scholarly journals The pH-dependence of second-order rate constants of enzyme modification may provide free-reactant pKa values

1977 ◽  
Vol 167 (3) ◽  
pp. 859-862 ◽  
Author(s):  
K Brocklehurst ◽  
H B F Dixon

1. Reactions of enzymes with site-specific reagents may involve intermediate adsorptive complexes formed by parallel reactions in several protonic states. Accordingly, a profile of the apparent second-order rate constant for the modification reaction (Kobs., the observed rate constant under conditions where the reagent concentration is low enough for the reaction to be first-order in reagent) against pH can, in general, reflect free-reactant-state molecular pKa values only if a quasi-equilibrium condition exists around the reactive protonic state (EHR) of the adsorptive complex. 2. Usually the condition for quasi-equilibrium is expressed in terms of the rate constants around EHR: (formula: see text) i.e. k mod. less than k-2. This often cannot be assessed directly, particularly if it is not possible to determine kmod. 3. It is shown that kmod. must be much less than k-2, however, if kobs. (the pH-independent value of kobs.) less than k+2. 4. Since probable values of k+2 greater than 10(6)M-1.S-1 and since values of kobs. for many modification reactions less than 10(6)M-1.S-1, the equilibrium assumption should be valid, and kinetic study of such reactions should provide reactant-state pKa values. 5. This may not apply to catalyses, because for them the value of kcat./Km may exceed 5 X 10(5)M-1.S-1. 6. The conditions under which the formation of an intermediate complex by parallel pathways may come to quasi-equilibrium are discussed in the Appendix.

1991 ◽  
Vol 275 (2) ◽  
pp. 335-339 ◽  
Author(s):  
H C Hawkins ◽  
R B Freedman

1. The number of reactive thiol groups in mammalian liver protein disulphide-isomerase (PDI) in various conditions was investigated by alkylation with iodo[14C]acetate. 2. Both the native enzyme, as isolated, and the urea-denatured enzyme contained negligible reactive thiol groups; the enzyme reduced with dithiothreitol contained two groups reactive towards iodoacetic acid at pH 7.5, and up to five reactive groups were detectable in the reduced denatured enzyme. 3. Modification of the two reactive groups in the reduced native enzyme led to complete inactivation, and the relationship between the loss of activity and the extent of modification was approximately linear. 4. Inactivation of PDI by alkylation of the reduced enzyme followed pseudo-first-order kinetics; a plot of the pH-dependence of the second-order rate constant for inactivation indicated that the essential reactive groups had a pK of 6.7 and a limiting second-order rate constant at high pH of 11 M-1.s-1. 5. Since sequence data on PDI show the presence within the polypeptide of two regions closely similar to thioredoxin, the data strongly indicate that these regions are chemically and functionally equivalent to thioredoxin. 6. The activity of PDI in thiol/disulphide interchange derives from the presence of vicinal dithiol groups in which one thiol group of each pair has an unusually low pK and high nucleophilic reactivity at physiological pH.


1984 ◽  
Vol 62 (3) ◽  
pp. 580-585 ◽  
Author(s):  
Parminder S. Surdhar ◽  
Rizwan Ahmad ◽  
David A. Armstrong

Spectral changes and rates of reaction of flavins and several radical species have been investigated at pH 7, 9, and 10 in the presence and absence of β-mercaptoethanol. The radicals •CO2−, eaq−, and [Formula: see text] reacted with FAD at pH 10 to give a spectrum of FAD •Fl− with rate constants of 7 ± 1 × 108 and 4 ± 1 × 108 M−1 s−1 for •CO2− and [Formula: see text] respectively. At pH 7 only •FlH was observed and at pH 9 a mixture of •FIH and •Fl−.Interactions between flavin radicals and sulphydryl at 10−4 M concentration did not cause perturbations in the uv–visible spectra until either the radical and/or the sulphydryl were ionized. With FAD at pH 9 or 10 and LFl at pH 10 the 370 nm peak of •Fl− was enhanced by about 15% and a second larger growth occurred near 450 nm in the presence of 10−4 to 10−2 M sulphydryl. We attribute this to the formation of labile intermediate RSHFl•−, which must also be involved in the reduction of Fl by [Formula: see text] at pH 9 or 10.The second order rate constant k13 for reaction of [Formula: see text] with FAD at pH 9 and 10 was found to be 4.2 ± 0.5 × 108 M−1 s−1 and 2.0 ± 0.4 × 108 M−1 s−1 respectively. The rate constant for the reaction between [Formula: see text] and LFl at pH 10 was slightly faster, 7 ± 1 × 108 M−1 s−1, probably reflecting the fact that LFl lacks the bulky negatively charged adenine dinucleotide group of FAD.


1983 ◽  
Vol 49 (03) ◽  
pp. 193-195 ◽  
Author(s):  
Torbjörn Nilsson

SummaryThe kinetics of the reaction between human plasma kallikrein and CĪ-esterase inhibitor was studied in a purified system. By monitoring the inhibition reaction for extended periods of time, it was found to proceed in two consecutive steps, a fast reversible second-order binding step followed by a slower, irreversible first-order transition. The rate constants in this reaction model were determined, as well as the dissociation constant of the initial, reversible enzyme-inhibitor complex. Thus, at 37° C the second-order rate constant was found to be 5 · 104 M -1 · s-1, the first order rate constant was 5 · 10-4 s-1 and the dissociation constant K was 1.5 · 10-8 M. Heparin (28 U/ml) and 6-aminohexanoic acid (10 mM) had no effect on the k1 of the of the reaction.


Blood ◽  
2008 ◽  
Vol 112 (11) ◽  
pp. 2029-2029
Author(s):  
S. Paul Bajaj ◽  
Amanda Sutton ◽  
Sreejesh Shanker ◽  
Amy E Schmidt ◽  
Sayeh Agah ◽  
...  

Abstract Factor VII (FVII) consists of an N-terminal γ-carboxyglutamic acid (Gla) domain followed by two epidermal growth factor-like (EGF1 and EGF2) domains and the C-terminal protease domain. Activation of FVII results in a two-chain FVIIa molecule consisting of a light chain (Gla-EGF1-EGF2 domains) and a heavy chain (protease domain) held together by a single disulfide bond. The complex of tissue factor (TF) and FVIIa activates FIX and FX during coagulation. FVIIa on its own is structurally more “zymogen-like” and when bound to TF it is more “active enzyme-like.” We obtained crystal structures of EGR-VIIa/soluble (s) TF (2.9 Å resolution), dansyl-EGR-VIIa/sTF (1.9 Å resolution) and benzamidine-VIIa/sTF (1.6 Å resolution). We also investigated the effect of TF binding on the S1, S2, and S3/S4 subsites (Schechter and Berger, BBRC, 27:157-162, 1967) in FVIIa. The affinity of variously inhibited FVIIa to sTF was also measured using Biacore technology. For obtaining second order inhibition rate constants, FVIIa ± soluble (s) TF was incubated with each inhibitor for various times, diluted several fold and assayed for the residual FVIIa activity. The second order rate constants were obtained by plotting the first order rate constants versus the inhibitor concentrations. These data are summarized in the table below. From these data it appears that all subsites are affected upon FVIIa binding to sTF. Since in the crystal structure of EGR-VIIa/sTF the P1 Arg residue is the only residue that makes contact with FVIIa, it follows that the S1 site is affected ~10-fold upon binding to sTF. Adding a dansyl group that partially occupies the S3/S4 position (1.9 Å structure) increases the second order rate constant 7-fold (2.41 versus 0.35) over that of EGR-ck. Moreover the addition of Pro (DFPR-ck) or Phe (DFFR-ck) residue occupying the S2 position increases the second order rate constant 357-fold and 1500- fold, respectively (125 and 525 versus 0.35). Thus, comparison of dEGR, DFPR, DFFR inhibition suggests that FVIIa prefers Phe at S2 and at S3/S4 positions, and that TF opens up the S1/S2/S3/S4 sites for substrate or inhibitor occupancy. These data are consistent with LTR (P3/P2/P1) residues in FX at its activation cleavage site as well as with LTR (P3/P2/P1) residues and FTR (P3/P2/P1) residues at the 145-146 and 180-181 FIX activation cleavage sites, respectively. Thus, these studies with chloromethylketone inhibitors have biologic relevance. For Biacore studies, sTF was amine coupled to a CM5 chip. The binding of unoccupied active site FVIIa in 5 mM calcium to sTF was characterized by a KD of 7 nM. Benzamidine (10 mM)-VIIa, p-aminobenzamidine (pAB, 1 mM)-VIIa, EGR-VIIa, dEGR-VIIa, DFPR-VIIa and DFFR-VIIa each bound to sTF with KD values ranging from 1- 2 nM. These affinity measurements indicate that the S1 site occupied FVIIa molecule (benzamidine-FVIIa, pAB-VIIa) has essentially the same conformation as the S1/S2/S3/S4 occupied FVIIa. This conclusion is consistent with similar crystal structures of variously inhibited FVIIa molecules complexed with sTF. The differential rates of incorporation of various chloromethylketone inhibitors could be due to the interaction of various residues (P1, P2, P3, P4) with the corresponding active subsites (S1/S2/S3/S4) of FVIIa. Additionally, the rate of incorporation of chloromethylketone inhibitors into FVIIa also involves the irreversible alkylation step, which could be faster for DFFR-ck and DFPR-ck. Once these inhibitors are incorporated, it appears that they induce the same conformation in FVIIa as achieved by S1 site occupancy alone. Thus S1 site occupancy in FVIIa induces the required conformation to modestly increase the affinity for TF. Second Order Rate Constants for Inhibition of FVIIa ± sTF with Various Chloromethylketone (ck) Inhibitors Inhibitor Minus sTF k (min−1 mM−1) Plus sTF k (min−1 mM−1) Fold Difference EGR-ck 0.04 0.35 8.8 dansyl EGR-ck 0.07 2.41 34.4 (D)FPR-ck 2.3 125 54.3 D)FFR-ck 5.6 525 93.8


1985 ◽  
Vol 63 (10) ◽  
pp. 2673-2678 ◽  
Author(s):  
Robert A. McClelland ◽  
Claude Moreau

Hydrolysis kinetics are reported for four spiro ortho esters: 3,4-dihydro-6-methoxy-1H-2-benzopyran-1-spiro-2′-1′,3′-dioxolane (13), its 1′,3′-dioxane analog (14), and the 6-unsubstituted versions of each (11 and 12). For comparison, also included are the diethoxy analogs: 1,1-diethoxy-3,4-dihydro-6-methoxy-1H-2-benzopyran (10) and the 6-unsubstituted compound (9). Product analysis implicates an initial opening of the dioxolane or dioxane ring in the spiro ortho esters, as expected on the basis of stereoelectronic considerations. The intermediate dialkoxycarbocations can be observed in HCl solutions. A detailed analysis has been carried out for the 6-methoxy systems to provide the rate constants k1, the second-order rate constant for H+-catalyzed formation of the cation from the ortho ester, k2, the first-order rate constant for water addition to the cation, and k−1, the first-order rate constant for ring closing of the cation to reform the ortho ester. The two spiro ortho esters are shown in this analysis to undergo reversible ring opening in their hydrolysis, in that values of k−1, are greater than k2. The differences, however, are not large, k−1/k2 being 1.2 (dioxolane, 13) and 3.8 (dioxane, 14). Comparison with the diethoxy ortho ester also reveals that the ring opening process (k1, rate constants) is inherently more difficult with the dioxolane, although not with the dioxane. An argument involving lone pair orientation is advanced to explain this.


1976 ◽  
Vol 159 (3) ◽  
pp. 513-527 ◽  
Author(s):  
J M Armstrong ◽  
D R Trentham

D-Glyceraldehyde 3-phosphate forms adducts with thiols. These adducts, which are presumed to be hemithioacetals, equilibrate rapidly with the unhydrated form of the aldehyde, which is the subtrate for D-glyceraldehyde 3-phosphate dehydrogenase. The adduct provides a substrate buffer system whereby a constant low free aldehyde concentration can be maintained during the oxidation of aldehyde by the enzyme and NAD+. With this system, the kinetics of the association of the aldehyde with the enzyme were examined. The rate profile for this reaction is a single exponential process, showing that all four active sites of the enzyme have equivalent and independent reactivity towards the aldehyde, with an apparent second-order rate constant of 5 × 10(7)M-1-S-1 at pH8.0 and 21 degrees C. The second-order rate constant becomes 8 × 10(7)M-1-S-1 when account is taken of the forward and reverse catalytic rate constants of the dehydrogenase. The pH-dependence of the observed rate constant is consistent with a requirement for the unprotonated form of a group of pK 6.1, which is the pK observed for second ionization of glyceraldehyde 3-phosphate. The rate of phosphorolysis of the acyl-enzyme intermediate during the steady-state oxidative phosphorylation of the aldehyde was studied, and is proportional to the total Pi concentration up to at least 1 mM-Pi at pH 7.5. The pH-dependence of the rate of NADH generation under these conditions can be explained by the rate law d[NADA]/dt = k[acy] holoenzyme][PO4(3-)-A1, where thioester bond, although kinetically indistinguishable rate equations for the reaction are possible. The rates of the phosphorolysis reaction and of the aldehyde-association reaction decrease with increasing ionic strength, suggesting that the active site of the enzyme has cationic groups which are involved in the reaction of the enzyme with anionic substrates.


1973 ◽  
Vol 51 (4) ◽  
pp. 582-587 ◽  
Author(s):  
M. L. Cotton ◽  
H. B. Dunford

In order to investigate the nature of compounds I and II of horseradish peroxidase, the kinetics were studied of ferrocyanide oxidation catalyzed by these compounds which were prepared from three different oxidizing agents. The pH dependence of the apparent second-order rate constant for ferrocyanide oxidation by compound I, prepared from ethyl hydroperoxide and m-chloroperbenzoic acid, was interpreted in terms of an ionization on the enzyme with a pKa = 5.3, identical to that reported previously for hydrogen peroxide. The second-order rate constant for the compound II-ferrocyanide reaction also showed the same pH dependence for the three oxidizing substrates. However, with more accurate results, the compound II-ferrocyanide reaction was reinterpreted in terms of a single ionization with pKa = 8.5. The same dependence of ferrocyanide oxidation on pH suggests structurally identical active sites for compounds I and II prepared from the three different oxidizing substrates.


1986 ◽  
Vol 64 (6) ◽  
pp. 1179-1183 ◽  
Author(s):  
Clifford A. Bunton ◽  
Angela Cuenca

Cationic micelles of cetyltrimethylammonium chloride and bromide (CTACl and CTABr) speed attack of water upon the 2,2′,4,4′,4″-pentamethoxytrityl cation by a factor of ca. 5. The first-order rate constant in water is 5.51 s−1 at 25.0 °C. Anionic micelles of sodium dodecyl sulfate (SDS) have little effect on this reaction, but they strongly inhibit attack of OH−. In water, second-order rate constants for attack of OH−, CN−, and N3− are, respectively, 235, 177, and 2.8 × 105 M−1 s−1. Rate constants of reaction in CTACl go through maxima with increasing [surfactant] and analysis of the data shows that second-order rate constants at the micellar surface are similar to those in water.


1992 ◽  
Vol 287 (3) ◽  
pp. 881-889 ◽  
Author(s):  
M Patel ◽  
I S Kayani ◽  
W Templeton ◽  
G W Mellor ◽  
E W Thomas ◽  
...  

1. 2-(N′-Acetyl-D-phenylalanyl)hydroxyethyl 2′-pyridyl disulphide (compound IV) (m.p. 59 degrees C; [alpha]D20 -6.6 degrees (c 1.2 in methanol)) was synthesized. 2. The results of a study of the pH-dependence of the second-order rate constant (k) for its reaction with the catalytic-site thiol group (Cys-25) of papain (EC 3.4.22.2) together with analogous kinetic data for the reactions of related time-dependent inhibitors, notably the L-enantiomer of compound (IV) (compound III) and the L- and D-enantiomers of 2-(N′-acetylphenylalanylamino)ethyl 2′-pyridyl disulphide (compounds I and II respectively), were used to assess the contributions of the (P1)-NH ... O = C < (Asp-158) and (P2) > C = O ... H-N-(Gly-66) hydrogen bonds and enantiomeric P2-S2 hydrophobic contacts in two manifestations of dynamic molecular recognition in papain-ligand association: (a) signalling to the catalytic-site region to provide for a (His-159)-IM(+)-H-assisted transition state and (b) the dependence of P2-S2 stereoselectivity on hydrogen-bonding interactions outside the S2 subsite. The analysis involved determination of the reactivities of individual ionization states of the reactions (pH-independent rate constants, k) and associated macroscopic pKa values and difference kinetic specificity energies (delta delta GKS = -RT1n(k1/k2), where k1 is the pH-independent second-order rate constant for reaction with one inhibitor and k2 is the analogous rate constant in the same ionization state for reaction with another inhibitor so that, when the structural change provides that k2 > k1, delta delta GKS is positive. 3. The kinetic data further illuminate the nature of the interdependence of binding interactions in papain first noted by Kowlessur, Topham, Thomas, O'Driscoll, Templeton & Brocklehurst [(1989) Biochem. J. 258, 755-764] in the S2 subsite, S1-S2 intersubsite and catalytic-site regions. Of particular note is the apparent dependence of the binding of the N-Ac-D-Phe moiety on the binding of the leaving group to (His-159)-Im+H and the fact that the resulting rate enhancement is more effective when (P1)-N-H is absent than when it is present. This result revealed by kinetic analysis goes beyond the conclusion suggested by model building that it is possible to make all of the binding contacts in complexes involving the D-enantiomers [(II) and (IV)] as in those involving the L-enantiomers [(I) and (III)].(ABSTRACT TRUNCATED AT 400 WORDS)


Sign in / Sign up

Export Citation Format

Share Document