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 antibodies, HyHEL-10 and HyHEL-26 (H10 and H26, respectively), share over 90% sequence homology and recognize with high affinity the same epitope on hen egg white lysozyme (HEL) but differ in degree of cross-reactivity with mutant lysozymes. The binding kinetics, as measured by BIAcore surface plasmon resonance, of monovalent Fab from both Abs (Fab10 and Fab26) to HEL and mutant lysozymes are best described by a two-step association model consistent with an encounter followed by docking that may include conformational changes. In their complexes with HEL, both Abs make the transition to the docked phase rapidly. For H10, the encounter step is rate limiting, whereas docking is also partially rate limiting for H26. The forward rate constants of H10 are higher than those of H26. The docking equilibrium as well as the overall equilibrium constant are also higher for H10 than for H26. Most of the free energy change of association (Delta G degrees) occurs during the encounter phase (Delta G1) of both Abs. H10 derives a greater amount and proportion of free energy change from the docking phase (Delta G2) than does H26. In the H10--HEL(R21Q) complex, a significant slowing of docking results in lowered affinity, a loss of most of Delta G2, and apparently faster dissociation. Slower encounter and docking cause lowered affinity and a loss of free energy change primarily in the encounter step (Delta G1) of H26 with mutant HEL(R21Q). Overall, in the process of complex formation with lysozyme, the mutations HEL(R21X) affect primarily the docking phase of H10 association and both phases of H26. Our results are consistent with the interpretation that the free energy barriers to conformational rearrangement are highest in H26, especially with mutant antigen.  相似文献   

2.
Using BIACORE SPR, we have examined the mechanism of temperature effects on the binding kinetics of two closely related antibody Fabs (H10 and H26) which recognize coincident epitopes on hen egg-white lysozyme (HEL), and whose association and dissociation kinetics are best described by the two-step conformational change model which we interpret as molecular encounter and docking. Time-course series data obtained at a series of six temperatures (6, 10, 15, 25, 30 and 37 degrees C) showed that temperature differentially affects the rate constants of the encounter and docking steps. Docking is more temperature-sensitive than the encounter step, and energetically less favorable at higher temperatures. At elevated temperatures, the time required for docking is longer and the apparent increase in off-rate reflects the greater proportion of the molecules failing to dock and remaining in the less stable encounter state. As a consequence, distribution of free energy change between the encounter and docking steps is altered. At physiological temperature (37 degrees C) the docking step of the H26 complex is energetically unfavorable and most complexes essentially do not dock. There is a significant decrease in total free energy change of the H26 complex at higher temperatures. Elevated temperature changes the rate-limiting step of H26--HEL association from the encounter to the docking step, but not that of H10--HEL. Our results indicate that the mechanism by which elevated temperature reduces the affinities of antigen--antibody complexes is to decrease the net docking rate, and/or stability of the docked complex; at higher temperatures, a smaller proportion of the complexes actually anneal to a more stable docked state. This mechanism may have broad applicability to other receptor--ligand complexes.  相似文献   

3.
4.
We studied the role of the 2 salt bridges (Asp143-Arg147 and Asp146-Arg150) in helix 1 of mouse prion protein (PrP) on the formation of the complex between PrP and the monoclonal antibody T2. We introduced 6 charge-changing mutations to the amino acid residues associated with the salt bridges. Analysis of the circular dichroism spectra of the mutant PrPs showed that the salt bridge mutations did not change the secondary structures. We analyzed the kinetics of the association and dissociation of the PrPs with the T2 antibody. The results showed that the association kinetics were not significantly different among the variants except Arg150Lys, while the dissociation rate of the neutralized-charge variants was 2 orders of magnitude higher than that of the wild type. These results indicate that salt bridges make the interaction of PrP with T2 tighter by slowing down dissociation.  相似文献   

5.
We have studied the rate and character with which testosterone (T) and 5 alpha-dihydrotestosterone (DHT) dissociate from the androgen receptor both within intact cultured genital skin fibroblasts of a subject with 5 alpha-reductase deficiency and after the androgen-receptor complexes have been extracted from the cells. Within the cells, the kinetics of the dissociative process for each hormone was first-order, but T dissociated four times faster than DHT. An Arrhenius plot of the variation of the dissociation rate constants with temperature for T was linear and yielded an activation energy of 28 kcal/mol. This value is identical with the one previously determined for activated DHT-receptor complexes. T-receptor complexes extracted from the cells dissociated with complex kinetics: at 37 degrees C the rate constants of the "fast" and "slow" components were 40 and 14 X 10(-3) min-1, respectively. In contrast, DHT-receptor complexes extracted from the cells dissociated with first-order kinetics and at a rate identical to that observed within cells, except after exposure to pyridoxal 5'-phosphate (5 mM) or concentration by Amicon (B-15) filtration, when their dissociation kinetics became complex. We interpret these data to mean that, within the cells, both T- and DHT-receptor complexes exist predominantly in the activated state whereas, when extracted from the cells, DHT-receptor complexes remain activated, unless perturbed, while T-receptor complexes become unstable spontaneously, probably by reverting to a preactivated state.  相似文献   

6.
Bombarda E  Roques BP  Mély Y  Grell E 《Biochemistry》2005,44(19):7315-7325
The kinetics of Zn(2+) binding by two point-mutated forms of the HIV-1 NCp7 C-terminal zinc finger, each containing tridentate binding motif HCC [Ser49(35-50)NCp7] or CCC [Ala44(35-50)NCp7], has been studied by stopped-flow spectrofluorimetry. Both the formation and dissociation rate constants of the complexes between Zn(2+) and the two model peptides depend on pH. The results are interpreted on the basis of a multistep reaction model involving three Zn(2+) binding paths due to three deprotonated states of the coordinating motif, acting as monodentate, bidentate, and tridentate ligands. For Ser49(35-50)NCp7 around neutral pH, binding preferentially occurs via the deprotonated Cys36 in the bidentate state also involving His44. The binding rate constants for the monodentate and bidentate states are 1 x 10(6) and 3.9 x 10(7) M(-)(1) s(-)(1), respectively. For Ala44(35-50)NCp7, intermolecular Zn(2+) binding predominantly occurs via the deprotonated Cys36 in the monodentate state with a rate constant of 3.6 x 10(7) M(-)(1) s(-)(1). In both mutants, the final state of the Zn(2+) complex is reached by subsequent stepwise ligand deprotonation and intramolecular substitution of coordinated water molecules. The rate constants for the intermolecular binding paths of the bidentate and tridentate states of Ala44(35-50)NCp7 and of the tridentate state of Ser49(35-50)NCp7 are much smaller than expected according to electrostatic considerations. This is attributed to conformational constraints required to achieve proper metal coordination during folding. The dissociation of Zn(2+) from both peptides is again characterized by a multistep process and takes place fastest via the protonated Zn(2+)-bound bidentate and monodentate states, with rate constants of approximately 0.3 and approximately 10(3) s(-)(1), respectively, for Ser49(35-50)NCp7 and approximately 4 x 10(-)(3) and approximately 500 s(-)(1), respectively, for Ala44(35-50)NCp7.  相似文献   

7.
Amayed P  Carlier MF  Pantaloni D 《Biochemistry》2000,39(40):12295-12302
Stathmin is an important protein that interacts with tubulin and regulates microtubule dynamics in a phosphorylation-controlled fashion. Here we show that the dissociation of guanosine 5'-diphosphate (GDP) from beta-tubulin is slowed 20-fold in the (tubulin)(2)-stathmin ternary complex (T(2)S). The kinetics of GDP or guanosine 5'-triphosphate (GTP) dissociation from tubulin have been monitored by the change in tryptophan fluorescence of tubulin upon exchanging 2-amino-6-mercapto-9-beta-ribofuranosylpurine 5'-diphosphate (S6-GDP) for tubulin-bound guanine nucleotide. At molar ratios of stathmin to tubulin lower than 0.5, biphasic kinetics were observed, indicating that the dynamics of the complex is extremely slow, consistent with its high stability. The method was used to characterize the effects of phosphorylation of stathmin on its interaction with tubulin. The serine-to-glutamate substitution of all four phosphorylatable serines of stathmin (4E-stathmin) weakens the stability of the T(2)S complex by about 2 orders of magnitude. The phosphorylation of serines 16 and 63 in stathmin has a more severe effect and weakens the stability of T(2)S 10(4)-fold. The rate of GDP dissociation is lowered only 7-fold and 4-fold in the complexes of tubulin with 4E-stathmin and diphosphostathmin, respectively. Sedimentation velocity studies support the conclusions of nucleotide exchange data and show that the T(2)S complexes formed between tubulin and 4E-stathmin or diphosphostathmin are less compact than the highly stable T(2)S complex. The correlation between the effect of phosphorylation of stathmin on the stability of T(2)S complex measured in vitro and on the function of stathmin in vivo is discussed.  相似文献   

8.
The kinetic mechanism of the binding and release of calcium by troponin and by the complexes troponin: tropomyosin, troponin:tropomyosin:actin, and troponin (TN)-tropomyosin (TM)-actin:myosin subfraction 1 (SF-1) was investigated using troponin labeled on the TN-I subunit with the fluorophore 4-(N-iodoac etoxyethyl-N-methyl)-7-nitrobenz-2-oxa-1,3-diazole. The apparent association constant is five to 10 times smaller for TN:TM:actin compared to TN:TM or TN and saturation of actin sites with SF-1 increased the binding constant approximately to the value for TN:TM. Kinetic measurements on TN or TN:TM fitted a single rate process for association or dissociation which is consistent with a model in which the calcium sites are equivalent and independent and each calcium induces a change in structure of the complex. TN:TM:actin gave biphasic transients for association and dissociation of calcium. The two binding sites are no longer equivalent and independent. The TN:TM:actin:SF-1 complex gave kinetic behavior essentially equivalent to TN:TM. The kinetics of calcium dissociation from the various complexes was also measured by the fluorescent calcium indicator quin 2, which gave the same values for the rate constants as for the labeled protein. The evidence is interpreted in terms of a model in which regulated actin can exist in two states and the binding of each calcium and SF-1 displaces the equilibrium between states. Formation of the complex of TN:TM with actin yielded an enhancement of the fluorescence of the labeled TN-I moiety of approximately 30%. The rate of constant for association of the complex decreased 6-fold in the presence of calcium while the rate constant for dissociation of the protein complex was essentially unchanged. Saturation of actin sites with SF-1 had no effect on the rate constant for association with TN:TM in the presence of calcium.  相似文献   

9.
Kinetics and mechanism in the reaction of gene regulatory proteins with DNA   总被引:28,自引:0,他引:28  
We have measured the kinetic properties of the Escherichia coli cAMP receptor protein (CAP) and lac repressor interacting with lac promoter restriction fragments. Under our reaction conditions (10 mM-Tris X HCl (pH 8.0 at 21 degrees C), 1 mM-EDTA, 10 microM-cAMP, 50 micrograms bovine serum albumin/ml, 5% glycerol), the association of CAP is at least a two-step process, with an initial, unstable complex formed with rate constant kappa a = 5(+/- 2.5) X 10(7) M-1 s-1. Subsequent formation of a stable complex occurs with an apparent bimolecular rate constant kappa a = 6.7 X 10(6) M-1 s-1. At low total DNA concentration, the dissociation rate constant for the specific CAP-DNA complex is 1.2 X 10(-4) s-1. The ratio of formation and dissociation rate constants yields an estimate of the equilibrium constant, Keq = 5 X 10(10) M-1, in good agreement with static results. We observed that the dissociation rate constant of both CAP-DNA and repressor-DNA complexes is increased by adding non-specific "catalytic" DNA to the reaction mixture. CAP dissociation by the concentration-dependent pathway is second-order in added non-specific DNA, consistent with either the simultaneous or the sequential participation of two DNA molecules in the reaction mechanism. The results imply a role for distal DNA in assembly-disassembly of specific CAP-DNA complexes, and are consistent with a model in which the subunits in the CAP dimer separate in the assembly-disassembly process. The dissociation of lac repressor-operator complexes was found to be DNA concentration-dependent as well, although in contrast to CAP, the reaction is first-order in catalytic DNA. Added excess operator-rich DNA gave more rapid dissociation than equivalent concentrations of non-specific DNA, indicating that the sequence content of the competing DNA influences the rate of repressor dissociation. The simplest interpretation of these observations is that lac repressor can be transferred directly from one DNA molecule to another. A comparison of the translocation rates calculated for direct transfer with those predicted by the one-dimensional sliding model indicates that direct transfer may play a role in the binding site search of lac repressor.  相似文献   

10.
The binding of oligomeric peptide-MHC (pMHC) complexes to cell surface TCR can be considered to approximate TCR-pMHC interactions at cell-cell interfaces. In this study, we analyzed the equilibrium binding of streptavidin-based pMHC oligomers (tetramers) and their dissociation kinetics from CD8(pos) T cells from 2C-TCR transgenic mice and from T cell hybridomas that expressed the 2C TCR or a high-affinity mutant (m33) of this TCR. Our results show that the tetramers did not come close to saturating cell-surface TCR (binding only 10-30% of cell-surface receptors), as is generally assumed in deriving affinity values (K(D)), in part because of dissociative losses from tetramer-stained cells. Guided by a kinetic model, the oligomer dissociation rate and equilibrium constants were seen to depend not only on monovalent association and dissociation rates (k(off) and k(on)), but also on a multivalent association rate (μ) and TCR cell-surface density. Our results suggest that dissociation rates could account for the recently described surprisingly high frequency of tetramer-negative, functionally competent T cells in some T cell responses.  相似文献   

11.
Simulation of the diffusional association of barnase and barstar.   总被引:2,自引:1,他引:1       下载免费PDF全文
The rate of protein association places an upper limit on the response time due to protein interactions, which, under certain circumstances, can be diffusion-controlled. Simulations of model proteins show that diffusion-limited association rates are approximately 10(6)-10(7) M-1 s-1 in the absence of long-range forces (Northrup, S. H., and H. P. Erickson. 1992. Kinetics of protein-protein association explained by Brownian dynamics computer simulations. Proc. Natl. Acad. Sci. U.S.A. 89:3338-3342). The measured association rates of barnase and barstar are 10(8)-10(9) M-1 s-1 at 50 mM ionic strength, and depend on ionic strength (Schreiber, G., and A. R. Fersht. 1996. Rapid, electrostatically assisted association of proteins. Nat. Struct. Biol. 3:427-431), implying that their association is electrostatically facilitated. We report Brownian dynamics simulations of the diffusional association of barnase and barstar to compute association rates and their dependence on ionic strength and protein mutation. Crucial to the ability to reproduce experimental rates is the definition of encounter complex formation at the endpoint of diffusional motion. Simple definitions, such as a required root mean square (RMS) distance to the fully bound position, fail to explain the large influence of some mutations on association rates. Good agreement with experiments could be obtained if satisfaction of two intermolecular residue contacts was required for encounter complex formation. In the encounter complexes, barstar tends to be shifted from its position in the bound complex toward the guanine-binding loop on barnase.  相似文献   

12.
T M Lohman 《Biochemistry》1984,23(20):4656-4665
The dissociation kinetics of cooperatively bound bacteriophage T4 gene 32 protein from a variety of single-stranded homopolynucleotides has been investigated by stopped-flow techniques. Irreversible dissociation of the complexes was induced by rapidly increasing the salt concentration and monitoring the increase in tryptophan fluorescence upon dissociation of the gene 32 protein. The dependence of the apparent dissociation rate constant on initial fractional saturation of the nucleic acid lattice as well as the observation of zero-order kinetics when the lattice is initially fully saturated with protein indicates that dissociation occurs only from the ends of protein clusters and not from doubly contiguous molecules. The data for the entire time course are quantitatively fit by a kinetics model specifying irreversible dissociation of only singly contiguously bound protein [Lohman, T.M. (1983) Biopolymers 22, 1697-1713]. This model is used to extract molecular rate constants for the dissociation of isolated, singly contiguously and doubly contiguously bound protein. It is also shown that the polynucleotide specificity observed for the cooperative binding constant, K omega, and the cooperativity itself are intrinsic properties of the dissociation rate of the various complexes.  相似文献   

13.
A novel analytical method based on the exact solution of equations of kinetics of unbranched first- and pseudofirst-order mechanisms is developed for application to the process of Esigma70 RNA polymerase (R)-lambdaPR promoter (P) open complex formation, which is described by the minimal three-step mechanism with two kinetically significant intermediates (I1, I2), [equation: see text], where the final product is an open complex RPo. The kinetics of reversible and irreversible association (pseudofirst order, [R] > [P]) to form long-lived complexes (RPo and I2) and the kinetics of dissociation of long-lived complexes both exhibit single exponential behavior. In this situation, the analytical method provides explicit expressions relating observed rate constants to the microscopic rate constants of mechanism steps without use of rapid equilibrium or steady-state approximations, and thereby provides a basis for interpreting the composite rate constants of association (ka), isomerization (ki), and dissociation (kd) obtained from experiment for this or any other sequential mechanism of any number of steps. In subsequent papers, we apply this formalism to analyze kinetic data obtained in the reversible and irreversible binding regimes of Esigma70 RNA polymerase (R)-lambdaP(R) promoter (P) open complex formation.  相似文献   

14.
The kinetics of charge recombination following photoexcitation by a laser pulse have been analyzed in the reaction center-light harvesting complex 1 (RC-LH1) purified from the photosynthetic bacterium Rhodobacter sphaeroides. In RC-LH1 core complexes isolated from photosynthetically grown cells P(+)Q(B)(-) recombines with an average rate constant, k approximately 0.3 s(-1), more than three times smaller than that measured in RC deprived of the LH1 (k approximately 1 s(-1)). A comparable, slowed recombination kinetics is observed in RC-LH1 complexes purified from a pufX-deleted strain. Slowing of the charge recombination kinetics is even more pronounced in RC-LH1 complexes isolated from wild-type semiaerobically grown cells (k approximately 0.2 s(-1)). Since the kinetics of P(+)Q(A)(-) recombination is unaffected by the presence of the antenna, the P(+)Q(B)(-) state appears to be energetically stabilized in core complexes. Determinations of the ubiquinone-10 (UQ(10)) complement associated with the purified RC-LH1 complexes always yield UQ(10)/RC ratios larger than 10. These quinone molecules are functionally coupled to the RC-LH1 complex, as judged from the extent of exogenous cytochrome c(2) rapidly oxidized under continuous light excitation. Analysis of P(+)Q(B)(-) recombination, based on a kinetic model which considers fast quinone equilibrium at the Q(B) binding site, indicates that the slowing down of charge recombination kinetics observed in RC-LH1 complexes cannot be explained solely by a quinone concentration effect and suggests that stabilization of the light-induced charge separation is predominantly due to interaction of the Q(B) site with the LH1 complex. The high UQ(10) complements detected in RC-LH1 core complexes, but not in purified light-harvesting complex 2 and in RC, are proposed to reflect an in vivo heterogeneity in the distribution of the quinone pool within the chromatophore bilayer.  相似文献   

15.
16.
Lasey RC  Liu L  Zang L  Ogawa MY 《Biochemistry》2003,42(13):3904-3910
Photoinduced electron-transfer (ET) occurs between a negatively charged metallopeptide, [Ru(bpy)(2)(phen-am)-Cys-(Glu)(5)-Gly](3-) = RuCE(5)G, and ferricytochrome c = Cyt c. In the presence of Cyt c, the triplet state lifetime of the ruthenium metallopeptide is shortened, and the emission decays via biexponential kinetics, which indicates the existence of two excited-state populations of ruthenium peptides. The faster decay component displays concentration-independent kinetics demonstrating the presence of a preformed peptide-protein complex that undergoes intra-complex electron-transfer. Values of K(b) = (3.5 +/- 0.2) x 10(4) M(-1) and k(obs)(ET)= (2.7 +/- 0.4) x 10(6) s(-1) were observed at ambient temperatures. The magnitude of k(obs)(ET) decreases with increasing solvent viscosity, and the behavior can be fit to the expression k(obs)(ET) proportional to eta(-alpha) to give alpha = 0.59 +/- 0.05. The electron-transfer process occurring in the preformed complex is therefore gated by a rate-limiting configurational change of the complex. The slower decay component displays concentration-dependent kinetics that saturate at high concentrations of Cyt c. Analysis according to rapid equilibrium formation of an encounter complex that undergoes unimolecular electron-transfer yields K(b)' = (2.5 +/- 0.7) x 10(4) M(-1) and k(obs')(ET)= (7 +/- 3) x 10(5) s(-1). The different values of k(obs)(ET) and k(obs')(ET) suggest that the peptide lies farther from the heme when in the encounter complex. The value of k(obs')(ET) is viscosity dependent indicating that the reaction occurring within the encounter complex is also configurationally gated. A value of alpha = 0.98 +/- 0.14 is observed for k(obs')(ET), which suggests that the rate-limiting gating processes in the encounter complex is different from that in the preformed complex.  相似文献   

17.
A Aulabaugh  J V Schloss 《Biochemistry》1990,29(11):2824-2830
N-Hydroxy-N-isopropyloxamate (IpOHA) is an exceptionally potent inhibitor of the Escherichia coli ketol-acid reductoisomerase. In the presence of Mg2+ or Mn2+, IpOHA inhibits the enzyme in a time-dependent manner, forming a nearly irreversible complex. Nucleotide, which is essential for catalysis, greatly enhances the binding of IpOHA by the reductoisomerase, with NADPH (normally present during the enzyme's rearrangement step, i.e., conversion of a beta-keto acid into an alpha-keto acid, in either the forward or reverse physiological reactions) being more effective than NADP. In the presence of Mg2+ and NADPH, IpOHA appears to bind to the enzyme in a two-step mechanism, with an initial inhibition constant of 160 nM and a maximum rate of formation of the tight, slowly reversible complex of 0.57 min-1 (values that give an association rate of IpOHA, at low concentration, of 5.9 X 10(4) M-1 s-1). The rate of exchange of [14C]IpOHA from an enzyme-[14C]IpOHA-Mg2(+)-NADPH complex with exogenous, unlabeled IpOHA has a half-time of 6 days (150 h). This dissociation rate (1.3 X 10(-6) s-1) and the association rate determined by inactivation kinetics define an overall dissociation constant of 22 pM. By contrast, in the presence of Mn2+ and NADPH, the corresponding association and dissociation rates for IpOHA are 8.2 X 10(4) M-1 s-1 and 3.2 X 10(-6) s-1 (half-time = 2.5 days), respectively, which define an overall dissociation constant of 38 pM. In the presence of NADP or in the absence of nucleotide (both in the presence of Mg2+), the enzyme-IpOHA complex is far more labile, with dissociation half-times of 28 and 2 h, respectively. In the absence of Mg2+ or Mn2+, IpOHA does not exhibit time-dependent inhibition of the reductoisomerase.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
We explored the use of a newly developed cuvette-based surface plasmon resonance (SPR) instrument (IBIS) to study peptide-protein interactions. We studied the interaction between the SH2 domain of lck and a phosphotyrosine peptide EPQY*EEIPIYL which was immobilized on a sensor chip. No indications for mass transport limitation (MTL) were observed when standard kinetic approaches were used. However, addition of competing peptide during dissociation revealed a high extent of rebinding. A dissociation rate constant (k(d)) of 0.6+/-0.1 s(-1) was obtained in the presence of large amounts of peptide. A simple bimolecular binding model, applying second-order kinetics for the cuvette system, could not adequately describe the data. Fits were improved upon including a step in the model which describes diffusion of the SH2 domain from the bulk to the sensor, especially for a surface with high binding capacity. From experiments in glycerol-containing buffers, it appeared that the diffusion rate decreased with higher viscosity. It is demonstrated that MTL during association and dissociation can be described by the same diffusion rate. A binding constant (K(D)) of 5.9+/-0.8 nM was obtained from the SPR equilibrium signals by fitting to a Langmuir binding isotherm, with correction for loss of free analyte due to binding. An association rate constant k(a) of 1.1(+/-0.2)x10(8) M(-1) x s(-1) was obtained from k(d)/K(D). The values for k(a) and k(d) obtained in this way were 2-3 orders larger than that from standard kinetic analysis, ignoring MTL. We conclude that in a cuvette the extent of MTL is comparable to that in a flow system.  相似文献   

19.
Reactivation of tetrameric porcine skeletal muscle lactic dehydrogenase after dissociation and extensive unfolding of the monomers by 6 M guanidine hydrochloride (Gdn . HCl) is characterized by sigmoidal kinetics, indicating a complex mechanism involving rate-limiting folding and association steps. For analysis of the association reactions, chemical cross-linking with glutaraldehyde may be used [Hermann, R., Jaenicke, R., & Rudolph, R. (1981) Biochemistry 20, 2195-2201]. The data clearly show that the formation of a dimeric intermediate is determined by a first-order folding reaction of the monomers with k1 = (8.0 +/- 0.1) x 10(-4) s-1. The rate constant of the association of dimers to tetramers which represents the second rate-limiting step on the pathway of reconstitution after guanidine denaturation, was then determined by reactivation and cross-linking experiments after dissociation in 0.1 M H3PO4 containing 1 M Na2SO4. The rate constant for the dimer association (which is the only rate-limiting step after acid dissociation) was k2 = (3.0 +/- 0.5) x 10(4) M-1 s-1. On the basis of the given two rate constants, the complete reassociation pattern of porcine lactic dehydrogenase after dissociation and denaturation in 6 M Gdn . HCl can be described by the kinetic model (formula: see text).  相似文献   

20.
Protein-protein recognition results from the assembly of complementary surfaces on two molecules that form a stable, noncovalent, specific complex. Our interest was to describe kinetic aspects of the recognition in order to understand the subtle molecular mechanism of association. R67 dihydrofolate reductase (DHFR) provides an ideal model to investigate kinetic parameters of protein-protein association since it is a homotetramer resulting from the pH-dependent dimerization of homodimers. We took advantage of the presence of a tryptophan residue at the dimer-dimer interface to monitor pH-dependent oligomerization of R67 DHFR using stopped-flow fluorescence techniques. Except for pH near neutrality where dissociation exhibited biphasic kinetics, association and dissociation followed monophasic kinetics fitted on a two-state model. Apparent rate constants of association k(on) and dissociation k(off) were determined at various pHs and pointed to the key role of a histidine located at the dimer-dimer interface in the pH control of tetramerization. The values of the tetramer-dimer equilibrium dissociation constant were calculated from the ratio k(off) /k(on) and correlated well with those previously measured at equilibrium. The thermodynamic parameters and the activation energies of both the association and dissociation were determined and indicated that the association is enthalpy driven and suggested that the formation of four hydrogen bonds (one per monomer) is responsible for the thermodynamic stability of the tetramer. Detailed analysis of the biphasic kinetics led to an original model, in which protonation of the tetramer is the triggering event for the dissociation process while the association involves primarily the unprotonated dimers.  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