scholarly journals Substrate Inhibition in the Hydrolysis of Hippuric Acid Esters by Carboxypeptidase A

1975 ◽  
Vol 53 (2) ◽  
pp. 283-294 ◽  
Author(s):  
Joe Murphy ◽  
John W. Bunting

The dependence of initial velocity upon substrate concentration has been examined in the carboxypeptidase A catalyzed hydrolysis of the following hippuric acid esters (at pH 7.5, 25°, ionic strength O.5): C6H5CONHCH2CO2CHRCO2H: R=CH3; CH2CH3;(CH2)2CH3; (CH2)3CH3; (CH2)5CH3; CH(CH3)2; CH2CH(CH3)2; C6H5; CH2C6H5. All of these esters display marked substrate inhibition of their enzymic hydrolyses. With the exception of R=CH3, the velocity-substrate concentration profiles for each of these esters can be rationalized by the formation of an E.S2 complex which, independent of the alcohol moiety of the ester, reacts approximately 25 times more slowly than the E.S complex. For most of these esters, the formation of E.S2 approximates ordered binding of the substrate molecules at the catalytic and inhibitory sites. While binding at the catalytic site is markedly dependent on the nature of the R group, binding of a second substrate molecule to E.S is not significantly affected by the nature of the R side chain. For R=C6H5, the D ester is neither a substrate nor a competitive inhibitor of the hydrolysis of the L-ester but can replace the L-ester at the binding site which is responsible for substrate inhibition. The kinetic analysis suggests that this behavior of D and L -enantiomers is also typical of the other esters examined (except possibly R=CH3). For R=CH3 only, substrate activation also seems to occur prior to the onset of substrate inhibition at higher substrate concentrations.

1978 ◽  
Vol 56 (16) ◽  
pp. 2188-2193
Author(s):  
John W. Bunting ◽  
Samuel S.-T. Chu

The kinetics of hydrolysis of a series of 10 new hippurate esters (C6H5CONHCH2CO2CRR1CO2H (I)) by bovine pancreatic carboxypeptidase A have been investigated at pH 7.5, 25 °C, and ionic strength 0.5. Pronounced substrate inhibition was displayed by I: R = H, R1 = C6H5(CH2)2, 3-indolylmethyl, 4-HOC6H4CH2, and 4-FC6H4 whereas pronounced substrate activation was observed for I: R = H, R1 = 4-CH3C6H4, 4-C2H5C6H4, 4-C6H5C6H4, 1-naphthyl, 2-naphthyl, and R = R1 = C2H5. In all cases substrate activation and substrate inhibition were shown to be consistent with ES2 complex formation similar to that previously observed for other hippurate esters. Kinetic parameters were evaluated for each ester and it is noted that ail 13 hippurate esters now known to display substrate inhibition have kcat/Km > 106 M−1 min−1, whereas kcat/km < 106 M−1 min−1 for all 9 hippurate esters known to display substrate activation. The enzymic specificity for the R1 unit of I suggests binding of R1 in a 'bent' hydrophobic pocket having a restricted entrance.


1974 ◽  
Vol 52 (23) ◽  
pp. 3829-3836 ◽  
Author(s):  
Joe Murphy ◽  
John W. Bunting

The hydrolyses of the O-hippuryl derivatives of glycolic acid (1a), 2-methyllactic acid (1b), and p-chloromandelic acid (1c) by bovine carboxypeptidase A display substrate activation. The hydrolyses of the latter two esters also display substrate inhibition at high substrate concentrations (>0.03 and >0.05 M respectively). Partial kinetic analyses are presented, and these phenomena are discussed in terms of reaction schemes which involve substrate binding at both activating and inhibiting regulatory sites.The hydrolysis of 1b by this enzyme is the first indication that the presence of a hydrogen atom on the α-carbon atom of the alcohol moiety is not obligatory for ester substrates of carboxypeptidase A. The binding of 1c at the catalytic site is approximately 1000 times weaker than for O-hippurylmandelic acid and indicates a dramatic influence for the p-chloro substituent on the binding of the phenyl ring of the mandelic acid unit.


1974 ◽  
Vol 52 (14) ◽  
pp. 2640-2647 ◽  
Author(s):  
John W. Bunting ◽  
Joe Murphy

The hydrolysis of each of the following esters by bovine carboxypeptidase A has been studied at pH 7.5, 25°, ionic strength 0.5: O-hippuryl-, O-phenaceturyl-, O-aceturyl-, O-(N-methylhippuryl)-, and O-(N-hippurylglycyl)-2-hydroxybutanoic acids, and 2-(3-benzoylpropanoxy)-, 2-benzoxyacetoxy-, and 2-(4-phenylbutanoxy)butanoic acids. Substrate inhibition occurs with only the hippuric and phenaceturic acid esters and in the six other cases simple Michaelis–Menten kinetics are observed. The relatively minor variations in the structures of the acid moieties of these esters lead to quite large variations in Km, although kcat seems to be relatively independent of the nature of the acid moiety. Binding modes of substrate molecules at both the catalytic and inhibitory sites are discussed in the light of these observations.


Biochemistry ◽  
1976 ◽  
Vol 15 (15) ◽  
pp. 3237-3244 ◽  
Author(s):  
John W. Bunting ◽  
Samuel S. T. Chu

1975 ◽  
Vol 53 (13) ◽  
pp. 1993-2004 ◽  
Author(s):  
John W. Bunting ◽  
Chester D. Myers

The anions of each of the following carboxylic acids exhibit uncompetitive inhibition of the hydrolysis of O-hippuryl-L-3-phenyllactic acid by bovine carboxypeptidase A at pH 7.5, 25°, ionic strength 0.2: hippuric acid, p-chloro- and p-nitrohippuric acids, hippurylglycine, carbobenzoxyglycine, phenaceturic acid, N'-(3-phenylpropanoyl)glycine, benzoxyacetic acid, 3-benzoylpropanoic acid, and O-hippuryl-D-mandelic acid. In each case, this uncompetitive inhibition is consistent with the ordered binding of substrate and inhibitor to the enzyme; i.e. the inhibitor binds to E.S but not to the free enzyme. Evidence is presented for the binding site for uncompetitive inhibitors being the same as for inhibitory ester substrate molecules. Comparison of the specificities of uncompetitive inhibitors and esters which display substrate inhibition provides evidence for a critical conformational change which controls the binding of uncompetitive inhibitors and inhibitory substrate molecules.D-Phenylalanine, D-leucine, D-p-nitrophenylalanine, glycyl-L-tyrosine, glycyl-L-phenylalanine, and glycyl-L-leucine are competitive inhibitors of the enzymic hydrolysis of O-hippuryl-L-3-phenyllactic acid, whereas the N-chloroacetyl derivatives of L-tyrosine, L-phenylalanine, and L-leucine are noncompetitive inhibitors. For the above D-amino acids, glycyl dipeptides, and N-chloroacetyl amino acids, the phenylalanine derivative in each case is a considerably stronger inhibitor than the corresponding leucine derivative. This preference is similar to that observed for the binding of peptide substrates but the reverse of that observed for ester substrates and simple mono- and dicarboxylate ion inhibitors.The peptide substrates carbobenzoxyglycylglycyl-L-phenylalanine and N-chloroacetyl-L-phenylalanine are noncompetitive inhibitors of the enzymic hydrolysis of O-hippuryl-L-3-phenyllactic acid. This clearly demonstrates the presence of different ester and peptide binding sites in this enzyme, which is consistent with conclusions from recent studies in other laboratories.


1987 ◽  
Vol 65 (8) ◽  
pp. 717-725 ◽  
Author(s):  
John F. Sebastian ◽  
Richard S. Hinks ◽  
Ralf V. Reuland

A variety of modifiers of carboxypeptidase A (CPA) have been investigated in an effort to understand the structural requirements of inhibitors and activators of peptidase activity. It is proposed that an understanding of the mechanism of action of reversible activators of the enzyme may bear on the long standing question of whether the detailed mechanism of peptidase activity is different from that of esterase activity. An analog of the activator 2,2-dimethyl-2-silapentane-5-sulfonate, 5,5-dimethylhexanoate, was found to be a competitive inhibitor of the CPA-catalyzed hydrolysis of benzoylglycyl-L-phenyl-alanine. The modifier 4-phenyl-3-butenoate (styrylacetic acid) was determined to be an activator. The sulfonates benzene-sulfonate, p-toluenesulfonate, phenylmethanesulfonate, 2-phenylethanesulfonate, and 3-phenylpropanesulfonate were all found to be activators.


1968 ◽  
Vol 107 (1) ◽  
pp. 97-102 ◽  
Author(s):  
D. T. Elmore ◽  
J. J. Smyth

1. α-N-Methyl-α-N-toluene-p-sulphonyl-l-lysine β-naphthyl ester (MTLNE) was synthesized as its hydrobromide and shown to be slowly hydrolysed by bovine pancreatic trypsin. The acylation step, however, is so much faster than deacylation of the acyl-enzyme that spectrophotometric measurement of the ‘burst’ of β-naphthol provides a convenient method for determining the absolute molarity of trypsin solutions. 2. By using the same stock solution of trypsin, application of this method at pH4·0 and pH7·0 as well as that of Bender et al. (1966) at pH3·7 gave concordant results. 3. Provided that [S]0>[E]0, the size of the ‘burst’ is independent of substrate concentration. 4. In the trypsin-catalysed hydrolysis of α-N-toluene-p-sulphonyl-l-arginine methyl ester, MTLNE functions as a powerful non-competitive inhibitor. 5. There is no detectable reaction between MTLNE and either bovine pancreatic α-chymotrypsin at pH4·0 or bovine thrombin at pH6·0.


1922 ◽  
Vol 4 (5) ◽  
pp. 487-509 ◽  
Author(s):  
John H. Northrop

1. The velocity of hydrolysis of gelatin by trypsin increases more slowly than the gelatin concentration and finally becomes nearly independent of the gelatin concentration. The relative velocity of hydrolysis of any two substrate concentrations is independent of the quantity of enzyme used to make the comparison. 2. The rate of hydrolysis is independent of the viscosity of the solution. 3. The percentage retardation of the rate of hydrolysis by inhibiting substances, is independent of the substrate concentration. 4. There is experimental evidence that the enzyme and inhibiting substance are combined to form a widely dissociated compound. 5. If the substrate were also combined with the enzyme, an increase in the substrate concentration should affect the equilibrium between the enzyme and the inhibiting substance. This is not the case. 6. The rate of digestion of a mixture of casein and gelatin is equal to the sum of the rates of hydrolysis of the two substances alone, as it should be if the rate is proportional to the concentration of free enzyme. This contradicts the saturation hypothesis. 7. If the reaction is followed by determining directly the change in the substrate concentration, it is found that this change agrees with the law of mass action; i.e., the rate of digestion is proportional to the substrate concentration.


Sign in / Sign up

Export Citation Format

Share Document