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.
The binding of divalent haptens to IgE-class antibodies leads predominantly to their oligomerization into open and closed dimers. Kinetics of the open dimer formation was investigated by fluorescence titrations of Fab fragments of monoclonal DNP-specific IgE antibodies with divalent haptens having different spacer length (gamma = 14-130 A). Binding was monitored by quenching of intrinsic tryptophan emission of the Fab. Addition of divalent haptens with short spacers (gamma = 14-21 A) to the Fabs at rates larger than a distinct threshold value caused a significant decrease of Fab-binding site occupation in the initial phase of the titration. This finding was interpreted to reflect a nonequilibrium state of hapten-Fab-binding. Such nonequilibrium titrations were analyzed by inserting a kinetic model into a theory of antibody aggregation as presented by Dembo and Golstein (Histamine release due to bivalent penicilloyl haptens. 1978. J. Immunol. 121, 345). Fitting of this model to the fluorescence titrations yielded dissociation rate constants of 7.8 x 10(-3) s-1 and 6 x 10(-3) s-1 for the Fab dimers formed by the flexible divalent haptens N alpha, N epsilon-di(dinitrophenyl)-L-lysine (gamma = 16 A) and bis(N beta-2,4-dinitrophenyl-alanyl)-meso-diamino-succinate (gamma = 21 A). Making the simplifying assumption that a single step binding equilibrium prevails, the corresponding dimer formation rate constants were calculated to be 1.9 x 10(5) M-1 s-1 and 1.1 x 10(4) M-1 s-1, respectively. In contrast, all haptens with spacers longer than 40 A (i.e., bis(N alpha-2,4-dinitrophenyl-tri-D-alanyl)-1,7-diamino-heptane, and di(N epsilon-2,4-dinitrophenyl)-6-aminohexanoate-aspartyl-(prolyl)n-L-l ysyl (n = 24, 27, 33) exhibit a relative fast dimerization rate of the Fab fragments (greater than 7 x 10(6) M-1 s-1). These observations were interpreted as being caused by orientational constraints set by the limited solid angle of the reaction between the macromolecular reactants. Thus, ligands having better access to the binding site would react faster.  相似文献   

2.
Cross-linking of antibodies constitutes a widespread initiation signal for their respective effector functions. Cross-linking IgE-class antibodies provide the triggering signal to mast cells for their degranulation process. To obtain a quantitative insight into these cross-linking processes, the interactions between a DNP-specific monoclonal antibody of the IgE class and a series of divalent DNP haptens with spacers of different length and flexibility have been studied by fluorescence titration experiments. These were analyzed by employing the theoretical model developed by Dembo and Goldstein [Dembo, M., & Goldstein, B. (1978) J. Immunol. 121, 345-353] in a fitting procedure. Equilibrium constants that describe the aggregation and ring-closure processes caused by divalent hapten binding have been used as free parameters. The intrinsic binding constants were determined by fluorescence titrations with corresponding monovalent haptens. The main results are the following: (1) The divalent haptens with a short and flexible spacer [i.e., N alpha, N epsilon-di-(DNP)-L-lysine,meso-bis[(DNP-beta-Ala)amino]succinate, and bis[(DNP-tri-D-Ala)amino]heptane, having a maximal DNP-DNP distance of gamma = 14, 21, and 45 A, respectively] effect aggregation of the antibodies mainly into closed dimers. (2) The divalent hapten family with long and rigid oligoproline spacers di(DNP)-Ahx-Asp-(Pro)n-Lys with n = 24, 27, and 33 (i.e., gamma = 100, 110, and 130 A) causes aggregation of the antibodies predominantly into closed dimers and trimers. The corresponding equilibrium constants of the respective ring-closure processes decrease significantly with longer spacer length. (3) Evidence was found that intramolecularly monomeric ring closure of the IgE antibodies is caused by haptens containing oligoproline spacers with n = 37 or 42 (gamma = 130-150 A). The equilibrium constant of the ring-closure process increases with spacer length. This increase in stability indicates a difference in the imposed strain. Furthermore, the latter results imply that the distance between the two binding sites of the IgE molecule lies in the range dictated by the rigid oligoproline part of the respective hapten's spacer, i.e., 115-130 A. (4) Nearly all oligomeric ring-closure processes proceed relatively slowly with an approximate lower limit of a half-life of 5-10 s. This slowing down of the aggregation and ring-closure processes most probably reflects steric factors.  相似文献   

3.
2-Keto-3-deoxygluconate-6-P exists as an euqilibrium of three forms at 25 degrees measurable by 13C NMR: alpha-furanose anomer (41%), beta-furanose anomer (50%), and open chain keto (9%). The three forms are interconverted rapidly (greater than 0.5 s-1) so that the unidirectional rates of furanose ring opening and closing can be quantitated by an NMR line broadening method. The 2-keto-3-deoxygluconate aldolase is specific for only one of these forms, the open chain keto form. The rates for ring opening calculated from the rapid kinetic enzyme system compare closely with those obtained with the NMR method.  相似文献   

4.
P Canioni  P J Cozzone 《Biochimie》1979,61(3):343-354
The low-field portion of the 360 MHz proton N.M.R. spectrum of native porcine pancreatic colipase has been studied as a function of pH over the pH range 2-12. Resonances associated with the 26 protons of the aromatic rings of the two histidines, two phenylalanines and three tyrosines have been identified and tentatively assigned to specific residues. Titrations of pH yielded apparent pKa's of 7.9, 6.9, 10.4, 10.3 and 11.3 for His I (His 30), His II (His 86), Tyr I (Tyr 56 or 57), Tyr II (Tyr 56 or 57) and Tyr III (Tyr 53) respectively (tentative assignments). The high pKa value of His 30 is attributed to the vicinity of Asp 31. The mobility of the aromatic ring of Tyr 53 is hindered and an upper bound of 500 s-1 on the rate of rotation can be estimated. The aromatic rings of the 2 other tyrosine residues and of the 2 phenylalanine residues can rotate freely on the N.M.R. time scale. The study of perturbations in titration profiles and chemical shift values reveals a specific interaction of His 86 with Tyr I and, to a lesser extent, Tyr II. The existence of this interaction indicates that the protein folding brings in close spatial vicinity two distant regions of the covalent structure to form a "hydrophobic-aromatic" site which might be involved in the binding of bile salt micelles to pancreatic colipase.  相似文献   

5.
R D Sheardy  E J Gabbay 《Biochemistry》1983,22(9):2061-2067
A series of diastereomeric dipeptide amides, containing an N-terminal L-lysyl residue and a C-terminal L- or D-amino acid with a derivatized aromatic ring on the side chain, was synthesized to determine the dependence of (1) the chirality of the N-terminal amino acid alpha-carbon and (2) the length of the N-terminal amino acid side chain for intercalation of the aromatic ring. The nature of the complex between the peptide and DNA (i.e., electrostatic, intercalative, or a combination of these) was determined by UV and CD studies, viscometric titrations, and 1H NMR studies. The results of these studies reveal distinct differences in the binding site of the aromatic rings of the various peptides. In particular, the results suggest that the alpha- and epsilon-amino groups of the lysyl residue bind electrostatically to adjacent phosphates on the DNA backbone in a stereospecific manner. As a result of this stereospecificity, the aromatic rings of the peptides with the L-L designation point toward the DNA helix, while those of the peptides of the L-D designation point away from the helix. This is completely consistent with previously reported work [Gabbay, E.J., Adawadkar, P. D., & Wilson, W. D. (1976) Biochemistry 15, 146; Gabbay, E. J., Adawadkar, P. D., Kapicak, L., Pearce, S., & Wilson, W. D. (1976) Biochemistry 15, 152]. The results also indicate a great dependence on the length of the side chain for intercalation of the aromatic ring. Specifically, if the side chain is long enough, and flexible enough, the aromatic ring can fully or partially intercalate, regardless of the chirality of the N-terminal amino acid alpha-carbon. However, if the side chain is too short, only partial intercalation is observed for peptides of the L-D designation, and no intercalation is observed for peptides of the L-D designation.  相似文献   

6.
In the radiolysis of aqueous formate-containing solutions a chain reaction (i, ii) proceeds in the presence of N2O. CO2-. + N2O + H2O----CO2 + N2 + .OH + OH- (i) .OH + HCO2-.----CO2-. + H2O (ii) The chain length depends on the dose rate and the N2O concentration but not on the formate concentration. Typically, G(CO2) approximately 140 molecules (100 eV)-1 is found, with an equivalent amount of N2, at a dose rate of 3 X 10(-3) Gy s-1. The rate constant for the rate-determining step in this chain reaction has been calculated at k(i) = 1600 dm3 mol-1 s-1. The possible relevance of this chain reaction in radiation biological studies is briefly discussed.  相似文献   

7.
E P Lennette  B V Plapp 《Biochemistry》1979,18(18):3933-3938
The reaction of the imidazole group of histidine hydantoin with bromoacetate was studied as a model for carboxymethylation of histidine residues in proteins. pK values of 6.4 and 9.1 (25 degrees C) and apparent heats of ionization of 7.8 and 8.7 kcal/mol were determined for the imidazole and hydantoin rings, respectively. At pH values corresponding to the isoelectric points for histidine hydantoin, the rates of carboxymethylation at 12, 25, 37, and 50 degrees C were determined; the modified hydantoins were hydrolyzed to the corresponding histidine derivatives for quantitative amino acid analysis. At pH 7.72 and 25 degrees C, the imidazole tele-N was alkylated (k = 3.9 X 10(-5) M-1 s-1) twice as fast as the pros-N. The monocarboxymethyl derivatives were carboxymethylated at the same rate at the pros-N (k = 2.1 X 10(-5) M-1 s-1) but 3 times faster at the tele-N (k = 11 X 10(-5) M-1 s-1). The enthalpies of activation determined for carboxymethylation of the imidazole ring and its monocarboxymethyl derivatives were similar (15.9 +/- 0.7 kcal/mol). delta S for the four carboxymethylations was -25 +/- 2 eu. The electrostatic component of delta S (delta S es) was calculated from the influence of the dielectric constant on the reaction rate at 25 degrees C. delta S es was slightly negative (-4 +/- 1 eu) for mono- or dicarboxymethylations, indicating some charge separation in the transition state. The nonelectrostatic entropy of activation was -21 +/- 2 eu for all four carboxymethylations.  相似文献   

8.
The redox reaction between cytochrome c-551 and its oxidase from the respiratory chain of pseudomonas aeruginosa was studied by rapid-mixing techniques at both pH7 and 9.1. The electron transfer in the direction of cytochrome c-551 reduction, starting with the oxidase in the reduced and CO-bound form, is monophasic, and the governing bimolecular rate constants are 1.3(+/- 0.2) x 10(7) M-1 . s-1 at pH 9.1 and 4 (+/- 1) x 10(6) M-1 . s-1 at pH 7.0. In the opposite direction, i.e. mixing the oxidized oxidase with the reduced cytochrome c-551 in the absence of O2, both a lower absorbance change and a more complex kinetic pattern were observed. With oxidized azurin instead of oxidized cytochrome c-551 the oxidation of the c haem in the CO-bound oxidase is also monophasic, and the second-order rate constant is 2 (+/- 0.7) x 10(6) M-1 . s-1 at pH 9.1. The redox potential of the c haem in the oxidase, as obtained from kinetic titrations of the completely oxidized enzyme with reduced azurin as the variable substrate, is 288 mV at pH 7.0 and 255 mV at pH 9.1. This is in contrast with the very high affinity observed in similar titrations performed with both oxidized azurin and oxidized cytochrome c-551 starting from the CO derivative of the reduced oxidase. It is concluded that: (i) azurin and cytochrome c-551 are not equally efficient in vitro as reducing substrates of the oxidase in the respiratory chain of Pseudomonas aeruginosa; (ii) CO ligation to the d1 haem in the oxidase induces a large decrease (at least 80 mV) in the redox potential of the c-haem moiety.  相似文献   

9.
The function of a highly mobile loop in Escherichia coli dihydrofolate reductase was studied by constructing a mutant (DL1) using cassette mutagenesis that had four residues deleted in the middle section of the loop (Met16-Ala19) and a glycine inserted to seal the gap. This part of the loop involves residues 16-20 and is disordered in the X-ray crystal structures of the apoprotein and the NADP+ binary complex but forms a hairpin turn that folds over the nicotinamide moiety of NADP+ and the pteridine moiety of folate in the ternary complex [Bystroff, C., & Kraut, J. (1991) Biochemistry 30, 2227-2239]. The steady-state and pre-steady-state kinetics and two-dimensional 1H NMR spectra were analyzed and compared to the wild-type protein. The kinetics on the DL1 mutant enzyme show that the KM value for NADPH (5.3 microM), the KM for dihydrofolate (2 microM), the rate constant for the release of the product tetrahydrofolate (10.3 s-1), and the intrinsic pKa value (6.2) are similar to those exhibited by the wild-type enzyme. However, the hydride-transfer rate declines markedly from the wild-type value of 950 s-1 to 1.7 s-1 for the DL1 mutant and when taken with data for substrate binding indicates that the loop contributes to substrate flux by a factor of 3.5 x 10(4). Thus, the mobility of loop I may provide a mechanism of recruiting hydrophobic residues which can properly align the nicotinamide and pteridine rings for the hydride-transfer process (a form of transition-state stabilization).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The interaction with closed circular supercoiled and linear DNA of bisphenanthridinium compounds substituted through both the meta and para positions of the 6-phenyl group, along with appropriate monomer intercalators as controls, has been investigated by viscometric titration. When CPK models for the phenanthridinium rings of the three bis-compounds are oriented in a parallel manner as a model for intercalation, their ring plane to ring plane distances are approximately 7 to 8 A (SR 2430), 11 A (SR 2193), and 15 A (SR 2166). In SR 2430 the two phenanthridines are linked through the para positions of the 6-phenyl group; this chain allows intercalation of the two rings at adjacent binding sites in DNA, but is not long enough to accommodate an excluded site. The viscometric titrations with both superhelical and linear DNA clearly indicate that SR 2430 gives results close to those of the monomer control compounds while SR 2193 and SR 2166 have approximately twice the unwinding angle and DNA length increase on binding to DNA as the monomer compounds. These phenanthridinium compounds, therefore, are capable of bisintercalation only if their linking groups are of sufficient length to allow an excluded binding site between base pairs. This conclusion is supported by DNA thermal denaturation experiments in the presence of these compounds.  相似文献   

11.
Myosin light chain kinase is activated by Ca2+/calmodulin. Insights into the kinetic mechanism of this activation by Ca2+/calmodulin have now been obtained using extrinsically labeled fluorescent calmodulin, a fluorescent peptide substrate, and a stopped-flow spectrophotofluorimeter. We employed spinach calmodulin labeled with the sulfhydryl-selective probe, 2-(4-maleimidoanilino)naphthalene-6-sulfonic acid, to measure changes in the fluorescence intensity of the 2-(4-maleimidoanilino)naphthalene-6-sulfonic acid-calmodulin upon binding to rabbit skeletal muscle myosin light chain kinase. The fluorescent peptide substrate KKRAARAC(sulfobenzo-furazan)SNVFS-amide was used to measure kinase activity. Our results showed that the binding interaction could be modeled as a two-step process: a bimolecular reaction with an association rate of 4.6 x 10(7) M-1 s-1 followed by an isomerization with a rate of 2.2 s-1. Phosphorylation of the peptide during stopped-flow experiments could be modeled by a two-step process with a catalytic association rate of 6.5 x 10(6) M-1 s-1 and a turnover rate of 10-20 s-1. Our results also indicated that kinase activity occurred too rapidly for the slower isomerization rate of 2.2 s-1 to be linked specifically to the activation process.  相似文献   

12.
Each actin molecule has one high affinity site which binds a divalent cation. It has been proposed that an isomerization of the actin molecule is involved in divalent cation exchange at this site ("isomerization model," Frieden, C. (1982) J. Biol. Chem. 257, 2882-2886); we have maintained that exchange is by a simple competitive mechanism (Estes, J. E., Selden, L. A., and Gershman, L. C. (1987) J. Biol. Chem. 262, 4952-4957). Here, using fluorescent-labeled actin, we measure the apparent rate constant for exchange (kapp) as a function of the ratio of free Ca2+ and Mg2+ concentrations, ([Ca]/[Mg]), and show that both models are consistent with the data. The major parameter controlling this relationship in the simple competitive exchange model, the ratio of the association rate constants for Ca2+ and Mg2+ to actin (kCa/kMg), is found to have a value of about 90. We have verified this parameter by direct measurements of kCa and kMg, finding that kCa = 1.9 x 10(7) M-1 s-1 and kMg = 2.3 x 10(5) M-1 s-1, consistent with the characteristics of the Ca2+ and Mg2+ aquo ions. The corresponding parameter derived from the isomerization model is not verifiable. We conclude that high affinity divalent cation exchange on actin proceeds by a simple competitive mechanism.  相似文献   

13.
To understand the extent of the cross-linking of proteins by the bifunctional reagent p-NN'-phenylenebismaleimide, a quantitative study of competing reactions has been undertaken. The two reactive maleimide rings of the bismaleimide are hydrolysed in mildly alkaline aqueous solutions much more rapidly than is the single maleimide ring of the monofunctional analogue N-ethylmaleimide. The kinetics of hydrolysis are second-order, depending on both imide and hydroxyl ion concentration in the pH range 8-10. The hydrolysis of the first imide ring of the bismaleimide is more rapid than the second, with second-order rate constants of 1600 M-1 . s-1 and 500 M-1 . s-1 respectively, at 25 degrees C. The half-times for hydrolysis of the first and second imide rings at pH 9.0 are therefore only 43s and 140s. Because it renders the maleimide ring unreactive towards cysteine, this rapid hydrolysis can limit the extent of cross-linking of proteins by the bismaleimide.  相似文献   

14.
Cd2+ blocks whole-cell calcium currents in frog sympathetic neurons by 50% at approximately 300 nM. Strong depolarizations rapidly reverse that blockade (tau = 1.3 ms at +120 mV). Reblocking follows bimolecular kinetics (rate = 1.2 x 10(8) M-1 s-1) at voltages where channels are mostly open (0 to +30 mV). The unblocking rate is approximately 50 s-1, so the dissociation constant calculated from the rate constants is approximately 400 nM. Steady-state block is strong at -80 mV, so closed channels can also be blocked. However, reblocking is extremely slow (tau = 1-2 s) at voltages where the channels are mostly closed. The rates for Cd2+ entry and exit are greater than 100-fold lower for closed channels than for open channels, and closed channels appear to be closed at both ends.  相似文献   

15.
The kinetics of CO and NO recombination with the giant approximately 3600-kDa hexagonal bilayer hemoglobin of Lumbricus terrestris and its subunits, the approximately 200-kDa dodecamer of globin chains (3 x chains (I + II + III + IV] (see preceding paper (Vinogradov S.N., Sharma, P.K., Qabor, A.N., Wall, J.S., Westrick, J.A., Simmons, J.H., and Gill, S.J. (1991) J. Biol. Chem. 266, 13091-13096], the 50-kDa disulfide-bonded trimer (chains II-IV), the monomer (chain I), and the approximately 30-kDa linker (chains VA, VB, and VI), were measured following photolysis over time scales ranging from picosecond to millisecond. CO recombination at 436 nm subsequent to excitation (9 ns) at 532 nm showed three phases covering a 100-fold range for the Hb, dodecamer, trimer, and linker protein. The proportion of the fast phase was 0.1-0.2 for the trimer, dodecamer, and Hb. The relative rates and amplitudes of the phases were not affected by changes in CO concentration or excitation intensity. The monomer showed a single phase with a rate of 2 x 10(6) M-1 s-1. The second-order reaction with NO showed two rates. The faster rate was 90 x 10(6) M-1 s-1 and accounted for approximately 0.7 of the reaction for all species except the monomer, where it accounted for the full reaction. The slower rate was 15 x 10(6) M-1 s-1 for all species except the monomer.  相似文献   

16.
The binding of NADH to porcine mitochondrial malate dehydrogenase in phosphate buffer at pH 7.5 has been studied by equilibrium and kinetic methods. Hyperbolic binding was obtained by fluorimetric titration of enzyme with NADH, in the presence or absence of hydroxymalonate. Identical results were obtained for titrations of NADH with enzyme in the presence or absence of hydroxymalonate, measured either by fluorescence emission intensity or by the product of intensity and anisotropy. The equilibrium constant for NADH dissociation was 3.8 +/- 0.2 micrometers, over a 23-fold range of enzyme concentration, and the value in the presence of saturating hydroxymalonate was 0.33 +/- 0.02 micrometer over a 10-fold range of enzyme concentration. The rate constant for NADH binding to the enzyme in the presence of hydroxymalonate was 3.6 X 10(7) M-1 s-1, while the value for dissociation from the ternary complex was 30 +/- 1 s-1. No limiting binding rate was obtained at pseudo-first order rate constants as high as 200 s-1, and the rate curve for dissociation was a single exponential for at least 98% of the amplitude. In addition to demonstrating that the binding sites are independent and indistinguishable, the absence of effects of enzyme concentration on the KD value indicates that NADH binds with equal affinity to monomeric and dimeric enzyme forms.  相似文献   

17.
1. The oscillations in the peroxidase-oxidase reaction in an open system with NADH as the hydrogen donor are caused by the reaction starting and stopping at critical concentrations of the substrates O2 and NADH. The existence of such critical concentrations is typical of branched chain reactions. 2. The critical concentrations of O2 and NADH that determine the initiation of the reaction are mutually dependent. 3. The branching reactions that determine these critical concentrations involve compounds I and II. 4. Superoxide may be involved in the branching reactions by reacting with NADH and ferriperoxidase. At pH 5.1 the rate constant for the latter reaction is determined as 1.5 . 10(5) M-1 . s-1, whereas for the former reaction only an upper limit for the rate constant of 3.5 . 10(4) M-1 . s-1 could be estimated. These relatively low rate constants suggest that alternative branching reactions may also be involved.  相似文献   

18.
NADH chemistry ancillary to the oscillatory peroxidase-oxidase (PO) reaction has been reexamined. Previously, (NAD)2 has been thought of as a terminal, inert product of the PO reaction. We now show that (NAD)2 is a central reactant in this system. Although we found traces of the dimer after several hours of the PO reaction, no accumulation of the dimer occurred, regardless of the reaction time or the number of oscillations. (NAD)2 can convert horseradish peroxidase (HRP) compound I (CpI) to compound II (CpII) with apparent rate constant (2.7 +/- 0.2) x 105 M-1.s-1 and CpII to HRP at 1 x 105 M-1.s-1. Moreover, a reduction of HRP compound III (CpIII) to CpI by (NAD)2 occurs with a rate constant faster than 5 x 106 M-1.s-1. The (NAD)2 reduction of CpIII provides an alternative to the reduction by NAD radical suggested by Yokota and Yamazaki. HRP catalyzes oxidation of alpha-NADH, not only the beta anomer as previously assumed. Rate constants of alpha- and beta-NADH reactions with CpI are (7.4 +/- 0.4) x 105 M-1.s-1, and (1.7 +/- 0.2) x 105 M-1.s-1, and with CpII are estimated as 5 x 104 M-1.s-1, and 4 x 104 M-1.s-1. Apparent rate constants of reduction of methylene blue (MB) to leuco-methylene blue (MBH) are 3.8 x 104 M-1.s-1 for NADH and 6.4 x 104 M-1.s-1 for NAD dimer, (NAD)2, while reoxidation of MBH proceeds at (2.1 +/- 0.2) x 103 M-1.s-1 All the rates were measured in 0.1 M acetate buffer, pH 5.1.  相似文献   

19.
This study examines mechanisms that regulate the activation of blood coagulation proteases on intact cell membranes. The activation of factor X by factors IXa and VIIIa assembled on viable monocytes is presented as a biologically relevant model for membrane-dependent proteolysis of coagulation zymogens. The hypothesis that this reaction is limited by diffusion was tested by comparing predicted with observed concentration dependence, temperature dependence, and effective rate coefficient. Rates of factor X catalysis were measured using a chromogenic substrate specific for the product, factor Xa. The value of KR and of K1/2, i.e. concentrations giving half-maximal rates in reciprocal functional titrations with substrate and enzyme, respectively, were directly correlated with the concentration of the titrated component. Arrhenius plots constructed over temperatures encompassing 10-35 degrees C were biphasic with downward concavity. Apparent activation energies were 6.01 +/- 0.93 and 35.84 +/- 8.9 kcal/mol for the interval above and below the inflection point, respectively. The effective rate coefficient calculated from apparent kinetic parameters was 3.58 +/- 0.1 x 10(12) M-1 s-1. This rate is similar to the maximal rate of collision between factor X molecules and the monocyte, i.e. 2.9 x 10(12) M-1 s-1 estimated from the steady-state von Smoluchowski equation for uniformly reacting spherical particles. The observed agreement between predicted and experimental results indicates that under biologically relevant conditions, the rate of factor X activation by the intrinsic protease is controlled by diffusion of factor X toward the catalytic site.  相似文献   

20.
The synthetic antithrombin-binding heparin pentasaccharide and a full-length heparin of approximately 26 saccharides containing this specific sequence have been compared with respect to their interactions with antithrombin and their ability to promote inhibition and substrate reactions of antithrombin with thrombin and factor Xa. The aim of these studies was to elucidate the pentasaccharide contribution to heparin's accelerating effect on antithrombin-proteinase reactions. Pentasaccharide and full-length heparins bound antithrombin with comparable high affinities (KD values of 36 +/- 11 and 10 +/- 3 nM, respectively, at I 0.15) and induced highly similar protein fluorescence, ultraviolet and circular dichroism changes in the inhibitor. Stopped-flow fluorescence kinetic studies of the heparin binding interactions at I 0.15 were consistent with a two-step binding process for both heparins, involving an initial weak encounter complex interaction formed with similar affinities (KD 20-30 microM), followed by an inhibitor conformational change with indistinguishable forward rate constants of 520-700 s-1 but dissimilar reverse rate constants of approximately 1 s-1 for the pentasaccharide and approximately 0.2 s-1 for the full-length heparin. Second order rate constants for antithrombin reactions with thrombin and factor Xa were maximally enhanced by the pentasaccharide only 1.7-fold for thrombin, but a substantial 270-fold for factor Xa, in an ionic strength-independent manner at saturating oligosaccharide. In contrast, the full-length heparin produced large ionic strength-dependent enhancements in second order rate constants for both antithrombin reactions of 4,300-fold for thrombin and 580-fold for factor Xa at I 0.15. These enhancements were resolvable into a nonionic component ascribable to the pentasaccharide and an ionic component responsible for the additional rate increase of the larger heparin. Stoichiometric titrations of thrombin and factor Xa inactivation by antithrombin, as well as sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the products of these reactions, indicated that pentasaccharide and full-length heparins similarly promoted the formation of proteolytically modified inhibitor during the inactivation of factor Xa by antithrombin, whereas only the full-length heparin was effective in promoting this substrate reaction of antithrombin during the reaction with thrombin.(ABSTRACT TRUNCATED AT 400 WORDS)  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