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 46 毫秒
1.
2.
Rate constants for the hydrolysis of L-proline methyl ester to form proline and methanol in D(2)O buffered at neutral pD and 25 degrees C and the deuterium enrichment of the proline product determined by electrospray ionization mass spectrometry are reported. The data give k(DO) = 5.3 +/- 0.5 M(-1) s(-1) as the second-order rate constant for carbon deprotonation of N-protonated proline methyl ester by deuterioxide ion in D(2)O at 25 degrees C and I = 1.0 (KCl). The data provide good estimates of carbon acidities of pK(a) = 21 for N-protonated proline methyl ester and pK(a) = 29 for proline zwitterion in water and of the second-order rate constant k(HO) = 4.5 x 10(-5) M(-1) s(-1) for carbon deprotonation of proline zwitterion by hydroxide ion at 25 degrees C. There is no detectable acceleration of the deprotonation of N-protonated proline methyl ester by the Br?nsted base 3-quinuclidinone in water, and it is not clear that such Br?nsted catalysis would make a significant contribution to the rate acceleration for deprotonation of bound proline at proline racemase. A comparison of the first-order rate constants k(HO)[HO(-)] = 4.5 x 10(-11) s(-1) for deprotonation of free proline zwitterion in water at pH 8 and k(cat) = 2600 s(-1) for deprotonation of proline bound to the active site of proline racemase at pH 8 shows that the enzymatic rate acceleration for proline racemase is ca. 10(13)-fold. This corresponds to a 19 kcal/mol stabilization of the transition state for deprotonation of the enzyme-bound carbon acid substrate by interaction with the protein catalyst. It is suggested that (1) much of the rate acceleration of the enzymatic over the nonenzymatic reaction in water may result from transfer of the substrate proline zwitterion from the polar solvent water to a nonpolar enzyme active site and (2) the use of thiol anions rather than oxygen anions as Br?nsted bases at this putative nonpolar enzyme active site may be favored, because of the smaller energetic price for desolvation of thiol anions than for desolvation of the more strongly solvated oxygen anions.  相似文献   

3.
The sulfomethylation of piperazine and the polyazamacrocycles, [9]aneN3, [12]aneN3, [12]aneN4, and [18]aneN6 with formaldehyde bisulfite in aqueous medium at various pH values is described. The number of methanesulfonate groups introduced into these structures was found to be largely determined by pH. At neutral pH, disubstituted products of [9]aneN3, [12]aneN3, [12]aneN4 are formed and, in the latter case, the trans-1,7-bis(methanesulfonate) isomer was predominant. Similarly, a single, symmetrical trisubstituted product was formed with [18]aneN6 at neutral pH. Monomethanesulfonated products of these same polyaza compounds were formed at more acidic pH's. These sulfomethylated products were used as an entry into a series of mono- and diacetate, phosphonate, and phosphinate derivatives of [9]aneN3, [12]aneN3, and [12]aneN4. The sulfonate groups may be converted to acetates without isolation of intermediates by using cyanide to displace the sulfonate(s) followed by acidic hydrolysis. The aminomethanesulfonates may also be oxidatively hydrolyzed by using aqueous triiodide as a prelude to the preparation of aminomethanephosphonates or aminomethanephosphinates.  相似文献   

4.
The second-order rate constants (kcat/Km) for the beta-glucosidase-catalyzed hydrolysis of aryl beta-D-glucopyranosides show a bell-shaped dependence of pH. The pKas that characterize this dependence are 4.4 (delta Hion approximately equal to 0) and 6.7 (delta Hion approximately equal to 0). In D2O these pKas are increased by 0.5 (+/- 0.1) unit, but there is no solvent isotope effect on the pH-independent second-order rate constant. Nath and Rydon [Nath, R. L., & Rydon, H. N. (1954) Biochem. J. 57, 1-10] examined the kinetics of the beta-glucosidase-catalyzed hydrolysis of a series of substituted phenyl glucosides. We have extended this study to include glucosides with phenol leaving groups of pKa less than 7. Br?nsted plots for this extended series were nonlinear for both kcat/Km and kcat. Br?nsted coefficients for those compounds with leaving groups of pKa greater than 7 (for kcat/Km) or pKa greater than 8.5 (for kcat) were nearly equal to -1.0, indicating substantial negative charge buildup on the leaving group in the transition state. The nonlinearity indicates an intermediate in the reaction. This was confirmed by partitioning experiments in the presence of methanol as a competing glucose acceptor. A constant product ratio, [methyl glucoside]/[glucose], was found with aryl glucoside substrates varying over 16,000-fold in reactivity (V/K), indicative of a common intermediate. Viscosity variation (in sucrose-containing buffers) was used to probe the extent to which the beta-glucosidase reactions are diffusion-controlled.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
A D Hall  A Williams 《Biochemistry》1986,25(17):4784-4790
Values of kcat and Km have been measured for the Escherichia coli alkaline phosphatase catalyzed hydrolysis of 18 aryl and 12 alkyl monophosphate esters at pH 8.00 and 25 degrees C. A Br?nsted plot of log (kcat/Km) (M-1 s-1) vs. the pK of the leaving hydroxyl group exhibits two regression lines: log (kcat/Km) = -0.19 (+/- 0.02) pKArOH + 8.14 (+/- 0.15) log (kcat/Km) = -0.19 (+/- 0.01) pKROH + 5.89 (+/- 0.17) Alkyl phosphates with aryl or large lipophilic side chains are not correlated by the above equations and occupy positions intermediate between the two lines. The observed change in effective charge on the leaving oxygen of the ester (-0.2) is very small, consistent with substantial electrophilic participation of the enzyme with this atom. Cyclohexylammonium ion is a noncompetitive inhibitor against 4-nitrophenyl phosphate substrate at pH 8.00, and neutral phenol is a competitive inhibitor (Ki = 82.6 mM); these data and the 100-fold larger reactivity of aryl over alkyl esters are consistent with the existence of a lipophilic binding site for the leaving group of the substrate. The absence of a major steric effect in kcat/Km for substituted aryl esters confirms that the leaving group in the enzyme--substrate complex points away from the surface of the enzyme. Arguments are advanced to exclude a dissociative mechanism (involving a metaphosphate ion) for the enzyme-catalyzed substitution at phosphorus.  相似文献   

6.
Tertiary sulfonamidomethyl esters of benzylpenicillin (4) were synthesised and evaluated as a new class of potential prodrugs for beta-lactam antibiotics. Their hydrolysis in aqueous buffers was studied by HPLC and reveal a U-shaped pH rate profile with a pH-independent process extending from ca. pH 2 to ca. pH 10. This pathway is characterised by kinetic data that are consistent with a unimolecular mechanism involving rate-limiting iminium ion formation and penicillinoate expulsion. Benzylpenicillin and the corresponding sulfonamide are the ultimate products detected and isolated, indicating that beta-lactam ring opening is much slower than ester hydrolysis. As expected from the high reactivity, benzylpenicillin esters (4) displayed similar in vitro antibacterial activity to benzylpenicillin itself. Compared to the benzylpenicillin derivatives, sulfonamidomethyl esters of benzoic, clofibric and valproic acids display a much higher stability, giving rise to a Br?nsted beta1g value of -0.96 and suggesting that tertiary sulfonamidomethyl esters may be useful prodrugs for carboxylic acid drugs with pKa > 4.  相似文献   

7.
An imidazole-containing tripodal polyamine ligand N(1)-(2-aminoethyl)-N(1)-(2-imidazol-1-ylethyl)-ethane-1,2-diamine (L) was prepared and its dinuclear zinc(II) complex [Zn(L)(H(2)O)](2)(ClO(4))(4).4H(2)O (1) was obtained and examined as a catalyst for the hydrolysis of 4-nitrophenyl acetate (NA). X-ray crystal structure analysis of the complex revealed that the complex features a dinuclear cation unit with a Zn...Zn distance of 8.34A and both Zn(II) centers adopt distorted trigonal-bipyramid geometry. The solution complexation investigation performed at 25 degrees C by means of potentiometric titration revealed that the mononuclear species [ZnL](2+) is predominating in the pH rage of 7.0-9.7 in the solution and the pK(a1) for the Zn-bound water is 8.50+/-0.01. Complex 1 promoted hydrolysis of NA showed a second-order rate constant of 0.046+/-0.004 M(-1)s(-1) at pH 9.0 in 10% (v/v) CH(3)CN aqueous solution at 25 degrees C. The pH-rate profile for the second-order rate constant of NA hydrolysis with complex 1 gave a sigmoidal curve. And the results show that in the hydrolysis process the two Zn(II) centers of the dinuclear deprotonated species do not cooperate with each other and the Zn-bound hydroxide servers as reactive nucleophile toward the ester.  相似文献   

8.
Narine AA  Watson JN  Bennet AJ 《Biochemistry》2006,45(30):9319-9326
The sialidase from Micromonospora viridifaciens has been found to catalyze the hydrolysis of aryl 2-thio-alpha-D-sialosides with remarkable efficiency: the first- and second-order rate constants, kcat and kcat/Km, for the enzyme-catalyzed hydrolysis of PNP-S-NeuAc are 196 +/- 5 s(-1) and (6.7 +/- 0.7) x 10(5) M(-1) s(-1), respectively. A reagent panel of eight aryl 2-thio-alpha-D-sialosides was synthesized and used to probe the mechanism for the M. viridifaciens sialidase-catalyzed hydrolysis reaction. In the case of the wild-type enzyme, the derived Br?nsted parameters (beta(lg)) on kcat and kcat/Km are -0.83 +/- 0.11 and -1.27 +/- 0.17 for substrates with thiophenoxide leaving groups of pKa values > or = 4.5. For the general-acid mutant, D92G, the derived beta(lg) value on kcat for the same set of leaving groups is -0.82 +/- 0.12. When the conjugate acid of the departing thiophenol was < or = 4.5, the derived Br?nsted slopes for both the wild-type and the D92G mutant sialidase were close to zero. In contrast, the nucleophilic mutant, Y370G, did not display a similar break in the Br?nsted plots, and the corresponding values for beta(lg), for the three most reactive aryl 2-thiosialosides, on kcat and kcat/Km are -0.76 +/- 0.28 and -0.84 +/- 0.04, respectively. Thus, for the Y370G enzyme glycosidic C-S bond cleavage is rate-determining for both kcat and kcat/Km, whereas, for both the wild-type and D92G mutant enzymes, the presented data are consistent with a change in rate-determining step from glycosidic C-S bond cleavage for substrates in which the pKa of the conjugate acid of the leaving group is > or = 4.5, to either deglycosylation (kcat) or a conformational change that occurs prior to C-S bond cleavage (kcat/Km) for the most activated leaving groups. Thus, the enzyme-catalyzed hydrolysis of 2-thiosialosides is strongly catalyzed by the nucleophilic tyrosine residue, yet the C-S bond cleavage does not require the conserved aspartic acid residue (D92) to act as a general-acid catalyst.  相似文献   

9.
The mechanism and substrate specificity of the phosphotriesterase from Pseudomonas diminuta have been examined. The enzyme hydrolyzes a large number of phosphotriester substrates in addition to paraoxon (diethyl p-nitrophenyl phosphate) and its thiophosphate analogue, parathion. The two ethyl groups in paraoxon can be changed to propyl and butyl groups, but the maximal velocity and Km values decrease substantially. The enzyme will not hydrolyze phosphomonoesters or -diesters. There is a linear correlation between enzymatic activity and the pKa of the phenolic leaving group for 16 paraoxon analogues. The beta value in the corresponding Br?nsted plot is -0.8. No effect on either Vmax or Vmax/Km is observed when sucrose is used to increase the relative solvent viscosity by 3-fold. These results are consistent with rate-limiting phosphorus-oxygen bond cleavage. A plot of log V versus pH for the hydrolysis of paraoxon shows one enzymatic group that must be unprotonated for activity with a pKa of 6.1. The deuterium isotope effect by D2O on Vmax and Vmax/Km is 2.4 and 1.2, respectively, and the proton inventory is linear, which indicates that only one proton is "in flight" during the transition state. The inhibition patterns by the products are consistent with a random kinetic mechanism.  相似文献   

10.
A M Davis  A C Regan  A Williams 《Biochemistry》1988,27(25):9042-9047
The title esters are demonstrated to be specific substrates of bovine pancreatic ribonuclease A (EC 3.1.27.5). The Br?nsted dependence of kcat/Km at pH 7.50 for the enzyme-catalyzed cyclization versus the pKa of the leaving phenol exhibits two regression lines of almost identical slope for respectively 2-chlorophenols and 2,6-unsubstituted phenols: log kcat/Km = -0.20 pKa ArOH + 5.47 (n = 5, r = 0.957); log kcat/Km = -0.17 pKa ArOH + 5.79 (n = 4, r = 0.965). Comparison of the Br?nsted beta 1g's with that for the standard reaction where imidazole catalyzes the cyclization (beta 1g = -0.59) indicates considerably less development of negative charge on the leaving oxygen in the enzyme case, providing experimental evidence for the hypothesis that electrophilic assistance is involved in catalysis. The existence of two essentially parallel Br?nsted correlations is not reflected in the standard reaction of substrate with imidazole. Modeling studies indicate that the phenyl ring of the substrate can take up a range of positions away from the active site; the presence of ortho chloro substituents considerably restricts the motion of the phenyl leaving group.  相似文献   

11.
Box-Asp-?Ahx-Ser-?Ahx-His-?Ahx-OEt(Linear-6), cyclo(-Asp-?Ahx-Ser-?Ahx-His-?Ahx-)(Cyclic-6) and poly(-Asp-?Ahx-Ser-?Ahx-His-?Ahx-)(Poly-6) were synthesized, and their catalytic actions in the hydrolysis of PNPA were investigated. Linear-6 was prepared by fragment condensations of three peptides having the sequence -Asp(OBzl)-?Ahx-, -Ser(Bzl)-?Ahx- and -His-?Ahx-, respectively, and subsequent debenzylation. Cyclic-6 and Poly-6 were obtained by cyclization and polymerization, respectively, of H-Asp(OBzl)-?Ahx-Ser(Bzl)-?Ahx-His-?Ahx-OH with DPPA, and subsequent deprotections. The reaction velocities in the hydrolysis of PNPA were all proportional to [E] or [S], and all peptides gave the bell-shaped pH-κcat profiles having optima around pH 8.2. The reaction velocity of Cyclic-6 was always larger than that of Linear-6 or Poly-6. The velocities of all reactions increased steadily with rise in temperature, and the Arrhenius' plots from T-kcat relations suggested that the activation energy for the reaction catalysed by Poly-6 is larger than that by Linear- or Cyclic-6. Brief results for the hydrolysis of other substrates, the solvent isotope effect, and the conformational study with c.d. measurements are also reported.  相似文献   

12.
Catalytic mechanism of hamster arylamine N-acetyltransferase 2   总被引:1,自引:0,他引:1  
Wang H  Liu L  Hanna PE  Wagner CR 《Biochemistry》2005,44(33):11295-11306
Arylamine N-acetyltransferases (NATs) catalyze an acetyl group transfer from AcCoA to primary arylamines, hydrazines, and hydrazides and play a very important role in the metabolism and bioactivation of drugs, carcinogens, and other xenobiotics. The reaction follows a ping-pong bi-bi mechanism. Structure analysis of bacterial NATs revealed a Cys-His-Asp catalytic triad that is strictly conserved in all known NATs. Previously, we have demonstrated by kinetic and isotope effect studies that acetylation of the hamster NAT2 is dependent on a thiolate-imidazolium ion pair (Cys-S(-)-His-ImH(+)) and not a general acid-base catalysis. In addition, we established that, after formation of the acetylated enzyme intermediate, the active-site imidazole, His-107, is likely deprotonated at physiological pH. In this paper, we report steady-state kinetic studies of NAT2 with two acetyl donors, acetyl coenzyme A (AcCoA) and p-nitrophenyl acetate (PNPA), and four arylamine substrates. The pH dependence of k(cat)/K(AcCoA) exhibited two inflection points at 5.32 +/- 0.13 and 8.48 +/- 0.24, respectively. The pK(a) at 5.32 is virtually identical with the previously reported pK(a) of 5.2 for enzyme acetylation, reaffirming that the first half of the reaction is catalyzed by a thiolate-imidazolium ion pair in the active site. The inflection point at 8.48 indicates that a pH-sensitive group on NAT2 is involved in AcCoA binding. A Br?nsted plot constructed by the correlation of log k(4) and log k(H)2(O) with the pK(a) for each arylamine substrate and water displays a linear free-energy relationship in the pK(a) range from -1.7 (H(2)O) to 4.67 (PABA), with a slope of beta(nuc) = 0.80 +/- 0.1. However, a further increase of the pK(a) from 4.67 (PABA) to 5.32 (anisidine) resulted in a 2.5-fold decrease in the k(4) value. Analysis of the pH-k(cat)/K(PABA) profile revealed a pK(a) of 5.52 +/- 0.14 and a solvent kinetic isotope effect (SKIE) of 2.01 +/- 0.04 on k(cat)/K(PABA). Normal solvent isotope effects of 4.8 +/- 0.1, 3.1 +/- 0.1, and 3.2 +/- 0.1 on the k(cat)/K(b) for anisidine, pABglu, and PNA, respectively, were also determined. These observations are consistent with a deacetylation mechanism dominated by nucleophilic attack of the thiol ester for arylamines with pK(a) values or=5.5. The general base is likely His-107 because the His-107 to Gln and Asn mutants were found to be devoid of catalytic activity. In contrast, an increase in pH-dependent hydrolysis of the acetylated enzyme was not observed over a pH range of 5.2-7.5. On the basis of these observations, a catalytic mechanism for the acetylation of arylamines by NAT2 is proposed.  相似文献   

13.
1. Dinitrophenol and maleate anions increase VATP on the 'washed', isolated, mitochondrial ATPase. Hydrolyses of iso-GTP and 2'-deoxy ATP are also stimulated, while hydrolyses of other nucleoside triphosphates (ITP, GTP etc.) are not. 2. Preincubation with ATP, iso-GTP or 2'-deoxy ATP results in a metastable enzyme form with a raised V and a reduced Km. Dinitrophenol stimulates both ATP and ITP hydrolyses by this form. 3. The Arrhenius plot of ATP (but not ITP) hydrolysis by the isolated ATPase shows a break at about 18 degrees C, apparently because the rate limiting step of hydrolysis changes as the temperature rises. 4. Adenylyl beta, gamma-imidodiphosphate (AdoPP[NH]P) inhibits ITP hydrolysis in a pseudofirst order reaction. Its binding is competitive with ITP. If the enzyme is preincubated with ATP, the rate of AdoPP[NH]P binding increases. It is concluded that AdoPP[NH]P inhibits by binding to the hydrolytic site of the enzyme. 5. We conclude that ATP hydrolysis is limited by diphosphate release and ITP hydrolysis by bond splitting. Energy release during ATP hydrolysis is maximal at the ATP binding step, and during ITP hydrolysis at bond splitting.  相似文献   

14.
The ability of an arginine residue to function as the active site acid catalyst in the fumarate reductase family of enzymes is now well-established. Recently, a dual role for the arginine during fumarate reduction has been proposed [Mowat, C. G., Moysey, R., Miles, C. S., Leys, D., Doherty, M. K., Taylor, P., Walkinshaw, M. D., Reid, G. A., and Chapman, S. K. (2001) Biochemistry 40, 12292-12298] in which it acts both as a Lewis acid in transition-state stabilization and as a Br?nsted acid in proton delivery. This proposal has led to the prediction that, if appropriately positioned, a water molecule would be capable of functioning as the active site Br?nsted acid. In this paper, we describe the construction and kinetic and crystallographic analysis of the Q363F single mutant and Q363F/R402A double mutant forms of flavocytochrome c(3), the soluble fumarate reductase from Shewanella frigidimarina. Although replacement of the active site acid, Arg402, with alanine has been shown to eliminate fumarate reductase activity, this phenomenon is partially reversed by the additional substitution of Gln363 with phenylalanine. This Gln --> Phe substitution in the inactive R402A mutant enzyme was designed to "push" a water molecule close enough to the substrate C3 atom to allow it to act as a Br?nsted acid. The 2.0 A resolution crystal structure of the Q363F/R402A mutant enzyme does indeed reveal the introduction of a water molecule at the correct position in the active site to allow it to act as the catalytic proton donor. The 1.8 A resolution crystal structure of the Q363F mutant enzyme shows a water molecule similarly positioned, which can account for its measured fumarate reductase activity. However, in this mutant enzyme Michaelis complex formation is impaired due to significant and unpredicted structural changes at the active site.  相似文献   

15.
Mutagenesis of the conserved tyrosine (Y370) of the Micromonospora viridifaciens sialidase to small amino acids changes the mechanism of catalysis from retention of anomeric configuration to inversion [Watson, J. N., et al. (2003) Biochemistry 42, 12682-12690]. For the Y370G mutant enzyme-catalyzed hydrolysis of a series of aryl sialosides and 3'-sialyllactose, the derived Br?nsted parameters (beta(lg)) on k(cat) and k(cat)/K(m) are -0.63 +/- 0.05 and -0.80 +/- 0.08, respectively. Thus, for the Y370G enzyme, glycosidic C-O bond cleavage is rate-determining. Analysis of the activity of the Y370G mutant and wild-type enzymes against a substrate [3,4-dihydro-2H-pyrano[3,2-c]pyridinium alpha-d-N-acetylneuraminide (DHP-alphaNeu5Ac)] whose hydrolysis cannot be accelerated by acid catalysis is consistent with these reactions proceeding via S(N)1 and S(N)2 mechanisms, respectively. The overall structure of the Y370G mutant sialidase active site is very similar to the previously reported wild-type structure [Gaskell, A., et al. (1995) Structure 3, 1197-1205], although removal of the tyrosine residue creates two significant changes to the active site. First, the anomeric oxygen atom of the hydrolysis product (beta-N-acetylneuraminic acid) and four water molecules bind in the large cavity created by the Y370G mutation. Second, the side chain of Asn310 moves to make a strong hydrogen bond to one of the bound water molecules.  相似文献   

16.
Watson JN  Dookhun V  Borgford TJ  Bennet AJ 《Biochemistry》2003,42(43):12682-12690
Mutagenesis of the conserved tyrosine (Y370) of the Micromonospora viridifaciens sialidase changes the mechanism of catalysis from retention of anomeric configuration to an unprecedented inverting mechanism in which water efficiently functions as the nucleophile. Three mutants, Y370A, Y370D, and Y370G, were produced recombinantly in Escherichia coli, and all are catalytically active against the activated substrate 4-methylumbelliferyl alpha-D-N-acetylneuraminide. The Y370D mutant was also shown to catalyze the hydrolysis of natural substrate analogues such as 3'-sialyllactose. A comparison of the pH-rate profiles for the wild-type and the Y370D mutant sialidase reveals no major differences, although with respect to the kinetic term k(cat)/K(m), an ionized form of the aspartate-370 enzyme is catalytically compromised. For the wild-type enzyme, the value of the Br?nsted parameter beta(lg) on k(cat) is 0.02 +/- 0.03, while for the Y370D mutant sialidase beta(lg) = -0.55 +/- 0.03 for the substrates with bad leaving groups. Thus, for the wild-type enzyme, a nonchemical step(s) is rate-limiting, but for the tyrosine mutant cleavage of the glycosidic C-O bond is rate-determining. The Br?nsted slopes derived for the kinetic parameter k(cat)/K(m) display a similar trend (beta(lg) -0.30 +/- 0.04 and -0.74 +/- 0.04 for the wild-type and Y370D, respectively). These results reveal that the tyrosine residue lowers the activation free energy for cleavage of 6'-sialyllactose, a natural substrate analogue, by more than 24.9 kJ mol(-1). Evidence is presented that the mutant sialidases operate by a dissociative mechanism, and the wild-type enzyme operates by a concerted mechanism.  相似文献   

17.
Non-enzymatic thioesterification of specific cysteinyl peptides with fatty acyl-CoA has been previously demonstrated in both liposomes and aqueous medium. To identify the molecular basis for the differential reactivity of polypeptides in aqueous solutions, 26 synthetic cysteinyl peptides encompassing the palmitoylation sites of well known proteins (protein zero, proteolipid protein, beta-adrenergic receptor, p21(K-ras), transferrin receptor, CD-4 and SNAP-25) and six small thiol compounds were incubated separately with [3H]palmitoyl-CoA, [14C]acetyl-CoA and p-nitrophenyl thioacetate (NPTA). For each peptide, both the observed reaction rate constant at pH 7.5 and the pH-independent rate constant (k(2)) were calculated, and reactivity of the attacking sulfhydryl group was characterized using the Br?nsted equation (log k(2)=beta(nuc) pK(a)+C). In general, peptides bearing basic and aromatic amino acid residues showed the lowest thiol pK(a)s, and consequently displayed the highest acylation rates. Reaction with palmitoyl-CoA was complicated to analyze because of the variable partition of peptides in the acyl chain donor/detergent micelles. In contrast, a linear Br?nsted relationship was found for the reaction of the peptides with the water-soluble acetyl-CoA (beta(nuc)=0.59). A similar beta(nuc) value was obtained with the neutral NPTA, indicating that electronic effects other than those responsible for the acid-base properties of the thiol are less important. Thus, the concentration of the thiolate anion appears to be the major factor influencing the rate of the nucleophilic substitution reaction. These findings and the fact that the acylation sites in most proteins are surrounded by basic amino acids may partially explain the specificity of non-enzymatic palmitoylation regarding the acceptor sequences.  相似文献   

18.
Thermolysin activity as well as its stability is remarkably enhanced by high concentration of neutral salts consisting of Na(+), K(+), Cl(-) and Br(-) in the synthesis and hydrolysis of N-carbobenzoxy-L-aspertyl-L-phenylalanine methyl ester and hydrolysis of N-[3-(2-furyl)acryloyl]-glycyl-L-leucine amide (FAGLA) [Inouye, K. (1992) J. Biochem. 112, 335-340]. However, effect of divalent salts on thermolysin activity has not been investigated systematically. In this study, effect of Co(2+) ion on thermolysin activity in the hydrolysis of FAGLA was examined. Thermolysin activity increased 3-4 times with increasing the Co(2+) concentration to 2 mM, but the enhanced activity was considerably reduced with higher Co(2+) concentration (2-18 mM). The activation-and-inhibition dual effects of Co(2+) ion were analysed kinetically. Release of the catalytic Zn(2+) ion from thermolysin, concomitantly occurred with the Co(2+)-dependent activation, was measured with a Zn(2+)-specific fluorescent probe. This indicates that the activation is caused by substituting Co(2+) ion for the catalytic Zn(2+) ion. Meanwhile, the Co(2+)-dependent activation was inhibited competitively by Zn(2+) ion (0.1-1.0 muM) added, similarly to that it is inhibited by higher concentration of Co(2+) ion. These lines of evidence provide a strategy for regulating thermolysin activity with Co(2+) and Zn(2+) ions.  相似文献   

19.
Fersht and co-workers have applied a linear free energy relation (Br?nsted equation) to analyze site-directed mutagenesis experiments involving the enzyme tyrosyl-tRNA synthetase and have suggested that the Br?nsted exponent is linearly correlated with the value of the reaction coordinate at the transition state. We point out that when the mutants differ solely through the formation or deletion of a hydrogen bond away from the reaction center, a linear free energy relation is expected only in limiting cases for which the Br?nsted relation exponent is 0, 1 or infinity. The results may be correlated with a conformational coordinate but not with the development of the reaction coordinate per se.  相似文献   

20.
Irradiation of trans-[RuCl(cyclam)(NO)](2+), cyclam is 1,4,8,11-tetraazacyclotetradecane, at pHs 1-7.4, with near UV light results in the release of NO and formation of trans-[Ru(III)Cl(OH)(cyclam)](+) with pH dependent quantum yields (from approximately 0.01 to 0.16 mol Einstein(-1)) lower than that for trans-[RuCl([15]aneN(4))(NO)](2+), [15]aneN(4) is 1,4,8,12-tetaazacyclopentadecane, (0.61 mol Einstein(-1)). After irradiation with 355 nm light, the trans-[RuCl([15]aneN(4))(NO)](2+) induces relaxation of the aortic ring, whereas the trans-[RuCl(cyclam)(NO)](2+) complex does not. The relaxation observed with trans-[RuCl([15]aneN(4))(NO)](2+) is consistent with a larger quantum yield of release of NO from this complex.  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