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.
Enzymatic digestion with a type IIP restriction endonuclease EcoRV was investigated on a DNA-immobilized 27-MHz quartz crystal microbalance (QCM). Real-time observations of both the enzyme binding process and the DNA cleavage process of EcoRV were followed by frequency (mass) changes on the QCM, which were dependent on divalent cations such as Ca(2+) or Mg(2+). In the presence of Ca(2+), the site-specific binding of EcoRV to DNA could be observed, without the catalytic process. On the other hand, in the presence of Mg(2+), both the binding of the enzyme to the specific DNA (mass increase) and the site-specific cleavage reaction (mass decrease) could be observed continuously from QCM frequency changes. From time courses of frequency (mass) changes, each kinetic parameter, namely binding rate constants (k(on)), dissociation rate constants (k(off)), dissociation constants (K(d)) of EcoRV to DNA, and catalytic rate constant (k(cat)) of the cleavage reaction, could be determined. The binding kinetic parameters of EcoRV in the presence of Ca(2+) were consistent with those of the binding process followed by the cleavage process in the presence of Mg(2+). The k(cat) value obtained by the QCM method was also consistent with that obtained by other methods. This study is the first to simultaneously determine k(on), k(off), and k(cat) for a type IIP restriction endonuclease on one device.  相似文献   

2.
The equilibrium binding and association kinetics of the fos-jun dimer (basic and leucine zipper domain) to the AP-1 DNA were studied using a quantitative assay. The basic-region and leucine zipper (bZip) domain of c-fos was expressed as a fusion protein with glutathione S-transferase, and it was bound to glutathione-agarose. The GST-fused fos bZip region was allowed to form a heterodimer with the bZip domain of c-jun, to which radiolabeled AP-1 nucleotides were added. After thorough washing, the gel-bound radioactivity was counted. The binding and dissociation rate constants (k(1) and k-(1)) of the fos-jun dimer and DNA could be obtained from a time-course experiment. The association binding constant (K(1)) was determined using both a thermodynamic equation and kinetic parameters. Nordihydroguaiaretic acid (NDGA), momordin I, natural product inhibitors of the fos-jun/DNA complex formation, was applied to this jun-GST-fused fos system and it was found to decrease the apparent equilibrium binding of dimer and DNA. The thermodynamic constant of dimer and inhibitor binding was also determined.  相似文献   

3.
4.
The interaction of the human adenovirus proteinase (AVP) and AVP-DNA complexes with the 11-amino acid cofactor pVIc was characterized. The equilibrium dissociation constant for the binding of pVIc to AVP was 4.4 microM. The binding of AVP to 12-mer single-stranded DNA decreased the K(d) for the binding of pVIc to AVP to 0.09 microM. The pVIc-AVP complex hydrolyzed the substrate with a Michaelis constant (K(m)) of 3.7 microM and a catalytic rate constant (k(cat)) of 1.1 s(-1). In the presence of DNA, the K(m) increased less than 2-fold, and the k(cat) increased 3-fold. Alanine-scanning mutagenesis was performed to determine the contribution of individual pVIc side chains in the binding and stimulation of AVP. Two amino acid residues, Gly1' and Phe11', were the major determinants in the binding of pVIc to AVP, while Val2' and Phe11' were the major determinants in stimulating enzyme activity. Binding of AVP to DNA greatly suppressed the effects of the alanine substitutions on the binding of mutant pVIcs to AVP. Binding of either or both of the cofactors, pVIc or the viral DNA, to AVP did not dramatically alter its secondary structure as determined by vacuum ultraviolet circular dichroism. pVIc, when added to Hep-2 cells infected with adenovirus serotype 5, inhibited the synthesis of infectious virus, presumably by prematurely activating the proteinase so that it cleaved virion precursor proteins before virion assembly, thereby aborting the infection.  相似文献   

5.
The kinetics of binding of bovine trypsin to a proteinaceous inhibitor of trypsin from buckwheat seeds (BWI-1a) has been studied. The association rate constant (k(ass)) was 2.2 x 10(6) M-1 x sec-1 and the dissociation rate constant (k(off)) of the enzyme--inhibitor complex was 3.5 x 10(-3) sec-1; the inhibition constant Ki was 1.5 nM. The inhibitor BWI-1a is of the slow, tightly binding type. The mechanism of the inhibition of bovine trypsin by the trypsin inhibitor BWI-1a was studied. The mechanism of inhibition was found to involve two steps according to the kinetic data.  相似文献   

6.
The kinetics of interaction of Esigma(70) RNA polymerase (R) with the lambdaP(R) promoter (P) were investigated by filter binding over a broad range of temperatures (7.3-42 degrees C) and concentrations of RNA polymerase (1-123 nM) in large excess over promoter DNA. Under all conditions examined, the kinetics of formation of competitor-resistant complexes (I(2), RP(o)) are single-exponential with first order rate constant beta(CR). Interpretation of the polymerase concentration dependence of beta(CR) in terms of the three step mechanism of open complex formation yields the equilibrium constant K(1) for formation of the first kinetically significant intermediate (I(1)) and the forward rate constant (k(2)) for the conformational change converting I(1) to the second kinetically significant intermediate I(2): R + P-->(K(1))<--I(1)(k(2))-->I(2). Use of rapid quench mixing allows K(1) and k(2) to be individually determined over the entire temperature range investigated, previously not possible at this promoter using manual mixing. Given the large (>60 bp) interface formed in I(1), its relatively small binding constant K(1) at 37 degrees C at this [salt] (approximately 6 x 10(6) M(-1)) strongly argues that binding free energy is used to drive large-scale structural changes in polymerase and/or promoter DNA or other coupled processes. Evidence for coupling of protein folding is provided by the large and negative activation heat capacity of k(a)[DeltaC(o,++)(a)= -1.5(+/-0.2)kcal K(-1)], now shown to originate directly from formation of I(1) [DeltaC(o)(1)= -1.4(+/-0.3)kcal K(-1)] rather than from the formation of I(2) as previously proposed. The isomerization I(1)-->I(2) exhibits relatively slow kinetics and has a very large temperature-independent Arrhenius activation energy [E(act)(2)= 34(+/-2)kcal]. This kinetic signature suggests that formation of the transition state (I(1)-I(2)++ involves large conformational changes dominated by changes in the exposure of polar and/or charged surface to water. Structural and biochemical data lead to the following hypotheses to interpret these results. We propose that formation of I(1) involves coupled folding of unstructured regions of polymerase (beta, beta' and sigma(70)) and bending of promoter DNA (in the -10 region). We propose that interactions with region 2 of sigma(70) and possibly domain 1 of beta induce a kink at the -11/-12 base pairs of the lambdaP(R) promoter which places the downstream DNA (-5 to +20) in the jaws of the beta and beta' subunits of polymerase in I(1). These early interactions of beta and beta' with the DNA downstream of position -5 trigger jaw closing (with coupled folding) and subsequent steps of DNA opening.  相似文献   

7.
PDC-109, the major protein of bovine seminal plasma, binds to sperm plasma membranes upon ejaculation and plays a crucial role in the subsequent events leading to fertilization. The binding process is mediated primarily by the specific interaction of PDC-109 with choline-containing phospholipids. In the present study the kinetics and mechanism of the interaction of PDC-109 with phospholipid membranes were investigated by the surface plasmon resonance technique. Binding of PDC-109 to different phospholipid membranes containing 20% cholesterol (wt/wt) indicated that binding occurs by a single-step mechanism. The association rate constant (k(1)) for the binding of PDC-109 to dimyristoylphosphatidylcholine (DMPC) membranes containing cholesterol was estimated to be 5.7 x 10(5) M(-1) s(-1) at 20 degrees C, while the values of k(1) estimated at the same temperature for the binding to membranes of negatively charged phospholipids such as dimyristoylphosphatidylglycerol (DMPG) and dimyristoylphosphatidic acid (DMPA) containing 20% cholesterol (wt/wt) were at least three orders of magnitude lower. The dissociation rate constant (k(-1)) for the DMPC/PDC-109 system was found to be 2.7 x 10(-2) s(-1) whereas the k(-1) values obtained with DMPG and DMPA was about three to four times higher. From the kinetic data, the association constant for the binding of PDC-109 to DMPC was estimated as 2.1 x 10(7) M(-1). The association constants for different phospholipids investigated decrease in the order: DMPC > DMPG > DMPA > DMPE. Thus the higher affinity of PDC-109 for choline phospholipids is reflected in a faster association rate constant and a slower dissociation rate constant for DMPC as compared to the other phospholipids. Binding of PDC-109 to dimyristoylphosphatidylethanolamine and dipalmitoylphosphatidylethanolamine, which are also zwitterionic, was found to be very weak, clearly indicating that the charge on the lipid headgroup is not the determining factor for the binding. Analysis of the activation parameters indicates that the interaction of PDC-109 with DMPC membranes is favored by a strong entropic contribution, whereas negative entropic contribution is primarily responsible for the rather weak interaction of this protein with DMPA and DMPG.  相似文献   

8.
During V(D)J recombination, RAG targeting to correct sites versus off-target sites relies on both DNA sequence features and on chromatin marks. Kinetic analysis using the first highly active full-length purified RAG1/RAG2 complexes has now allowed us to define the important catalytic features of this complex. We found that the overall rate of nicking, but not hairpinning, is critical for the discrimination between correct (optimal) versus off-target (suboptimal) sites used in human T-cell lymphomas, and we show that the C-terminal portion of RAG2 is required for this. This type of kinetic analysis permits us to analyze only the catalytically active RAG complex, in contrast to all other methods, which are unavoidably confounded by mixture with inactive RAG complexes. Moreover, we can distinguish the two major features of any enzymatic catalysis: the binding constant (K(D)) and the catalytic turnover rate, k(cat). Beyond a minimal essential threshold of heptamer quality, further suboptimal heptamer deviations primarily reduce the catalytic rate constant k(cat) for nicking. Suboptimal nonamers reduce not only the binding of the RAG complex to the recombination site (K(D)) but also the catalytic rate constant, consistent with a tight interaction between the RAG complex and substrate during catalysis. These features explain many aspects of RAG physiology and pathophysiology.  相似文献   

9.
One physiological significance of the red blood cell (RBC) structure is that NO binding of Hb is retarded by encapsulation with the cell membrane. To clarify the mechanism, we analyzed Hb-vesicles (HbVs) with different intracellular Hb concentrations, [Hb](in), and different particle sizes using stopped-flow spectrophotometry. The apparent NO binding rate constant, k(on)('(NO)), of HbV at [Hb](in) = 1 g/dl was 2.6 x 10(7) m(-1) s(-1), which was almost equal to k(on)((NO)) of molecular Hb, indicating that the lipid membrane presents no obstacle for NO binding. With increasing [Hb](in) to 35 g/dl, k(on)('(NO)) decreased to 0.9 x 10(7) m(-1) s(-1), which was further decreased to 0.5 x 10(7) m(-1) s(-1) with enlarging particle diameter from 265 to 452 nm. For CO binding, which is intrinsically much slower than NO binding, k(on)('(CO)) did not change greatly with [Hb](in) and the particle diameter. Results obtained using diffusion simulations coupled with elementary binding reactions concur with these tendencies and clarify that NO is trapped rapidly by Hb from the interior surface region to the core of HbV at a high [Hb](in), retarding NO diffusion toward the core of HbV. In contrast, slow CO binding allows time for further CO-diffusion to the core. Simulations extrapolated to larger particles (8 mum) showing retardation even for CO binding. The obtained k(on)('(NO)) and k(on)('(NO)) yield values similar to those reported for RBCs. In summary, the intracellular, not extracellular, diffusion barrier is predominant due to the rapid NO binding that induces a rapid sink of NO from the interior surface to the core, retarding further NO diffusion and binding.  相似文献   

10.
Conlan LH  Dupureur CM 《Biochemistry》2002,41(50):14848-14855
Restriction enzymes serve as important model systems for understanding the role of metal ions in phosphodiester hydrolysis. To this end, a number of laboratories have reported dramatic differences between the metal ion-dependent and metal ion-independent DNA binding behaviors of these systems. In an effort to illuminate the underlying mechanistic details which give rise to these differences, we have quantitatively dissected these equilibrium behaviors into component association and dissociation rates for the representative PvuII endonuclease and use these data to assess the stoichiometry of metal ion involvement in the binding process. The dependence of PvuII cognate DNA on Ca(II) concentration binding appears to be cooperative, exhibiting half-saturation at 0.6 mM metal ion and yielding an n(H) of 3.5 +/- 0.2 per enzyme homodimer. Using both nitrocellulose filter binding and fluorescence assays, we observe that the cognate DNA dissociation rate (k(-)(1) or k(off)) is very slow (10(-)(3) s(-)(1)) and exhibits a shallow dependence on metal ion concentration. DNA trap cleavage experiments with Mg(II) confirm the general irreversibility of DNA binding relative to cleavage, even at low metal ion concentrations. More dramatically, the association rate (k(1) or k(on)) also appears to be cooperative, increasing more than 100-fold between 0.2 and 10 mM Ca(II), with an optimum value of 2.7 x 10(7) M(-)(1) s (-)(1). Hill analysis of the metal ion dependence of k(on) indicates an n(H) of 3.6 +/- 0.2 per enzyme dimer. This value is consistent with the involvement in DNA association of two metal ions per subunit active site, a result which lends new strength to arguments for two-metal ion mechanisms in restriction enzymes.  相似文献   

11.
Asgeirsson B  Cekan P 《FEBS letters》2006,580(19):4639-4644
Temperature imposes limits on where life can thrive and this is evident in the evolution of the basic structural properties of proteins. Cold-adaptation of enzymes is one example, where the catalytic rate constant (k(cat)) is increased compared with hot-acclimated homologous under identical assay conditions. Trypsin I from Atlantic cod (Gadus morhua) has catalytic efficiency (k(cat)/K(m)) for amide hydrolysis that is 17-fold larger than observed for bovine trypsin. Here, the individual rate-constants for association of substrate (k(1)), dissociation of substrate (k(-1)), and acylation of the enzyme (k(2)) have been determined using benzoyl-Arg-p-nitroanilide or benzyloxycarbonyl-Gly-Pro-Arg-p-nitroanilide as substrates. Rather unexpectedly, by far the largest difference (37-fold increase) was observed in k(1), the rate constant for binding of substrate. The cold-adaptation of the dissociation and catalytic steps were not as prominent (increased by 3.7-fold). The length of substrate did have an effect by increasing the reaction rate by 70-fold, and again, the step most affected was the initial binding-step.  相似文献   

12.
Inhibition of the pre-steady-state burst of nucleotide incorporation by a single incorrect nucleotide (nucleotide discrimination) was measured with the Klenow fragment of DNA polymerase I [KF(exo+)]. For the eight mispairs studied on three DNA sequences, only low levels of discrimination ranging from none to 23-fold were found. The kinetics of dNTP incorporation into the 9/20-mer at low nucleotide concentrations was also determined. A limit of greater than or equal to 250 s-1 was placed on the nucleotide off-rate from the KF(exo+)-9/20-dTTP complex in accord with nucleotide binding being at equilibrium in the overall kinetic sequence. The influence of the relatively short length of the 9/20-mer on the mechanism of DNA replication fidelity was determined by remeasuring important kinetic parameters on a 30/M13-mer with high homology to the 9/20-mer. Pre-steady-state data on the nucleotide turnover rates, the dATP(alpha S) elemental effect, and the burst of dAMP misincorporation into the 30/M13-mer demonstrated that the kinetics were not affected by the length of the DNA primer/template. The effects on fidelity of two site-specific mutations, KF(polA5) and KF(exo-), were also examined. KF(polA5) showed an increased rate of DNA dissociation and a decreased rate of polymerization resulting in less processive DNA synthesis. Nevertheless, with at least one misincorporation event, that of dAMP into the 9/20-mer, KF(polA5) displays an increased replication fidelity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Double-stranded (ds) oligodeoxynucleotides (29mers) containing an O6-ethylguanine (O6-EtGua) flanked 5' and 3' by different bases (5'..TGT..3'; 5'..CGG..3', 5'..GGT..3'; 5'..GGG..3'; 5'..GGA..3') were synthesized to investigate the binding and repair characteristics of recombinant human O6-alkylguanine-DNA alkyltransferase (AT) in vitro. The apparent association constant (KA(app)) of AT to the oligomers and the repair rate constant for O6-EtGua (k) respectively, were determined by gel retardation and a monoclonal antibody-based filter binding assay. When ds- or single-stranded (ss) oligomers with or without O6-EtGua were used, no major differences in KA(app) values were observed with either substrate: KA(app) values for native AT were 7.1 and 8.4 x 10(5) M(-1) respectively, for unmodified and [O6-EtGua]-containing ds-oligomers. The corresponding values for ss-oligomers were 1.0 and 4.9 x 10(5) M(-1). The N-terminal first 56 amino acids of AT only exert a limited influence on DNA binding; the KA(app) values for an N-terminally truncated AT protein (1.1 x 10(5) M(-1)) and native AT were of the same order. Moreover, KA(app) was hardly affected by Cys(145)-methylated AT (2.0 x 10(5) M(-1)). The k-values (6.5-11.5 x 10(6) M(-1)s(-1)) were not significantly dependent on nucleotide sequence. k-values of 5.3 and 4.0 x 10(6) M(-1)s(-1) respectively, were obtained with the N-terminally truncated AT protein and for repair of the postreplicative mispair [O6-EtGua]: T by native AT. The low KA(app), the negligible influence on O6 of ethylation, and the minor modulation KA(app) and k by varying the bases flanking O6-EtGua, all indicate that the binding of AT to DNA is non-specific and mediated mainly by ionic interactions [reduced KA(app) and k-values at increased ionic strength]. Surplus DNA reduces the rate of O6-EtGua repair in ds-oligomers by competitive binding of AT molecules. The reaction mechanism of AT with DNA in vivo requires further investigation.  相似文献   

14.
The chemical kinetic mechanism of kinesin (K) is considered by using a consensus scheme incorporating biochemically defined open, closed and trapped states. In the absence of microtubules, the dominant species is a trapped K*ADP state, which is defined by its ultra-slow release of ADP (off rate, k(off) approximately 0.002 s(-1)) and weak microtubule binding (dissociation constant, K(d) approximately 10-20 microM). Once bound, this trapped state equilibrates with a strongly binding open state that rapidly releases ADP (k(off) approximately 300 s(-1)). After ADP release, Mg*ATP binds (on rate, k(on) approximately 2 microM(-1)s(-1)) driving formation of a closed state that is defined by hydrolysis competence and by strong binding to microtubules. Hydrolysis (k(hyd) approximately 100-300 s(-1)) and phosphate release (k(off)>100 s(-1)) both occur in this microtubule-bound closed state. Phosphate release acts as a gate that controls reversion to the trapped K*ADP state, which detaches from the microtubule, completing the cycle.  相似文献   

15.
1. The results of kinetic studies in vitro are difficult to apply to metabolic reactions in vivo. 2. In living vertebrates reaction rates are usually first-order and for a particular reaction the rate "constant", k, varies with the several thousand-fold variations in metabolic rate. 3. Therefore, in the kinetic equation (formula: see text) since K varies with metabolic rate, Vmax and/or Km will aslo vary. 4. However, reaction rates for a series of different substrates were similar in animals varying widely in metabolic rate if corrections were made for differences in blood flow. 5. The observation that metabolic rate (reaction rate) is directly dependent on blood flow (Coulson et al., 1977) allowed derivation of new kinetic constants which were valid in vivo. 6. Introduction of a flow term into the observed first-order equations yields an "affinity constant", K, or a "flow constant", KF, somewhat analogous to Km, which allows one to predict reaction rates in animals of different metabolic rates. 7. The defining equations are (formula: see text) where V = reaction velocity in mmol/hr/kg tissue, [S] = millimolar substrate concentration in the blood, k = In 2/tau (tau the half-life) and F = blood flow in 1/hr/kg tissue. 8. Thus, K = k/F = 1/KF. The KF's for the initial step in the degradation of 17 amino acids in rats, dogs, lizards, turtles and alligators were similar, demonstrating the similarity of enzyme affinities in different species.  相似文献   

16.
Yessotoxin (YTX) is a generic name for a group of lipophilic compounds recently discovered and chemically characterized. Association measurements were done in a resonant mirror biosensor. The instrument detects changes in the refractive index and/or thickness occurring within a few hundred nanometers form the sensor surface where a molecule is attached. We used aminosilane surfaces where phosphodiesterase 3',5'-cyclic-nucleotide-specific from bovine brain (PDEs) was immobilized. Over this immobilized ligand different amounts of YTX were added and typical association curve profiles were observed. These association curves fit a pseudo-first-order kinetic equation where the apparent association rate constant (k(on)) can be calculated. The value of this constant increases with YTX concentration. From the representation of k(on) versus YTX concentration we obtained the association rate constant (k(ass)) 248+/-40 M(-1)s(-1) and the dissociation rate constant (k(diss)) 9.36 x 10(-4)+/-1.72 x 10(-4)s(-1). From these values the kinetic equilibrium dissociation constant (K(D)) for YTX-PDEs association can be calculated. The value of this last constant is 3.74 x 10(-6)+/-8.25 x 10(-8)M YTX. The PDE-YTX association was used as a method suitable for determination of the toxin concentration in a shellfish sample. The assay had sufficient sensitivity and can be used on simple shellfish extracts.  相似文献   

17.
The physical and chemical reaction pathways of the metabolite model compound benzo(a)pyrene-7,8-diol-9,10-epoxide (BPDE) in aqueous (double-stranded) DNA solutions was investigated as a function of temperature (0-30 degrees C), pH (7.0-9.5), sodium chloride concentration (0-1.5M) and DNA concentration in order to clarify the relationships between the multiple reaction mechanisms of this diol epoxide in the presence of nucleic acids. The reaction pathways are (1) noncovalent intercalative complex formation with DNA, characterized by the equilibrium constant K, and Xb the fraction of molecules physically bound; (2) accelerated hydrolysis of BPDE bound to DNA; (3) covalent binding to DNA; and (4) hydrolysis of free BPDE(kh). The DNA-induced hydrolysis of BPDE to tetraols and the covalent binding to DNA are parallel pseudo-first-order reactions. Following the rapid (millisecond time scale) noncovalent complex formation between BPDE and DNA, a much slower (approximately minutes) H+-dependent (either specific or general acid catalysis) formation of a DNA-bound triol carbonium ion (rate constant k3) occurs. At pH 7.0 the activation energy of k3 is 8.7 +/- 0.9 kcal/mol, which is lower than the activation energy of hydrolysis of free BPDE in buffer solution (14.2 +/- 0.7 kcal/mol), and which thus partially accounts for the acceleration of hydrolysis of BPDE upon complexation with DNA. The formation of the triol carbonium ion is followed by a rapid reaction with either water to form tetraols (rate constant kT), or covalent binding to DNA (kc). The fraction of BPDE molecules which undergo covalent binding is fcov approximately equal to kc/(kc + kT) = 0.10 and is independent of the overall BPDE reaction rate constant k = kh(1 - Xb) + k3Xb if Xb----1.0, or is independent of Xb as long as k3Xb much greater than kh(1 - Xb). Thus, at Xb = 0.9, fcov is independent of pH (7.0-9.5) even though k exhibits a 70-fold variation in this pH range and k----kh above pH 9 (k3 = kh). Similarly, fcov is independent of temperature (0-30 degrees C), while k varies by a factor of approx. 3. In the range of 0-1.5 M NaCl, fcov decreases from 0.10 to 0.04. These variations are attributed to a combination of salt-induced variations in the factors k3, Xb and the ratio kc/kT.  相似文献   

18.
Escherichia coli DNA photolyase (photoreactivating enzyme) is a flavoprotein. The enzyme binds to DNA containing pyrimidine dimers in a light-independent step and, upon illumination with 300-600 nm radiation, catalyzes the photosensitized cleavage of the cyclobutane ring thus restoring the integrity of the DNA. We have studied the binding reaction using the techniques of nitrocellulose filter binding and flash photolysis. The enzyme binds to dimer-containing DNA with an association rate constant k1 estimated by two different methods to be 1.4 X 10(6) to 4.2 X 10(6) M-1 S-1. The dissociation of the enzyme from dimer-containing DNA displays biphasic kinetics; for the rapidly dissociating class of complexes k2 = 2-3 X 10(-2) S-1, while for the more slowly dissociating class k2 = 1.3 X 10(-3) to 6 X 10(-4) S-1. The equilibrium association constant KA, as determined by the nitrocellulose filter binding assay and the flash photolysis assay, was 4.7 X 10(7) to 6 X 10(7) M-1, in reasonable agreement with the values predicted from k1 and k2. From the dependence of the association constant on ionic strength we conclude that the enzyme contacts no more than two phosphodiester bonds upon binding; this strongly suggests that the pyrimidine dimer is the main structural determinant of specific photolyase-DNA interaction and that nonspecific ionic interactions do not contribute significantly to substrate binding.  相似文献   

19.
We have demonstrated that amino acids E (323), Y (324), E (330), and V (331) from the factor Va heavy chain are required for the interaction of the cofactor with factor Xa and optimum rates of prothrombin cleavage. We have also shown that amino acid region 332-336 contains residues that are important for cofactor function. Using overlapping peptides, we identified amino acids D (334) and Y (335) as contributors to cofactor activity. We constructed recombinant factor V molecules with the mutations D (334) --> K and Y (335) --> F (factor V (KF)) and D (334) --> A and Y (335) --> A (factor V (AA)). Kinetic studies showed that while factor Va (KF) and factor Va (AA) had a K D for factor Xa similar to the K D observed for wild-type factor Va (factor Va (WT)), the clotting activities of the mutant molecules were impaired and the k cat of prothrombinase assembled with factor Va (KF) and factor Va (AA) was reduced. The second-order rate constant of prothrombinase assembled with factor Va (KF) or factor Va (AA) for prothrombin activation was approximately 10-fold lower than the second-order rate constant for the same reaction catalyzed by prothrombinase assembled with factor Va (WT). We also created quadruple mutants combining mutations in the amino acid region 334-335 with mutations at the previously identified amino acids that are important for factor Xa binding (i.e., E (323)Y (324) and E (330)V (331)). Prothrombinase assembled with the quadruple mutant molecules displayed a second-order rate constant up to 400-fold lower than the values obtained with prothrombinase assembled with factor Va (WT). The data demonstrate that amino acid region 334-335 is required for the rearrangement of enzyme and substrate necessary for efficient catalysis of prothrombin by prothrombinase.  相似文献   

20.
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   

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