react with the epsilon-amino group of Lys-13 in the same molecule to form an amide bond. The last step may involve some conformational change of the backbone of lysozyme and be slower than the hydrolysis reaction of the alpha-carboxyl group activated by EDC itself. However, acylimidazole groups are stable against hydrolysis at pH 5. This may afford enough time to allow the epsilon-amino group of Lys-13 to attack the acylimidazole group of Leu-129.  相似文献   

7.
Kinetics of hydrolysis of poly(alkylene amidophosphate)s with amino acids or dipeptides as the side groups was studied by 31P NMR at pH 1.5, 6.5, and 8.5. The direction of hydrolysis and the relative rate coefficients of breaking P-O bonds in the main chain and P-N bonds in the side groups depend strongly on the pH of the medium of hydrolysis. The P-N (amide) bond hydrolyzes much faster than the P-O (ester) bond in acidic and close to neutral conditions (negligible P-O hydrolysis), whereas above pH > or = 8.5 these differences are much smaller. For instance, for 4-Ala the rate coefficients of hydrolysis are equal (in H2O at 37 degrees C and pH 8.5) to 1.9 x 10(-8) s(-1) and 1.0 x 10(-9) s(-1) for the P-N and P-O bonds, respectively, quite different from the values found for the low molecular model 2 (0 and 1.4 x 10(-7) s(-1), respectively).  相似文献   

8.
E S Lightcap  C J Halkides  P A Frey 《Biochemistry》1991,30(42):10307-10313
mu-Monothiopyrophosphate (MTP) binds monovalent and divalent metal ions with dissociation constants (Kd) similar to those for pyrophosphate (PPi). The values of Kd for metal-MTP complexes are the following, as measured kinetically in the hydrolysis of MTP (microM): Mg2+, 32 +/- 4; Mn2+, 5.4 +/- 1.4; and Co2+, 27 +/- 15. The thermodynamically measured (EPR) values for Mg2+ and Co2+ are 28 +/- 13 microns and 11 +/- 4 microM, respectively; and the Kd for the complex MnPPi is 3.4 +/- 0.5 microM. The metal-MTP complexes undergo hydrolysis at rates modestly faster or slower than the rate at which MTP itself reacts. The complexes MgMTP2-, CoMTP2-, and MnMTP2- undergo hydrolytic cleavage with release of thiophosphate with observed first-order rate constants of 1.6 x 10(-2) min-1, 2.3 x 10(-2) min-1, and 0.6 x 10(-2) min-1, respectively, at 35 degrees C, compared with 1.1 x 10(-2) min-1 for MTP4- under the same conditions. Alkali metal cations also stimulate or retard the hydrolysis of MTP. At 25 degrees C and pH 12.2, the observed rate constant for tetramethylammonium MTP4- is 2.1 x 10(-3) min-1, and the estimated rate constants (min-1) for saturating alkali metals under the same conditions are as follows: Li+, 0.25 x 10(-3); Na+, 3.9 x 10(-3), K+, 6.7 x 10(-3); and Cs+, 6.7 x 10(-3). Divalent metal ions markedly retard the hydrolysis of MTP at pH 7 and 8 because complexation shifts the pH rate profile more than 2 pH units toward the acid side.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
We have studied the reaction kinetics of ten manganese porphyrins, differing in their meso substituents, with peroxynitrite (ONOO-) and carbonate radical anion (CO3.) using stopped-flow and pulse radiolysis, respectively. Rate constants for the reactions of Mn(III) porphyrins with ONOO- ranged from 1 x 10(5) to 3.4 x 10(7) m(-1) s(-1) and correlated well with previously reported kinetic and thermodynamic data that reflect the resonance and inductive effects of the substituents on the porphyrin ring. Rate constants for the reactions of Mn(III) porphyrins with CO3. ranged from 2 x 10(8) to 1.2 x 10(9) m(-1)s(-1) at pH 相似文献   

10.
Krogh BO  Shuman S 《Biochemistry》2000,39(21):6422-6432
Vaccinia topoisomerase forms a covalent DNA-(3'-phosphotyrosyl)-enzyme intermediate at sites containing the sequence 5'-CCCTT downward arrow. The covalently bound topoisomerase can religate the CCCTT strand to a 5'-OH-terminated polynucleotide or else transfer the strand to a non-DNA nucleophile such a water or glycerol. Here, we report that vaccinia topoisomerase also catalyzes strand transfer to hydrogen peroxide. The observed alkaline pH-dependence of peroxidolysis is consistent with enzyme-mediated attack by peroxide anion on the covalent intermediate. The reaction displays apparent first-order kinetics. From a double-reciprocal plot of k(obs) versus [H(2)O(2)] at pH 10, we determined a rate constant for peroxidolysis of 6.3 x 10(-)(3) s(-)(1). This rate is slower by a factor of 200 than the rate of topoisomerase-catalyzed strand transfer to a perfectly aligned 5'-OH DNA strand but is comparable to the rate of DNA strand transfer across a 1-nucleotide gap. Strand transfer to 2% hydrogen peroxide is 300 times faster than strand transfer to 20% glycerol and approximately 2000 times faster than topoisomerase-catalyzed hydrolysis of the covalent intermediate. Hydroxylamine is also an effective nucleophile in topoisomerase-mediated strand transfer (k(obs) = 6.4 x 10(-)(4) s(-)(1)). The rates of the peroxidolysis, hydroxylaminolysis, glycerololysis, and hydrolysis reactions catalyzed by the mutant enzyme H265A were reduced by factors of 100-700, in accordance with the 100- to 400-fold rate decrements in DNA cleavage and religation by H265A. We surmise that vaccinia topoisomerase catalyzes strand transfer to DNA and non-DNA nucleophiles via a common reaction pathway in which His-265 stabilizes the scissile phosphate in the transition state rather than acting as a general acid or base.  相似文献   

11.
Chesnokova LS  Witt SN 《Biochemistry》2005,44(33):11224-11233
Hsp70 chaperones are heterotropic allosteric systems in which ATP and misfolded or aggregated polypeptides are the activating ligands. To gain insight into the mechanism by which ATP and polypeptides regulate Hsp70 chaperone activity, the effect of a short peptide on the K(M) for ATP was analyzed using the Escherichia coli Hsp70 called DnaK. In the absence of peptide, the K(-P)(M) for ATP is 52 +/- 11 nM, whereas this value jumps to 14.6 +/- 1.6 microM in the presence of saturating peptide. This finding supports a mechanism in which ATP binding drives the chaperone in one direction and peptide binding pushes the chaperone back in the opposite direction (and thus increases K(M)), according to ATP + DnaK.P <==> ATP.DnaK.P <==> ATP.DnaK* + P, where ATP.DnaK.P is an intermediate from which competing ATP hydrolysis occurs (ATP.DnaK.P --> ADP.DnaK.P). We show that this branched mechanism can even explain how DnaK hydrolyzes ATP in the absence of peptide and that the true rate constant for DnaK-mediated ATP hydrolysis (k(hy)) in the absence of peptide may be as high as 0.5 s(-)(1) (rather than 5 x 10(-)(4) s(-)(1) as often stated in the literature). What happens is that a conformational equilibrium outcompetes ATP hydrolysis and effectively reduces the concentration of the intermediate by a factor of a thousand, resulting in the following relation: k(cat) = k(hy)/1000 = 5 x 10(-)(4) s(-)(1). How polypeptide substrates and the co-chaperone DnaJ modulate DnaK to achieve its theoretical maximal rate of ATP hydrolysis, which we suggest is 0.5 s(-)(1), is discussed.  相似文献   

12.
K Kobayashi  Y Harada  K Hayashi 《Biochemistry》1991,30(34):8310-8315
The reactions of the monodehydroascorbate radical (As.-) with various biological molecules were investigated by pulse radiolysis. As.- reacted with both fully reduced and semiquinone forms of hepatic NADH-cytochrome b5 reductase with second-order rate constants of 4.3 x 10(6) and 3.7 x 10(5) M-1 s-1, respectively, at pH 7.0. In contrast, no reaction of As.- with ferrous cytochrome b5 could be detected by pulse radiolysis, whereas the oxidation of cytochrome b5 by As.- was observed by ascorbate-ascorbate oxidase method. This suggests that the rate constant of As.- with the ferrous cytochrome b5 must be several orders in magnitude smaller than that of the disproportionation of As.-. On the other hand, As.- reduced Fe3+EDTA with a second-order rate constant of 4.0 x 10(6) M-1 s-1 but did not reduce ferric hemoproteins such as metmyoglobin, methemoglobin, and cytochrome b5 by either the pulse radiolysis or the ascorbate-ascorbate oxidase method.  相似文献   

13.
ATP hydrolysis in the presence of polyoxomolybdates at pH levels of 6, 4, and 2 has been investigated with a help of high pressure liquid chromatography (HPLC) analyses, 31P- and 1H NMR measurements, and isothermal titration calorimetry (ITC). The polyoxomolybdates-induced ATP-hydrolysis proceeded satisfactorily in pH < 6 media at 20 degrees C with an optimum pH level of 4, while it was significantly depressed at low temperature of < or = 5 degrees C. At pH levels of 6 and 4, ADP was a main product, and the involvement of [(PO4)2Mo5O15](6-)-like ATP-molybdate complex as an intermediate was implied. At pH 2 ATP was decomposed to AMP with small generation of ADP through the formation of the ATP-molybdate complex isostructural with [(O3POPO3)Mo6O18(H2O)4]4- as an intermediate. The ITC result at pH 4 showed an occurrence of two types of the exothermic binding reactions between molybdate and ATP with binding constants (K) of 6.61x10(4) and 9.40x10(3) M(-1) and molar enthalpy values (deltaH) of -6.32x10(4) and -4.73x10(3) J mol(-1), respectively. Together with the results of 1H NMR measurements, it is deduced that the molybdates interact with not only phosphate sites in the ATP side-chain, but also adenine-ring with an accompanying aggregation of molybdates at pH 4.  相似文献   

14.
The self-assembling behavior, at physiological pH, of the amphiphile peptide (C18)(2)L5CCK8 in nanostructures is reported. Stable aggregates presenting a critical micellar concentration of 2 x 10(-6) mol kg(-1), and characterized by water exposed CCK8 peptide in beta-sheet conformation, are obtained. Small angle neutron scattering experiments are indicative for a 3D structure with dimensions > or =100 nm. AFM images confirm the presence of nanostructures. Fluorescence experiments indicating the sequestration of pyrene, chosen as drug model, and the anticancer Doxorubicin within the nanostructures are reported.  相似文献   

15.
Bidwai AK  Ok EY  Erman JE 《Biochemistry》2008,47(39):10458-10470
The spectrum of the ferric heme domain of the direct oxygen sensor protein from Escherichia coli ( EcDosH) has been measured between pH 3.0 and 12.6. EcDosH undergoes acid denaturation with an apparent p K a of 4.24 +/- 0.05 and a Hill coefficient of 3.1 +/- 0.6 and reversible alkaline denaturation with a p K a of 9.86 +/- 0.04 and a Hill coefficient of 1.1 +/- 0.1. Cyanide binding to EcDosH has been investigated between pH 4 and 11. The EcDosH-cyanide complex is most stable at pH 9 with a K D of 0.29 +/- 0.06 microM. The kinetics of cyanide binding are monophasic between pH 4 and 8. At pH >or=8.5, the reaction is biphasic with the fast phase dependent upon the cyanide concentration and the slow phase independent of cyanide. The slow phase is attributed to conversion of denatured EcDosH to the native state, with a pH-independent rate of 0.052 +/- 0.006 s (-1). The apparent association rate constant for cyanide binding to EcDosH increases from 3.6 +/- 0.1 M (-1) s (-1) at pH 4 to 520 +/- 20 M (-1) s (-1) at pH 11. The dissociation rate constant averages (8.6 +/- 1.3) x 10 (-5) s (-1) between pH 5 and 9, increasing to (1.4 +/- 0.1) x 10 (-3) s (-1) at pH 4 and (2.5 +/- 0.1) x 10 (-3) s (-1) at pH 12.2. The mechanism of cyanide binding is consistent with preferential binding of the cyanide anion to native EcDosH. The reactions of imidazole and H 2O 2 with ferric EcDosH were also investigated and show little reactivity.  相似文献   

16.
Matsuno H  Furusawa H  Okahata Y 《Biochemistry》2005,44(7):2262-2270
Catalytic DNA cleavage reactions by an ATP-dependent deoxyribonuclease (DNase) from Micrococcus luteus were monitored directly with a DNA-immobilized 27-MHz quartz-crystal microbalance (QCM). The 27-MHz QCM is a very sensitive mass-measuring device in aqueous solution, as the frequency decreases linearly with increasing mass on the electrode at a nanogram level. Three steps in ATP-dependent DNA hydrolysis reactions, including (1) binding of DNase to the end of double-stranded DNA (dsDNA) on the QCM electrode (mass increase), (2) degradation of one strand of dsDNA in the 3' --> 5' direction depending on ATP (mass decrease), and (3) release of the enzyme from the nonhydrolyzed 5'-free-ssDNA (mass decrease), could be monitored stepwise from the time dependencies of QCM frequency changes. Kinetic parameters for each step were obtained as follows. The binding constant (K(a)) of DNase to the dsDNA was determined as (28 +/- 2) x 10(6) M(-)(1) (k(on) = (8.0 +/- 0.3) x 10(3) M (-)(1) s(-)(1) and k(off) = (0.29 +/-0.01) x 10(-)(3) s(-)(1)), and it decreased to (0.79 +/- 0.16) x 10(6) M(-)(1) (k'(on) = (2.3 +/- 0.2) x 10(3) M (-)(1) s(-)(1) and k'(off) = (2.9 +/- 0.1) x 10(-)(3) s(-)(1)) for the completely nonhydrolyzed 5'-free ssDNA. This is the reason the DNase bound to the dsDNA substrate can easily release from the nonhydrolyzed 5'-free-ssDNA after the complete hydrolysis of the 3' --> 5' direction of the complementary ssDNA. K(a) values depended on the DNA structures on the QCM, and the order of these values was as follows: the dsDNA having a 4-base-mismatched base-pair end (3) > the dsDNA having a 5' 15-base overhanging end (2) > the dsDNA having a blunt end (1) > the ssDNA having a 3'-free end (4) > the ssDNA having a 5'-free end (5). Thus, DNase hardly recognized the free 5' end of ssDNA. Michaelis-Menten parameters (K(m) for ATP and k(cat)) of the hydrolysis process also could be obtained, and the order of k(cat)/K(m) was as follows: the dsDNA having a blunt end (1) approximately the dsDNA having a 4-base-mismatched base-pair end (3) > the ssDNA having a free 3' end (4) > the ssDNA having a free 5' end (5). Thus, DNase could not recognize and not hydrolyze the free 5' end of ssDNA. The DNA hydrolysis reaction could be driven by dATP and GTP (purine base) as well as ATP, whereas the cleavage efficiency was very low driven with UTP, CTP (pyrimidine base), ADP, and AMP.  相似文献   

17.
Time-resolved absorption spectroscopy has been employed to study the directionality and rate of charge transfer in W-Y and Ac-W-Y dipeptides as a function of pH. Excitation with 266-nm nanosecond laser pulses produces both W (or [WH](+), depending on pH) and Y. Between pH 6 and 10, W to was found to oxidize Y with k(X)=9.0x10(4) s(-1) and 1.8x10(4) s(-1) for the W-Y and Ac-W-Y dipeptide systems, respectively. The intramolecular charge transfer rate increases as the pH is lowered over the range 6>pH>2. For 10W-Y(-) (Y(-), tyrosinate anion), with a rate constant of k(X)=1.2x10(5) s(-1). The dependence of charge transfer directionality between W and Y on pH is important to the enzymatic function of several model and natural biological systems as discussed here for ribonucleotide reductase.  相似文献   

18.
Rate constants have been determined for hydrolysis of the acetate, glutarate, and phthalate monoesters of 2-hydroxy-1,10-phenanthroline in water at 30°C and μ = 0.1 M with KCl. The hydrolysis reactions of the esters are hydroxide ion catalyzed at pH > 9. The phthalate and glutarate monoesters have in addition pH-independent reactions from pH 5.5 to 9 that involve intramolecular participation by the neighboring carboxylate anion. The pH-independent reaction of the glutarate monoester is 5-fold faster than that of the phthalate monoester. The plots of log kobsd vs pH for hydrolysis of the carboxyl substituted esters are bell shaped at pH < 5, which indicates a rapid reaction of the zwitterionic species (carboxyl anion and protonated phenanthroline nitrogen). The divalent metal ions, Cu2+, Ni2+, Zn2+, and Co2+, complex strongly with the esters; saturation occurs at metal ion concentrations less than 0.01 M. The 1:1 metal ion complexes have greatly enhanced rates of hydrolysis; the second-order rate constants for the OH reactions are increased by factors of 105 to 108 by the metal ion. The pH-rate constant profiles for the phthalate and glutarate ester metal ion complexes have a sigmoidal region below pH 6 that can be attributed to a metal ion-promoted carboxylate anion nucleophilic reaction. The carboxyl group reactions are enhanced 102 - to 103 -fold by the metal ions, which allows the neighboring group reaction to be competitive with the favorable metal ion-promoted OH reaction at pH < 6, but not at pH > 6. The half-lives of the pH-independent neighboring carboxyl group reactions of the Cu(II) complexes at 30°C are l2 s. The other metal ion complexes are only slightly less reactive (half-lives vary from 2.5 to 40 s). These are the most rapid neighboring carboxyl group reactions that have been observed in ester hydrolysis.  相似文献   

19.
Electron-capture gas-liquid chromatography was used to study the spontaneous hydrolysis of heroin in phosphate buffer (pH 6.4 and pH 7.4) at 23 degrees C. Aliquots of solution were taken over a 24-h period. After extraction at pH 8.9 into propan-2-ol (10%)-ethyl acetate, deacetylated products were made into hepafluorobutyrate derivatives which were analyzed quantitatively using nalorphine as the internal standard. Heroin decomposes to 06-monoacetylmorphine (06-MAM) under these conditions. Further decomposition to morphine was not observed. Spontaneous hydrolysis was faster at pH 7.4 (first-order rate constant, 9.6 x 10-5 min-1) than at pH 6.4 (first-order rate constant, 3.0 x 10-5 min-1). In 24 h, the decomposition to 06-MAM was 13 and 4%, respectively.  相似文献   

20.
Use of rigorous equilibration kinetics to evaluate rate constants for the Fe(CN)6 4- reduction of horse-heart cytochrome c in the oxidized form, cyt c (III), has shown that limiting kinetics do not apply with concentrations of Fe(CN)6 4- (the reactant in excess) in the range 2-10 x 10(-4) M, I = 0.10 M (NaCl). The reaction conforms to a first-order rate law in each reactant, and at 25 degrees C, pH 7.2 (Tris), it is concluded that K for association prior to electron transfer is less than 200 M-1. From previous studies at 25 degrees C, ph 7.0 (10(-1) M phosphate), I = 0.242 M (NaCl), a value K = 2.4 x 10(3) M-1 has been reported. Had such a value applied, some or all of the redox inactive complexes Mo(CN)8 4-, Co(CN)6 3-, Cr(CN)6 3-, Zr(C2O4)4 4- present in amounts 5-20 x 10(-4) M would have been expected to associate at the same site and partially block the redox process. No effect on rats was observed. With the reductants Fe(CN)5(4-NH2-py)3- and Fe(CN)5(imid)3-, reactions proceeded to greater than 90% completion and rate laws were again first order in each reactant. Rate constants (M-1 sec-1) at 25 degrees C, pH 7.2 (Tris), I = 0.10 M (NaCl), are Fe(CN)6 4- (3.5 x 10(4)), Fe(CN)5(4-NH2py)3- (6.7 x 10(5), and Fe(CN)5(imid)3- (4.2 x 10(5). Related reactions in which cyt c(II) is oxidized are also first order in each reactant, Fe(CN)6 3- (9.1 x 10(6)), Fe(CN)5(NCS)3- (1.3 x 10(6)), Fe(CN)5(4-NH2py)2- (3.8 x 10(6) at pH 9.4), and Fe(CN)5(NH3)2- (2.75 x 10(6) at ph 8). Redox inactive Co(CN)6 3- (1.0 x 10(-3) M) has no effect on the reaction of Fe(CN)6 3- which suggests that a recent interpretation for the Fe(CN)6 3- oxidation of cyt c(II), I = 0.07 M, may also require reappraisal.  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