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 15 毫秒
1.
The Q7K/L331/L631 HIV-1 protease mutant was expressed in Escherichia coli and the effect of binding a substrate-analog inhibitor, acetyl-pepstatin, was investigated by fluorescence spectroscopy and molecular dynamics. The dimeric enzyme has four intrinsic tryptophans, located at positions 6 and 42 in each monomer. Fluorescence spectra and acrylamide quenching experiments show two differently accessible Trp populations in the apoenzyme with k(q1) = 6.85 x 10(9) M(-1) s(-1) and k(q2) = 1.88 x 10(9) M(-1) s(-1), that merge into one in the complex with k(q) = 1.78 x 10(9) M(-1) s(-1). 500 ps trajectory analysis of Trp X1/X2 rotameric interconversions suggest a model to account for the observed Trp fluorescence. In the simulations, Trp6/Trp6B rotameric interconversions do not occur on this timescale for both HIV forms. In the apoenzyme simulations, however, both Trp42s and Trp42Bs are flipping between X1/X2 states; in the complexed form, no such interconverions occur. A detailed investigation of the local Trp environments sampled during the molecular dynamics simulation suggests that one of the apoenzyme Trp42B rotameric interconversions would allow indole-quencher contact, such as with nearby Tyr59. This could account for the short lifetime component. The model thus interprets the experimental data on the basis of the conformational fluctuations of Trp42s alone. It suggests that the rotameric interconversions of these Trps, located relatively far from the active site and at the very start of the flap region, becomes restrained when the apoenzyme binds the inhibitor. The model is thus consistent with associating components of the fluorescence decay in HIV-1 protease to ground state conformational heterogeneity.  相似文献   

2.
3.
The fluorescent properties of 2-aminopurine (2-AP) incorporated in an RNA sequence are used to study the structural dynamics and local changes of the retroviral RNA structure. Using 2-AP, the conformational states of the unpaired loop adenine in avian leucosis virus RNA were studied upon its interaction with aminoglycoside antibiotics. The intensity of 2-AP fluorescence in the monomeric RNA hairpin was higher than in both RNA dimers. The intensity of fluorescence in the extended dimer was significantly lower than in the kissing loop dimer. The finding was be explained by the fact that stacking contacts in the extended dimer produce a more compact loop structure than in the kissing loop dimer. When the binding of aminogycoside antibiotics with the kissing loop dimer RNA was analyzed, only tobramycin increased the intensity of 2-AP fluorescence almost threefold. The results showed that 2-AP fluorescence is suitable for detecting local changes in complexes of retroviral RNA with ligands.  相似文献   

4.
Metal-ion and sequence dependent changes in the stacking interactions of bases surrounding abasic (AB) sites in 10 different DNA duplexes were examined by incorporating the fluorescent nucleotide probe 2-aminopurine (2-AP), opposite to the site (AB-APopp) or adjacent to the site (AB-APadj) on either strand. A detailed study of the fluorescence emission and excitation spectra of these AB duplexes and their corresponding parent duplexes indicates that AB-APoppis significantly less stacked than 2-AP in the corresponding normal duplex. In general, AB-APadjon the AB strand is stacked, but AB-APadjon the opposite strand shows destabilized stacking interactions. The results also indicate that divalent cation binding to the AB duplexes contributes to destabilizaton of the base stacking interactions of AB-APopp, but has little or no effect on the stacking interactions of AB-APadj. Consistent with these results, the fluorescence of AB-APoppis 18-30-fold more sensitive to an externally added quenching agent than the parent normal duplex. When uracil DNA glycosylase binds to AB-APoppin the presence of 2.5 mM MgCl2, a 3-fold decrease in fluorescence is observed ( K d = 400 +/- 90 nM) indicating that the unstacked 2-APoppbecomes more stacked upon binding. On the basis of these fluorescence studies a model for the local base stacking interactions at these AB sites is proposed.  相似文献   

5.
Stopped-flow measurements have been employed to study the kinetics of the conformational changes in TetR (B) induced by tetracycline binding with and without Mg(2+) ions. Result of stopped-flow fluorometry measurements at pH 8.0 indicate conformational changes in the helix-turn-helix motif in the N-terminal domain and in the C-terminal inducer binding domain. Binding of tetracycline (Tc) to TetR in the absence of Mg(2+) can be described by a simple kinetics process, which is limited to the first step association without any unimolecular conformational change step upon Tc binding. The rate constants for this process are equal to 2.0 x 10(5) M(-)(1) s(-)(1) and 2.1 s(-)(1) for the forward and backward reaction, respectively, and gave the binding constant K(a) = 0.96 x 10(5) M(-)(1). The kinetics of [Tc-Mg](+) binding to TetR can be described by reactions in which the first step describes the association characterized by the rate constants k(a) = 1.4 x 10(5) M(-)(1) s(-)(1) and k(d) = 2.2 x 10(-)(2) s(-)(1) and binding constant K(a) = 6.3 x 10(6) M(-)(1). The first step of [Tc-Mg](+) association is followed by at least three conformational change steps, which occur in the inducer binding site and then propagate to the surroundings of Trp75 and Trp43 residues. The rate constants for the forward, k(c), and backward, k(-)(c), reaction for each of these conformational steps have been determined. The thermodynamics of the binding of tetracycline with and without Mg(2+) to TetR was investigated by isothermal titration calorimetry (ITC) at pH 8.0 and 25 degrees C. The measurement shows that TetR dimer possesses two equivalent binding sites for tetracycline, characterized by binding constant K(a) = 9.0 x 10(6) M(-)(1) and K(a) = 7.0 x 10(4) M(-)(1) for Tc with and without Mg(2+), respectively. The binding of the inducer to TetR, in the presence and absence of Mg(2+) ion, is an enthalpy-driven reaction characterized by DeltaH = -51 kJ mol(-)(1) and DeltaH = -33 kJ mol(-)(1), respectively. The entropy change, DeltaS, for the interaction in the presence of Mg(2+) is equal to -38.9 J K(-)(1) mol(-)(1), and for the tetracycline alone, it was estimated at -17.6 J K(-)(1) mol(-)(1).  相似文献   

6.
The fluorescent base analogue 2-aminopurine (2-AP) is commonly used to study specific conformational and protein binding events involving nucleic acids. Here, combinations of steady-state and time-resolved fluorescence spectroscopy of 2-AP were employed to monitor conformational transitions within a model hairpin RNA from diverse structural perspectives. RNA substrates adopting stable, unambiguous secondary structures were labeled with 2-AP at an unpaired base, within the loop, or inside the base-paired stem. Steady-state fluorescence was monitored as the RNA hairpins made the transitions between folded and unfolded conformations using thermal denaturation, urea titration, and cation-mediated folding. Unstructured control RNA substrates permitted the effects of higher-order RNA structures on 2-AP fluorescence to be distinguished from stimulus-dependent changes in intrinsic 2-AP photophysics and/or interactions with adjacent residues. Thermodynamic parameters describing local conformational changes were thus resolved from multiple perspectives within the model RNA hairpin. These data provided energetic bases for construction of folding mechanisms, which varied among different folding-unfolding stimuli. Time-resolved fluorescence studies further revealed that 2-AP exhibits characteristic signatures of component fluorescence lifetimes and respective fractional contributions in different RNA structural contexts. Together, these studies demonstrate localized conformational events contributing to RNA folding and unfolding that could not be observed by approaches monitoring only global structural transitions.  相似文献   

7.
The photophysics of the complex forming reaction of Ca2+ and Fura-2 are investigated using steady-state and time-resolved fluorescence measurements. The fluorescence decay traces were analyzed with global compartmental analysis yielding the following values for the rate constants at room temperature in aqueous solution with BAPTA as Ca2+ buffer: k01 = 1.2 x 10(9)s-1, k21 = 1.0 x 10(11) M-1 s-1, k02 = 5.5 x 10(8) s-1, k12 = 2.2 x 10(7) s-1, and with EGTA as Ca2+ buffer: k01 = 1.4 x 10(9) s-1, k21 = 5.0 x 10(10) M-1 s-1, k02 = 5.5 x 10(8) s-1, k12 = 3.2 x 10(7) s-1. k01 and k02 denote the respective deactivation rate constants of the Ca2+ free and bound forms of Fura-2 in the excited state. k21 represents the second-order rate constant of binding of Ca2+ and Fura-2 in the excited state, whereas k12 is the first-order rate constant of dissociation of the excited Ca2+:Fura-2 complex. The ionic strength of the solution was shown not to influence the recovered values of the rate constants. From the estimated values of k12 and k21, the dissociation constant K*d in the excited state was calculated. It was found that in EGTA Ca2+ buffer pK*d (3.2) is smaller than pKd (6.9) and that there is negligible interference of the excited-state reaction with the determination of Kd and [Ca2+] from fluorimetric titration curves. Hence, Fura-2 can be safely used as an Ca2+ indicator. From the obtained fluorescence decay parameters and the steady-state excitation spectra, the species-associated excitation spectra of the Ca2+ free and bound forms of Fura-2 were calculated at intermediate Ca2+ concentrations.  相似文献   

8.
Acrylamide quenching is widely used to monitor the solvent exposure of fluorescent probes in vitro. Here, we tested the utility of this technique to discriminate local RNA secondary structures using the fluorescent adenine analogue 2-aminopurine (2-AP). Under native conditions, the solvent accessibilities of most 2-AP-labeled RNA substrates were poorly resolved by classical single-population models; rather, a two-state quencher accessibility algorithm was required to model acrylamide-dependent changes in 2-AP fluorescence in structured RNA contexts. Comparing 2-AP quenching parameters between structured and unstructured RNA substrates permitted the effects of local RNA structure on 2-AP solvent exposure to be distinguished from nearest neighbor effects or environmental influences on intrinsic 2-AP photophysics. Using this strategy, the fractional accessibility of 2-AP for acrylamide ( f a) was found to be highly sensitive to local RNA structure. Base-paired 2-AP exhibited relatively poor accessibility, consistent with extensive shielding by adjacent bases. 2-AP in a single-base bulge was uniformly accessible to solvent, whereas the fractional accessibility of 2-AP in a hexanucleotide loop was indistinguishable from that of an unstructured RNA. However, these studies also provided evidence that the f a parameter reflects local conformational dynamics in base-paired RNA. Enhanced base pair dynamics at elevated temperatures were accompanied by increased f a values, while restricting local RNA breathing by adding a C-G base pair clamp or positioning 2-AP within extended RNA duplexes significantly decreased this parameter. Together, these studies show that 2-AP quenching studies can reveal local RNA structural and dynamic features beyond those that can be measured by conventional spectroscopic approaches.  相似文献   

9.
The unfolding and refolding of creatine kinase (ATP:creatine N-phosphotransferase (CK), EC 2.7.3.2) during denaturation and reactivation by trifluoroethanol (TFE) have been studied. Significant aggregation was observed when CK was denatured at TFE concentrations between 10% and 40% (v/v). 50% TFE (v/v) was used to study the denaturation and unfolding of CK. The activity loss of CK was a very quick process, as was the marked conformational changes during denaturation followed by fluorescence emission spectra and far-ultraviolet CD spectra. DTNB modification and size exclusion chromatography were used to find that CK dissociated and was in its monomer state after denaturation with 50% TFE. Reactivation and refolding were observed after 80-fold dilution of the denatured CK into 0.05 M Tris-HCl buffer, pH 8.0. The denatured CK recovered about 38% activity following a two phase course (k(1)=4.82+/-0.41x10(-3) s(-1), k(2)=0.60+/-0.01x10(-3) s(-1)). Intrinsic fluorescence maximum intensity changes showed that the refolding process also followed biphasic kinetics (k(1)=4.34+/-0.27x10(-3) s(-1), k(2)=0.76+/-0.02x10(-3) s(-1)) after dilution into the proper solutions. The far-ultraviolet CD spectra ellipticity changes at 222 nm during the refolding process also showed a two phase course (k(1)=4.50+/-0.07x10(-3) s(-1), k(2)=1.13+/-0.05x10(-3) s(-1)). Our results suggest that TFE can be used as a reversible denaturant like urea and GuHCl. The 50% TFE induced CK denaturation state, which was referred to as the 'TFE state', and the partially refolded CK are compared with the molten globule state. The aggregation caused by TFE during denaturation is also discussed in this paper.  相似文献   

10.
Coiled coils consist of two or more amphipathic a-helices wrapped around each other to form a superhelical structure stabilized at the interhelical interface by hydrophobic residues spaced in a repeating 3-4 sequence pattern. Dimeric coiled coils have been shown to often form in a single step reaction in which association and folding of peptide chains are tightly coupled. Here, we ask whether such a simple folding mechanism may also apply to the formation of a three-stranded coiled coil. The designed 29-residue peptide LZ16A was shown previously to be in a concentration-dependent equilibrium between unfolded monomer (M), folded dimer (D), and folded trimer (T). We show by time-resolved fluorescence change experiments that folding of LZ16A to D and T can be described by 2M (k1)<==>(k(-1)) D and M + D (k2)<==>(k(-2)) T. The following rate constants were determined (25 degrees C, pH 7): k1 = 7.8 x 10(4) M(-1) s(-1), k(-1) = 0.015 s(-1), k2 = 6.5 x 10(5) M(-1) s(-1), and k(-2) = 1.1 s(-1). In a separate experiment, equilibrium binding constants were determined from the change with concentration of the far-ultraviolet circular dichroism spectrum of LZ16A and were in good agreement with the kinetic rate constants according to K(D) = k1/2k(-1) and K(T) = k2/k(-2). Furthermore, pulsed hydrogen-exchange experiments indicated that only unfolded M and folded D and T were significantly populated during folding. The results are compatible with a two-step reaction in which a subpopulation of association competent (e.g., partly helical) monomers associate to dimeric and trimeric coiled coils.  相似文献   

11.
Heredia VV  Thomson J  Nettleton D  Sun S 《Biochemistry》2006,45(24):7553-7562
The transient kinetics of glucose binding to glucokinase (GK) was studied using stopped-flow fluorescence spectrophotometry to investigate the underlying mechanism of positive cooperativity of monomeric GK with glucose. Glucose binding to GK was shown to display biphasic kinetics that fit best to a reversible two-step mechanism. GK initially binds glucose to form a transient intermediate, namely, E* x glucose, followed by a conformational change to a catalytically competent E x glucose complex. The microscopic rate constants for each step were determined as follows: on rate k1 of 557 M(-1) s(-1) and off rate k(-1) of 8.1 s(-1) for E* x glucose formation, and forward rate k2 of 0.45 s(-1) and reverse rate k(-2) of 0.28 s(-1) for the conformational change from E* x glucose to E x glucose. These results suggest that the enzyme conformational change induced by glucose binding is a reversible, slow event that occurs outside the catalytic cycle (kcat = 38 s(-1)). This slow transition between the two enzyme conformations modulated by glucose likely forms the kinetic foundation for the allosteric regulation. Furthermore, the kinetics of the enzyme conformational change was altered in favor of E x glucose formation in D2O, accompanied by a decrease in cooperativity with glucose (Hill slope of 1.3 in D2O vs 1.7 in H2O). The deuterium solvent isotope effects confirm the role of the conformational change in the magnitude of glucose cooperativity. Similar studies were conducted with GK activating mutation Y214C at the allosteric activator site that is likely involved in the protein domain rearrangement associated with glucose binding. The mutation enhanced equilibrium glucose binding by a combination of effects on both the formation of E* x glucose and an enzyme conformational change to E x glucose. Kinetic simulation by KINSIM supports the conclusion that the kinetic cooperativity of GK arises from slow glucose-induced conformational changes in GK.  相似文献   

12.
The kinetics of promoter binding and open complex formation in bacteriophage T7 RNA polymerase was investigated using 2-aminopurine (2-AP) modified promoters. 2-AP serves as an ideal probe to measure the kinetics of open complex formation because its fluorescence is sensitive to both base-unpairing and base-unstacking and to the nature of the neighboring bases. All four 2-AP bases in the TATA box showed an increase in fluorescence with similar kinetics upon binding to the T7 RNA polymerase, indicating that the TATA sequence becomes unpaired in a concerted manner. The 2-AP at -4 showed a peculiarly large increase in fluorescence upon binding to the T7 RNA polymerase. Based on the recent crystal structure of the T7 RNA polymerase-DNA complex, we propose that the large fluorescence increase is due to unstacking of the 2-AP base at -4 from the guanine at -5, during open complex formation. The unstacking may be a critical event in directing and placing the template strand correctly in the T7 RNA polymerase active site upon promoter melting for template directed RNA synthesis. Based on equilibrium fluorescence and stopped-flow kinetic studies, we propose that a fast form of T7 RNA polymerase binds promoter double-stranded DNA by a three-step mechanism. The initial collision complex or a closed complex, ED(c) is formed with a K(d) of 1.8 microm. This complex isomerizes to an open complex, ED(o1), in an energetically unfavorable reaction with an equilibrium constant of 0.12. The ED(o1) further isomerizes to a more stable open complex, ED(o2), with a rate constant around 300 s(-1). Thus, in the absence of the initiating nucleotide, GTP, the overall equilibrium constant for closed to open complex conversion is 0.5 and the net rate of open complex formation is nearly 150 s(-1).  相似文献   

13.
Dash C  Phadtare S  Deshpande V  Rao M 《Biochemistry》2001,40(38):11525-11532
We present here the first report of a hydrophilic peptidic inhibitor, ATBI, from an extremophilic Bacillus sp. exhibiting a two-step inhibition mechanism against the aspartic proteases, pepsin and F-prot from Aspergillus saitoi. Kinetic analysis shows that these proteases are competitively inhibited by ATBI. The progress curves are time-dependent and consistent with slow-tight binding inhibition: E + I right arrow over left arrow (k(3), k(4)) EI right arrow over left arrow (k(5), k(6)) EI. The K(i) values for the first reversible complex (EI) of ATBI with pepsin and F-prot were (17 +/- 0.5) x 10(-9) M and (3.2 +/- 0.6) x 10(-6) M, whereas the overall inhibition constant K(i) values were (55 +/- 0.5) x 10(-12) M and (5.2 +/- 0.6) x 10(-8) M, respectively. The rate constant k(5) revealed a faster isomerization of EI for F-prot [(2.3 +/- 0.4) x 10(-3) s(-1)] than pepsin [(7.7 +/- 0.3) x 10(-4) s(-1)]. However, ATBI dissociated from the tight enzyme-inhibitor complex (EI) of F-prot faster [(3.8 +/- 0.5) x 10(-5) s(-1)] than pepsin [(2.5 +/- 0.4) x 10(-6) s(-1)]. Comparative analysis of the kinetic parameters with pepstatin, the known inhibitor of pepsin, revealed a higher value of k(5)/k(6) for ATBI. The binding of the inhibitor with the aspartic proteases and the subsequent conformational changes induced were monitored by exploiting the intrinsic tryptophanyl fluorescence. The rate constants derived from the fluorescence data were in agreement with those obtained from the kinetic analysis; therefore, the induced conformational changes were correlated to the isomerization of EI to EI. Chemical modification of the Asp or Glu by WRK and Lys residues by TNBS abolished the antiproteolytic activity and revealed the involvement of two carboxyl groups and one amine group of ATBI in the enzymatic inactivation.  相似文献   

14.
We have investigated the kinetics of complex formation between bacteriophage MS2 coat protein subunits and synthetic RNA fragments encompassing the natural translational operator site, or the consensus sequences of three distinct RNA aptamer families, which are known to bind to the same site on the protein. Reactions were assayed using stopped-flow fluorescence spectroscopy and either the intrinsic tryptophan fluorescence of the protein or the signals from RNA fragments site-specifically substituted with the fluorescent adenosine analogue 2'-deoxy, 2-aminopurine. The kinetics observed were independent of the fluorophore being monitored or its position within the complex, indicating that the data report global events occurring during complex formation. Competition assays show that the complex being formed consists of a single coat protein dimer and one RNA molecule. The binding reaction is at least biphasic. The faster phase, constituting 80-85 % of the amplitude, is a largely diffusion driven RNA-protein interaction (k1 approximately 2x10(9) M(-1) s(-1)). The salt dependence of the forward reaction and the similarities of the on-rates of lower-affinity RNA fragments are consistent with a diffusion-controlled step dominated by electrostatic steering. The slower phase is independent of reactant concentration, and appears to correspond to isomerisation of the coat protein subunit(s) prior to RNA binding (k(iso) approximately 0.23 s(-1)). Measurements with a coat protein mutant (Pro78Asn) show that this phase is not due to cis-trans isomerisation at this residue. The conformational changes in the protein ligand during formation of an RNA-protein complex might play a role in the triggering of capsid self-assembly and a model for this is discussed.  相似文献   

15.
16.
MutS protein initiates mismatch repair with recognition of a non-Watson-Crick base-pair or base insertion/deletion site in DNA, and its interactions with DNA are modulated by ATPase activity. Here, we present a kinetic analysis of these interactions, including the effects of ATP binding and hydrolysis, reported directly from the mismatch site by 2-aminopurine fluorescence. When free of nucleotides, the Thermus aquaticus MutS dimer binds a mismatch rapidly (k(ON)=3 x 10(6) M(-1) s(-1)) and forms a stable complex with a half-life of 10 s (k(OFF)=0.07 s(-1)). When one or both nucleotide-binding sites on the MutS*mismatch complex are occupied by ATP, the complex remains fairly stable, with a half-life of 5-7 s (k(OFF)=0.1-0.14 s(-1)), although MutS(ATP) becomes incapable of (re-)binding the mismatch. When one or both nucleotide-binding sites on the MutS dimer are occupied by ADP, the MutS*mismatch complex forms rapidly (k(ON)=7.3 x 10(6) M(-1) s(-1)) and also dissociates rapidly, with a half-life of 0.4 s (k(OFF)=1.7 s(-1)). Integration of these MutS DNA-binding kinetics with previously described ATPase kinetics reveals that: (a) in the absence of a mismatch, MutS in the ADP-bound form engages in highly dynamic interactions with DNA, perhaps probing base-pairs for errors; (b) in the presence of a mismatch, MutS stabilized in the ATP-bound form releases the mismatch slowly, perhaps allowing for onsite interactions with downstream repair proteins; (c) ATP-bound MutS then moves off the mismatch, perhaps as a mobile clamp facilitating repair reactions at distant sites on DNA, until ATP is hydrolyzed (or dissociates) and the protein turns over.  相似文献   

17.
本文利用荧光、紫外差光谱研究了根霉葡萄糖淀粉酶在盐酸胍变性后的复性、复活动力学。结果表明,该酶在小于4mol/L盐酸胍中变性是可逆的,其复性过程遵循一级反应方程。酶复活过程是由两个一级反应组成的复合反应,构象变化速度与复活过程中较快的反应速度相差无几,这可能是在Trp及Tyr微区的构象变化基本完成之后,酶活力恢复还没有完成造成的。  相似文献   

18.
The self-complementary DNA decamer duplex d(CTGAATTCAG)2 and its modified counterpart d(CTGA[2AP]TTCAG)2, where the innermost adenine (6-aminopurine) has been replaced with the fluorescent analogue 2-aminopurine (2AP), have been studied by fluorescence and NMR spectroscopy and simulated by molecular dynamics. Both decamers are recognized and cleaved by the EcoRI restriction endonuclease. 2D NMR results show that both decamers have a standard B-type conformation below 20 degrees C, though a disturbance exists to the 5' side of the 2AP site which may originate from increased local mobility. The fluorescence and fluorescence anisotropy decays of both decamers, as well as the one containing 2AP in only one chain, were studied as a function of temperature. The data show that the 2AP base exists in a temperature-dependent distribution of states and shows rapid motions, suggesting interconversion among these states on a time scale of about 10(-10) s. The integrated fluorescence of the decamer with 2AP in both chains shows a large increase around the helix melting temperature whereas the decamer with one 2AP shows only a mild increase, showing that the mixed helix has a different structural transition as sensed by the 2AP base. The data suggest a model of conformational states which have distinct fluorescence decay times. The various states may differ in the degree of base stacking. Fluctuations in the degree of stacking of the A or 2AP base are supported by molecular dynamics simulations, which additionally show that the 2AP-T or A-T base pair hydrogen bonds remain intact during these large motions.  相似文献   

19.
20.
The kinetic mechanism of action of Draculin on activated Factor X (FXa) is established. Draculin inhibits activated Factor X within seconds of incubation at near equimolar concentration (2-6 times on molar basis). Fitting the data to the equation for a tight-binding inhibitor gives a value for K(i)(K(d)) = 14.8+/-1.5 nM. The formation of the Draculin-FXa complex can be explained by a two-step mechanism, where for the first, reversible step, k(on) = 1.117 (+/- 0.169, S.E.M.) x 10(6) M(-1)s(-1) and k(off) = 15.388 (+/- 1.672) x 10(-3) s(-1), while for the second, irreversible step, which is concentration-independent, k(2) = 0.072 s(-1). K(d) obtained from k(off)/k(on) = 13.76 nM. Lineweaver-Burk plot shows a noncompetitive behavior. This noncompetitive mode of inhibition of Draculin is supported by the observation that Draculin, at concentrations giving complete inhibition, does not impair binding of p-aminobenzamidine to FXa. Moreover, under the same conditions, Draculin induces <14% decrease of the fluorescence intensity of the p-aminobenzamidine-FXa complex. We conclude that Draculin is a noncompetitive, tight-binding inhibitor of FXa, a characteristic so far unique amongst natural FXa inhibitors.  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