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.
Hydrolytic reactions of 2',3'-O-methyleneadenos-5'-yl bis(2',5'-di-O-methylurid-3'-yl) phosphate (1), a sugar O-alkylated trinucleoside 3',3',5'-monophosphate, have been followed by RP HPLC over a wide pH range. Under neutral and mildly acidic conditions, the only reaction observed was a pH-independent cleavage of the O-C5' bond of the 5'-linked nucleoside. Under more alkaline conditions nucleophilic attack by hydroxide ion starts to compete. The reaction is first order in [OH(-)] and becomes predominant at pH 10. Each of the 3'-linked nucleosides is displaced 2.9 times as readily as the 5'-linked one. To determine the beta(lg) value for the hydroxide ion catalyzed hydrolysis of 1, two diesters (2a,b) having 2',3'-O-methyleneadenosine (7) and 2',5'-di-O-methyluridine (4) as leaving groups were hydrolyzed under alkaline conditions. Since the beta(lg) value for this reaction is known, DeltapK(a) between 4 and 7 could be calculated. The beta(lg) for the hydrolysis of 1 was estimated to be -0.5 with use of this information. The mechanisms of the partial reactions and the role of leaving group properties in ribozyme reactions of large ribozymes are discussed.  相似文献   

2.
Hydrolytic reactions of 3'-amino-3'-deoxyuridylyl-3',5'-uridine (2a), an analogue of uridylyl-3',5'-uridine having the 3'-bridging oxygen replaced with nitrogen, have been followed by RP HPLC over a wide pH range. The only reaction taking place under alkaline conditions (pH > 9) is hydroxide ion-catalyzed hydrolysis (first-order in [OH(-)]) to a mixture of 3'-amino-3'-deoxyuridine 3'-phosphoramidate (7) and uridine (4). The reaction proceeds without detectable accumulation of any intermediates. At pH 6-8, a pH-independent formation of 3'-amino-3'-deoxyuridine 2'-phosphate (3) competes with the base-catalyzed cleavage. Both 3 and in particular 7 are, however, rather rapidly dephosphorylated under these conditions to 3'-amino-3'-deoxyuridine (5). In all likelihood, both 3 and 7 are formed by an intramolecular nucleophilic attack of the 2'-hydroxy function on the phosphorus atom, giving a phosphorane-like intermediate or transition state. Under moderately acidic conditions (pH 2-6), the predominant reaction is acid-catalyzed cleavage of the P-N3' bond (first-order in [H(+)]) that yields an equimolar mixture of 5 and uridine 5'-phosphate (6). The reaction is proposed to proceed without intramolecular participation of the neighboring 2'-hydroxyl group. Under more acidic conditions (pH < 2), hydrolysis to 3 and 4 starts to compete with the cleavage of the P-N bond, and this reaction is even the fastest one at pH < 1. Formation of 2'-O,3'-N-cyclic phosphoramidate as an intermediate appears probable, although its appearance cannot be experimentally verified. The rate constants for various partial reactions have been determined. The reaction mechanisms and the effect that replacing the 3'-oxygen with nitrogen has on the behavior of the phosphorane intermediate are discussed.  相似文献   

3.
The hydrolysis of diadenosine 5',5'-triphosphate to AMP and ADP has been studied over a wide pH-range. Under acidic conditions the reaction shows a first-order dependence on the hydronium ion concentration. Below pH 3 the rate-increase begins to level off. From pH 6 to 9 the hydrolysis is slow and pH-independent. Base-catalysed hydrolysis is observed in NaOH-solutions. Under alkaline conditions an intramolecular nucleophilic attack on the phosphate producing 3',5'-cAMP is also observed, but it is slower than the intermolecular reaction. Depurination of the adenosine moieties competes with the hydrolysis both under acidic and alkaline conditions, but the mechanisms are different. The temperature-dependence of the hydrolysis of Ap(3)A and the depurination of adenosine moieties were studied under acidic conditions, and the activation parameters of the reactions were calculated. The results of the work reflect the fact that the negatively charged polyphosphate group is very resistant towards nucleophilic attack. An efficient catalysis is only observed under acidic conditions, where the phosphate group becomes protonated. General acids or bases did not catalyse the hydrolysis. Furthermore, hydroxide ion catalysed cleavage is only observed at high base concentrations and other negatively charged nucleophiles did not attack the phosphate groups of diadenosine polyphosphates.  相似文献   

4.
The synthesis of the C5' tert-butyl ketone of thymidine 1a and 2'-deoxyguanosine 2 is achieved by reaction of 5'-C-cyano derivatives with tert-butyl lithium followed by acid hydrolysis. The 5'R configuration is assigned by X-ray crystal structure determination of an opportunely protected derivative of 1a. The (5'S)-isomers of both nucleosides are not stable, and a complete decomposition occurs in the reaction medium. The photochemistry of 1a and 2 effectively produced the thymidin-5'-yl radical and the 2'-deoxyguanosin-5'-yl radical, respectively. In the thymidine system, the C5' radical is fully quenched in the presence of a physiological concentration of thiols. In the 2'-deoxyguanosine system, the C5' radical undergoes intramolecular attack onto the C8-N7 double bond of guanine leading ultimately to the 5',8-cyclo-2'-deoxyguanosine derivative. The cyclization of the 2'-deoxyguanosin-5'-yl radical occurs with a rate constant of ca. 1x10(6) s-1 and is highly stereoselective affording only the (5'S)-diastereomer.  相似文献   

5.
Hydrolytic reactions of 2',3'-O-methyleneadenos-5'-yl 2',5'-di-O-methylurid-3'-yl 5'-O-methylurid-3'(2')-yl phosphate (1a,b) have been followed by RP-HPLC over a wide pH range to evaluate the feasibility of occurrence of phosphate-branched RNA under physiological conditions. At pH <2, where the decomposition of is first order in [H3O+], the P-O5' bond is cleaved 1.5 times as rapidly as the P-O3' bond. Under these conditions, the reaction probably proceeds by an attack of the 2'-OH on the phosphotriester monocation. Over a relatively wide range from pH 2 to 5, the hydrolysis is pH-independent, referring to rapid initial deprotonation of the attacking 2'-OH followed by general acid catalyzed departure of the leaving nucleoside. The P-O5' bond is cleaved 3 times as rapidly as the P-O3' bond. At pH 6, the reaction becomes first order in [HO-], consistent with an attack of the 2'-oxyanion on neutral phosphate. The product distribution is gradually inversed: in 10 mmol L(-1) aqueous sodium hydroxide, cleavage of the P-O3' bond is favored over P-O5' by a factor of 7.3. The results of the present study suggest that the half-life for the cleavage of under physiological conditions is only 100 s. Even at pH 2, where is most stable, the half-life for its cleavage is less than one hour and the isomerization between and is even more rapid than cleavage. The mechanisms of the partial reactions are discussed.  相似文献   

6.
The reaction catalyzed by the protein phosphatase-1 (PP1) has been examined by linear free energy relationships and kinetic isotope effects. With the substrate 4-nitrophenyl phosphate (4NPP), the reaction exhibits a bell-shaped pH-rate profile for kcat/KM indicative of catalysis by both acidic and basic residues, with kinetic pKa values of 6.0 and 7.2. The enzymatic hydrolysis of a series of aryl monoester substrates yields a Br?nsted beta(lg) of -0.32, considerably less negative than that of the uncatalyzed hydrolysis of monoester dianions (-1.23). Kinetic isotope effects in the leaving group with the substrate 4NPP are (18)(V/K) bridge = 1.0170 and (15)(V/K) = 1.0010, which, compared against other enzymatic KIEs with and without general acid catalysis, are consistent with a loose transition state with partial neutralization of the leaving group. PP1 also efficiently catalyzes the hydrolysis of 4-nitrophenyl methylphosphonate (4NPMP). The enzymatic hydrolysis of a series of aryl methylphosphonate substrates yields a Br?nsted beta(lg) of -0.30, smaller than the alkaline hydrolysis (-0.69) and similar to the beta(lg) measured for monoester substrates, indicative of similar transition states. The KIEs and the beta(lg) data point to a transition state for the alkaline hydrolysis of 4NPMP that is similar to that of diesters with the same leaving group. For the enzymatic reaction of 4NPMP, the KIEs are indicative of a transition state that is somewhat looser than the alkaline hydrolysis reaction and similar to the PP1-catalyzed monoester reaction. The data cumulatively point to enzymatic transition states for aryl phosphate monoester and aryl methylphosphonate hydrolysis reactions that are much more similar to one another than the nonenzymatic hydrolysis reactions of the two substrates.  相似文献   

7.
We here present a theoretical study of the alkaline hydrolysis of a phosphodiester (methyl p-nitrophenyl phosphate or MpNPP) in the active site of Escherichia coli alkaline phosphatase (AP), a monoesterase that also presents promiscuous activity as a diesterase. The analysis of our simulations, carried out by means of molecular dynamics (MD) simulations with hybrid quantum mechanics/molecular mechanics (QM/MM) potentials, shows that the reaction takes place through a D(N)A(N) or dissociative mechanism, the same mechanism employed by AP in the hydrolysis of monoesters. The promiscuous activity observed in this superfamily can be then explained on the basis of a conserved reaction mechanism. According to our simulations the specialization in the hydrolysis of phosphomonoesters or phosphodiesters, developed in different members of the superfamily, is a consequence of the interactions established between the protein and the oxygen atoms of the phosphate group and, in particular, with the oxygen atom that bears the additional alkyl group when the substrate is a diester. A water molecule, belonging to the coordination shell of the Mg(2+) ion, and residue Lys328 seem to play decisive roles stabilizing a phosphomonoester substrate, but the latter contributes to increase the energy barrier for the hydrolysis of phosphodiesters. Then, mutations affecting the nature or positioning of Lys328 lead to an increased diesterase activity in AP. Finally, the capacity of this enzymatic family to catalyze the reaction of phosphoesters having different leaving groups, or substrate promiscuity, is explained by the ability of the enzyme to stabilize different charge distributions in the leaving group using different interactions involving either one of the zinc centers or residues placed on the outer side of the catalytic site.  相似文献   

8.
Reaction pathways and free energy barriers for alkaline hydrolysis of the highly neurotoxic insecticide 2-trimethylammonioethyl methylphosphonofluoridate and related organophosphorus compounds were studied by performing first-principles electronic structure calculations on representative methylphosphonofluoridates, (RO)CH3P(O)F, in which R = CH2CH2N+(CH3)3, CH3, CH2CH2C(CH3)3, CH2CH2CH(CH3)2, CH(CH3)CH2N+(CH3)3, and CH(CH3)CH2N(CH3)2. The dominant reaction pathway was found to be associated with a transition state in which the attacking nucleophile OH- and the leaving group F- are positioned on opposite sides of the plane formed by the three remaining atoms attached to the phosphorus in order to minimize the electrostatic repulsion between these two groups. The free energy barriers calculated for the rate-determining step of the dominant pathway are 12.5 kcal/mol when R = CH2CH2N+(CH3)3, 15.5 kcal/mol when R = CH3, 17.9 kcal/mol when R = CH2CH2C(CH3)3, 16.5 kcal/mol when R = CH2CH2CH(CH3)2, 13.4 kcal/mol when R = CH(CH3)CH2N+(CH3)3, and 18.7 kcal/mol when R = CH(CH(3))CH(2)N(CH(3))(2). The calculated free energy barriers are in good agreement with available experimentally derived activation free energies, i.e. 14.7 kcal/mol when R = CH(3), 13.4 kcal/mol when R = CH2CH2N+(CH3)3, and 13.9 kcal/mol when R = CH(CH3)CH2N+(CH3)3. A detailed analysis of the calculated energetic results and available experimental data suggests that the net charge of the molecule (M) being hydrolyzed is a prominent factor affecting the free energy barrier (DeltaG) for the alkaline hydrolysis of phosphodiesters, phosphonofluoridates, and related organophosphorus compounds. The electrostatic interactions between the attacking nucleophile OH- and the molecule M being hydrolyzed favor such an order of the free energy barrier: DeltaG(M(+)+OH-) < DeltaG(M0+OH-) < DeltaG(M(-)+OH-), where M+, M0, and M- represent the cationic, neutral, and anionic molecules, respectively. The change of the substituent R in (RO)CH(3)P(O)F from CH3 to CH2CH2N+(CH3)3 is associated with both the electrostatic and steric effects on the free energy barrier, but the electrostatic effect dominates the substituent shift of the free energy barrier. This helps to better understand why the alkaline hydrolysis of (RO)CH3P(O)F with R = CH2CH2N+(CH3)3 and CH(CH3)CH2N+(CH3)3 is significantly faster than that with R = CH3. The effect of electrostatic interaction also helps to understand why the rate constants for the alkaline hydrolysis of phosphodiesters, such as intramolecular second messenger adenosine 3',5'-phosphate (cAMP), are generally smaller than those for the alkaline hydrolysis of the phosphonofluoridates and related phosphotriesters.  相似文献   

9.
A comprehensive study on the gas-phase alkaline hydrolysis of cephalosporins by using the semiempirical AM1 method was carried out. Cephalothin was the model compound used on account of the presence of a good leaving group at C(3′). According to the results obtained, the hydrolysis process takes place via a twostep reaction mechanism that involves the formation of an intermediate with a fully open β-lactam ring that still preserves the acetate group. Likewise, the exo methylene end product is chiefly formed by nucleophilic attack on the β-lactam carbonyl group of cephalosporins containing a good leaving group at C(3′). On the other hand, the alternative mechanism involving hydrolysis of the ester function in the side chain at 3′ and subsequent hydrolysis of the resulting β-lactam yieds essentially the corresponding enamine. The presence of a first solvation layer consisting of five water molecules showed that, even though some potential barriers are slightly increased, the mechanism involved is identical to that of the gas-phase hydrolysis of this antibiotic. © John Wiley & Sons, Inc.  相似文献   

10.
The alkaline hydrolysis of title esters possessing acidic leaving groups follows an E1cB mechanism involving the participation of an "extra extended" p-oxo ketene intermediate. For the hydrolysis of the 2,4-dinitrophenyl ester kinetic data, activation parameters and trapping of the intermediate clearly indicate that the dissociative pathway carries the reaction flux. Break in the Bronsted plot of the apparent second-order rate constants versus the pK(a) of the leaving group suggests that the reaction mechanism changes from E1cB to B(Ac)2 for esters having pK(a) higher than about 6.  相似文献   

11.
The synthesis and conformational analysis of 2'-O,5-dimethyluridylyl(3'-5')-2'-O,5-dimethyluridine (1a), the analogue of thymidylyl(3'-5')thymidine (TpT; 1b) in which a methoxy group replaces each 2'-alpha-hydrogen atom, are described. In comparison with TpT, such modification increases the population of the C3'-endo conformer of the sugar ring puckering at the 5'- and 3'-ends from 30 to 75% and from 37 to 66%, respectively. Photolyses of 1a and TpT at 254 nm are qualitatively comparable (the cis-syn cyclobutane pyrimidine dimer and the (6-4) photoproduct are formed), although it is significantly faster in the case of 1a. These results are explained by the increased propensity of the modified dinucleotide to adopt a base-stacked conformation geometry reminiscent of that for TpT.  相似文献   

12.
To obtain detailed data on the kinetics of hydrolytic reactions of triester-like nucleoside 5'-O-aryl-N-alkylphosphoramidates, potential prodrugs of antiviral nucleoside monophosphates, the hydrolysis of diastereomeric (Rp/Sp) thymidine 5'-{O-phenyl-N-[(1S)-2-oxo-2-methoxy-1-methylethyl]phosphoramidate} (3), a phosphoramidate derived from the methyl ester of L-alanine, has been followed by reversed-phase HPLC over the range from Ho=0 to pH 8 at 90 degrees C. According to the time-dependent product distributions, the hydrolysis of 3 proceeds at pH<4 by two parallel routes, namely by nucleophilic displacement of the alaninyl ester moiety by a water molecule and by hydrolysis of the carboxylic ester linkage that allows intramolecular attack of the carboxy group on the phosphorus atom, thereby resulting in the departure of either thymidine or phenol without marked accumulation of any intermediates. Both routes represent about half of the overall disappearance of 3. The departure of phenol eventually leads to the formation of thymidine 5'-phosphate. At pH>5, the predominant reaction is hydrolysis of the carboxylic ester linkage followed by intramolecular displacement of a phenoxide ion by the carboxylate ion and hydrolysis of the resulting cyclic mixed anhydride into an acyclic diester-like thymidine 5'-phosphoramidate. The latter product accumulated quantitatively without any indication of further decomposition. Hydroxide-ion-catalyzed P--OPh bond cleavage of the starting material 3 occurred as a side reaction. Comparative measurements with thymidine 5'-{N-[(1S)-2-oxo-2-methoxy-1-methylethyl]phosphoramidate} (4) revealed that, under acidic conditions, this diester-like compound is hydrolyzed by P--N bond cleavage three orders of magnitude more rapidly than the triester-like 3. At pH>5, the stability order is reversed, with 3 being hydrolyzed six times as rapidly as 4. Mechanisms of the partial reactions are discussed.  相似文献   

13.
Several analogues of diuridine phosphate (UpU) were synthesized in order to investigate why replacing the 2'-hydroxyl with a 2'-amino group prevents hydrolysis. These analogues were designed to investigate what influence the 2'-substituent and 5'-leaving group have upon the rate of hydrolysis. All the analogues were considerably more labile than UpU toward acid-base-catalyzed hydrolysis. In the pH region from 6 to 9, the rate of hydrolysis of uridylyl (3'-5') 5'-thio-5'-deoxyuridine (UpsU) hydrolysis rose, in a log linear fashion, from a value of 5 x 10(-)(6) s(-)(1) at pH 6 to 3200 x 10(-)(6) s(-)(1) at pH 9, indicating that attack on the phosphorus by the 2'-oxo anion is rate-limiting in the hydrolysis mechanism. In contrast, the rate of uridylyl (3'-5') 5'-amino-5'-deoxyuridine (UpnU) hydrolysis fell from a value of 1802 x 10(-)(6) s(-)(1) at pH 5 to 140 x 10(-)(6) s(-)(1) at pH 7.5, where it remained constant up to pH 11.5, thus indicating an acid-catalyzed reaction. The analogue 2'-amino-2'-deoxyuridylyl (3'-5') 5'-thio-5'-deoxyuridine (amUpsU) was readily hydrolyzed above pH 7, in contrast to the hydrolytic stability of amUpT, with rates between 85 x 10(-)(6) s(-)(1) and 138 x 10(-)(6) s(-)(1). The hydrolysis of 2'-amino-2'-deoxyuridylyl (3'-5') 5'-amino-5'-deoxythymidine (amUpnT) rose from 17 x 10(-)(6) s(-)(1) at pH 11.5 to 11 685 x 10(-)(6) s(-)(1) at pH 7.0, indicating an acid-catalyzed reaction, where protonation of the 5'-amine is rate limiting. The cleavage rates of UpsU, UpnU, and amUpsU were accelerated in the presence of Mg(2+), Zn(2+), and Cd(2+) ions, but a correlation with interaction between metal ion and leaving group could only be demonstrated for amUpsU. UpsU and UpnU are also substrates for RNase A with UpsU having similar Michaelis-Menten parameters to UpU. In contrast, UpnU is more rapidly degraded with an approximate 35-fold increase in catalytic efficiency, which is reflected purely in an increase in the value of k(cat).  相似文献   

14.
Abstract— The photoreaction of 8-methoxypsoralen (8-MOP) with thymidine in solid film state yielded two 4', 5'-monoadducts (a pair of diastereomers) and three 3,4-monoadducts. The stereochemistry of two 4', 5'-monoadducts was found to be cis-syn and trans-syn and one 3,4-monoadduct was cis-anti. In addition to these monoadducts, 3,4-, 4', 5'-biadducts were also formed during the reaction, but the isolation of each isomer of these adducts was not successful; however, the formation of these biadducts was confirmed by UV, IR, TLC and photosplitting experiments.  相似文献   

15.
To assess the extent to which steric constraints may influence the product distribution of the reactions of the large ribozymes, phosphate-branched oligonucleotides of varying length and sequence have been synthesized and their alkaline hydrolysis studied over a wide temperature range. At low temperatures, the branching trinucleoside-3',3',5'-monophosphate moiety is hydrolyzed almost exclusively by P-O3' fission. At higher temperatures, P-O5' fission competes, accounting at most for 22% of the overall reaction. The results suggest that steric constraints imposed by the secondary structure of the reaction site may significantly contribute to the observed regioselectivity of the transesterification reactions catalyzed by the large ribozymes.  相似文献   

16.
The enzymatic synthesis of thymidine from 2-deoxy-D-ribose-5-phosphate is achieved, in a one-pot two-step reaction using phosphoribomutase (PRM) and commercially available thymidine phosphorylase (TP). In the first step the sugar-5-phosphate is enzymatically rearranged to alpha-2-deoxy-D-ribose-1-phosphate. Highly active PRM is easily obtained from genetically modified overproducing E. coli cells (12,000 units/84 mg protein) and is used without further purification. In the second step thymine is coupled to the sugar-1-phosphate. The thermodynamically unfavorable equilibrium is shifted to the product by addition of MnCl(2) to precipitate inorganic phosphate. In this way the overall yield of the beta-anomeric pure nucleoside increases from 14 to 60%. In contrast to uracil, cytosine is not accepted by TP as a substrate. Therefore, 2'-deoxy-cytidine is obtained by functional group transformations of the enzymatically prepared 2'-deoxy-uridine. The method has been demonstrated by the synthesis of [2',5'-(13)C(2)]- and [1',2',5'-(13)C(3)]thymidine as well as [1',2',5'-(13)C(3)]2'-deoxyuridine and [3',4'-(13)C(2)]2'-deoxycytidine. In addition the nucleoside bases thymine and uracil are tetralabeled at the (1,3-(15)N(2),2,4-(13)C(2))-atomic positions. All compounds are prepared without any scrambling or dilution of the labeled material and are thus obtained with a very high isotope enrichment (96-99%). In combination with the methods that have been developed earlier it is concluded that each of the (13)C- and (15)N-positions and combination of positions of the pyrimidine deoxynucleosides can be efficiently labeled starting from commercially available and highly (13)C- or (15)N-enriched formaldehyde, acetaldehyde, acetic acid, potassium cyanide, methylamine hydrochloride, and ammonia.  相似文献   

17.
Beta-sultams are the sulfonyl analogues of beta-lactams, and N-acyl beta-sultams are novel inactivators of the class C beta-lactamase of Enterobacter cloacae P99. They sulfonylate the active site serine residue to form a sulfonate ester which subsequently undergoes C-O bond fission and formation of a dehydroalanine residue by elimination of the sulfonate anion as shown by electrospray ionization mass spectroscopy. The analogous N-acyl beta-lactams are substrates for beta-lactamase and undergo enzyme-catalyzed hydrolysis presumably by the normal acylation-deacylation process. The rates of acylation of the enzyme by the beta-lactams, measured by the second-order rate constant for hydrolysis, kcat/K(m), and those of sulfonylation by the beta-sultams, measured by the second-order rate constant for inactivation, k(i), both show a similar pH dependence to that exhibited by the beta-lactamase-catalyzed hydrolysis of beta-lactam antibiotics. Electron-withdrawing groups in the aryl residue of the leaving group of N-aroyl beta-lactams increase the rate of alkaline hydrolysis and give a Bronsted beta(lg) of -0.55, indicative of a late transition state for rate-limiting formation of the tetrahedral intermediate. Interestingly, the corresponding Bronsted beta(lg) for the beta-lactamase-catalyzed hydrolysis of the same substrates is -0.06, indicative of an earlier transition state for the enzyme-catalyzed reaction. By contrast, although the Bronsted beta(lg) for the alkaline hydrolysis of N-aroyl beta-sultams is -0.73, similar to that for the beta-lactams, that for the sulfonylation of beta-lactamase by these compounds is -1.46, compatible with significant amide anion expulsion/S-N fission in the transition state. In this case, the enzyme reaction displays a later transition state compared with hydroxide-ion-catalyzed hydrolysis of the beta-sultam.  相似文献   

18.
5'-Nucleotides of A and U with the phosphate activated with 1-methyladenine generate RNA oligomers containing 40-50 monomers in 1 day in reactions catalyzed by montmorillonite. The corresponding monomers of C give oligomers that are 20-25-mers in length after a 9-day reaction. It was not possible to determine the chain lengths of the oligomers of G since they did not give well-defined bands on gel electrophoresis. Co-oligomers of A and U as well as A, U, G, and C were also prepared. The oligo(A)s formed were separated by gel electrophoresis, and the bands of the 7-39-mers were isolated, the 3',5'-phosphodiester bonds were cleaved by RNase T(2), and the terminal phosphate groups were cleaved with alkaline phosphatase. HPLC analysis revealed that the proportions of A(5)'pp(5)'A, A, A(2)'pA, and A(2)'pA(2)'pA formed were almost the same for the long and shorter oligomers. A similar structure analysis performed on the oligo(U)s established that the proportions of U(5)'pp(5)'U, U, U(2)'pU, U(2)'pU(2)'pU, U(2)'pU(2)'pU(2)'pU, and U(2)'pU(2)'pU(2)'pU(2)'pU did not vary with chain length. The structural analysis of the oligomers of A revealed that 74% of the phosphodiester bonds were 3',5'-linked a value slightly greater than 67% observed when imidazole was the activating group. 61% of the bonds in the U oligomers were 3',5'-linked, which is almost 3 times greater than the 20% measured when imidazole was the activating group. The potential significance of these data to the origin and early evolution of life is discussed.  相似文献   

19.
A theoretical study of the water-assisted alkaline hydrolysis of 2-azetidinone, 3-formylamino-2-azetidinone and 3-formylamino-2-azetidine-1-sulfonate ion is carried out at the B3LYP/6-31+G* level. The effect of bulk solvent is taken into account using the PCM solvation model while specific solvent effects are represented by the inclusion of an ancillary water molecule along the reaction profile. The calculated free energy barriers in solution are in reasonable agreement with experimental values. The observed substituent effects due to the presence of the 3-formylamino and the SO(3) groups attached to the beta-lactam ring are crucial factors determining the hydrolysis of monobactam antibiotics.  相似文献   

20.
The presence of a leaving group at C-4 of monobactams is usually considered to be a requirement for mechanism-based inhibition of human leukocyte elastase by these acylating agents. We report that second-order rate constants for the alkaline hydrolysis and elastase inactivation by N-carbamoyl monobactams are independent of the pKa of the leaving group at C-4. Indeed, the effect exerted by these substituents is purely inductive: electron-withdrawing substituents at C-4 of N-carbamoyl-3,3-diethylmonobactams increase the rate of alkaline hydrolysis and elastase inactivation, with Hammett pI values of 3.4 and 2.5, respectively, which indicate the development of a negative charge in the transition-states. The difference in magnitude between these pI values is consistent with an earlier transition-state for the enzymatic reaction when compared with that for the chemical process. These results suggest that the rate-limiting step in elastase inactivation is the formation of the tetrahedral intermediate, and that beta-lactam ring-opening is not concerted with the departure of a leaving group from C-4. Monobactam sulfones emerged as potent elastase inhibitors even when the ethyl groups at C-3, required for interaction with the primary recognition site, are absent. For one such compound, a 1 : 1 enzyme-inhibitor complex involving porcine pancreatic elastase has been examined by X-ray crystallography and shown to result from serine acylation and sulfinate departure from the beta-lactam C-4.  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