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 250 毫秒
1.
Salts of pyrilium, thiopyrilium and selenopyrilium derivatives at pH 7.5 and temperature of 25 degrees C are studied for their effect on the catalytic activity of acetyl cholinesterase (EC 3.1.1.7) of human blood erythrocytes and butyryl cholinesterase (EC 3.1.1.8) of horse blood serum which is measured by the method of potentiometric titration. All enumerated salts are established to be strong reversible inhibitors of mixed-type cholinesterases, that is testified by small values of the inhibitory constants: competitive Ki, noncompetitive K'i and generalized K epsilon. Pyrilium and selenopyrilium salts inhibit acetyl cholinesterase of human blood erythrocytes to a higher extent than butyryl cholinesterase of horse blood serum, and thiopyrilium salts inhibit the latter to the highest extent. By the value of the inhibitory effect on acetyl cholinesterase of human blood erythrocytes thiopyrilium salts exceed the analogous pyrilium salts, whereas in experiments with butyl cholinesterase of horse blood serum there is an opposite dependence.  相似文献   

2.
The reversible inhibition of horse blood serum butyrylcholinesterase (Ce 3.1.1.8) hydrolysis of ion substrates of acetyl- and butyrylthiocholines and non-ion substrate of indophenylacetate by N-methyl-4-piperidinylbenzylate and tacrine (1,2,3,4,-tetrahydro-9-aminoacridine) and phosphate buffer and ethanol influence on this process are investigated. The values of competitive Ki, uncompetitive K'i and generalized K sigma inhibitory constants are determined. It is shown that the inhibition effect and reversible inhibition type depend not only on the inhibitor and substrate nature but also on the phosphate buffer concentration and ethanol presence in the reaction mixture.  相似文献   

3.
Transient kinetic data of ATP binding and cleavage by cardiac myosin subfragment 1 (S1) were obtained by fluorescence stopped flow and analyzed by using computer modeling based on a consecutive, reversible two-step mechanism: (formula: see text) where M1 and M12 denote myosin species with enhanced fluorescence and K'O = K0/(K0[ATP] + 1). The kinetic constants K0, k12, k23, and k32 and the fractional contributions of M1 and M12 to the total fluorescence are analyzed over a range of systematically varied solution parameters. The initial ATP binding equilibrium (K0), which decreases with increasing pH, is facilitated by a positively charged protein residue with a pK of 7.1. An active-site charge of +1.5 is determined from the ionic strength dependence. The rate constants k12, k23, and k32 also exhibit pK's near neutrality but increase with increasing pH. The majority of the large (-54 kJ/mol) negative free energy of ATP binding occurs upon S1 isomerization, k12, and a large increase in entropy (183 J/kmol at 15 degrees C) is associated with the cleavage step. The equilibrium constant for the cleavage step, K2, is determined as 3.5 at pH 7.0, 15 degrees C, and 200 mM ionic strength. There are no significant changes in fractional contributions to total fluorescence enhancement due to solvent-dependent conformational changes of S1 in these data. When values for the combined rate constants are calculated and compared with those determined by graphical analysis, it is observed that graphical analysis overestimates the binding rate constant (K0k12) by 25% and the hydrolysis rate constant (k23 + k32) by as much as 30%.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
The catalytic constant (k(cat)) and the second-order association constant of compound II with reducing substrate (k(5)) of horseradish peroxidase C (HRPC) acting on phenols and anilines have been determined from studies of the steady-state reaction velocities (V(0) vs. [S(0)]). Since k(cat)=k(2)k(6)/k(2)+k(6), and k(2) (the first-order rate constant for heterolytic cleavage of the oxygen-oxygen bond of hydrogen peroxide during compound I formation) is known, it has been possible to calculate the first-order rate constant for the transformation of each phenol or aniline by HRPC compound II (k(6)). The values of k(6) are quantitatively correlated to the sigma values (Hammett equation) and can be rationalized by an aromatic substrate oxidation mechanism in which the substrate donates an electron to the oxyferryl group in HRPC compound II, accompanied by two proton additions to the ferryl oxygen atom, one from the substrate and the other the protein or solvent. k(6) is also quantitatively correlated to the experimentally determined (13)C-NMR chemical shifts (delta(1)) and the calculated ionization potentials, E (HOMO), of the substrates. Similar dependencies were observed for k(cat) and k(5). From the kinetic analysis, the absolute values of the Michaelis constants for hydrogen peroxide and the reducing substrates (K(M)(H(2)O(2)) and K(M)(S)), respectively, were obtained.  相似文献   

5.
We link two-allele population models by Haldane and Fisher with Kimura's diffusion approximations of the Wright-Fisher model, by considering continuous-state branching (CB) processes which are either independent (model I) or conditioned to have constant sum (model II). Recent works by the author allow us to further include logistic density-dependence (model III), which is ubiquitous in ecology. In all models, each allele (mutant or resident) is then characterized by a triple demographic trait: intrinsic growth rate r, reproduction variance sigma and competition sensitivity c. Generally, the fixation probability u of the mutant depends on its initial proportion p, the total initial population size z, and the six demographic traits. Under weak selection, we can linearize u in all models thanks to the same master formula u = p + p(1 - p)[g(r)s(r) + g(sigma)s(sigma) + g(c)s(c)] + o(s(r),s(sigma),s(c), where s(r) = r' - r, s(sigma) = sigma-sigma' and s(c) = c - c' are selection coefficients, and g(r), g(sigma), g(c) are invasibility coefficients (' refers to the mutant traits), which are positive and do not depend on p. In particular, increased reproduction variance is always deleterious. We prove that in all three models g(sigma) = 1/sigma and g(r) = z/sigma for small initial population sizes z. In model II, g(r) = z/sigma for all z, and we display invasion isoclines of the 'mean vs variance' type. A slight departure from the isocline is shown to be more beneficial to alleles with low sigma than with high r. In model III, g(c) increases with z like ln(z)/c, and g(r)(z) converges to a finite limit L > K/sigma, where K = r/c is the carrying capacity. For r > 0 the growth invasibility is above z/sigma when z < K, and below z/sigma when z > K, showing that classical models I and II underestimate the fixation probabilities in growing populations, and overestimate them in declining populations.  相似文献   

6.
7.
Biphenyl-4-acyoxylate-4'-N-butylcarbamates 1-8 are synthesized from 4,4'-biphenol and are characterized as the pseudosubstrate inhibitors of acetylcholinesterase. In other words, the inhibitors bind to the enzyme and react with the enzyme to form the tetrahedral intermediates for the K(i) steps, and then the tetrahedral intermediates exclude the leaving groups to form a common N-butycarbamyl enzyme intermediate for the k(c) steps. Due to a linear character of the 4,4'-biphenyl moiety, the 4'-N-butylcarbamate moieties of the inhibitors react with the Ser200 residue of the enzyme while the 4-acyoxylate moieties of the inhibitors, on the other hand, should fit in the peripheral anionic site of the enzyme, which is located at the mouth of the deep active site gorge. Thus, carbamates with varied acyl substituents at the 4-position of the biphenyl ring are good candidates for probing the quantitative structure activity relationships for the peripheral anionic site of the enzyme. The fact that the pK(i), log k(c), and log K(i) values are correlated with neither the Taft substituent constant (sigma*) nor the Taft steric constant (E(s)) indicates that the 4-acyoxylate moieties of the inhibitors are too far away from the reaction center. However, the pK(i), log k(c), and log K(i) values are linearly correlated with the Hansch hydrophobicity constant, pi. The intensity constants (psi) for these correlations are 0.16, -0.035, and 0.13, respectively. These results indicate that interactions between the 4-acyoxylate groups of the inhibitors and the peripheral anionic site of the enzyme are mainly hydrophobic ones. The correlation results are slightly improved by using the two-parameter correlations with the Taft substituent steric constant, E(s), and pi. For pK(i), log k(c), and log K(i)-E(s)-pi correlations, the psi values are 0.21, -0.021, and 0.19, respectively; the intensity constants for steric effect (delta) are 0.08, 0.022, and 0.10, respectively. Besides hydrophobic interactions, the two-parameter correlations also suggest that little steric hindrance occurs for the bulkier inhibitors to pass by the peripheral anionic site of the enzyme.  相似文献   

8.
1. The influence of halogen substituents on the 1,2-dioxygenation of catechols was investigated. The results obtained with the two isoenzymes pyrocatechase I and pyrocatechase II from the haloarene-utilizing Pseudomonas sp. B 13 and the pyrocatechase from benzoate-induced cells of Alcaligenes eutrophus B.9 were compared. 2. Substituents on catechol were found to interfere with O2 binding by the two isoenzymes from Pseudomonas sp. B 13, whereas the Km value for catechol kept constant at different O2 concentrations. 3. Electron-attracting substituents decreased the Km values for catechols. 4. Results from binding studies with substituted catechols demonstrated narrow stereospecificities of pyrocatechase I from pseudomonas sp. B 13 and the pyrocatechase from alcaligenes eutrophus B.9. In contrast, a low steric hindrance by substituents in the binding of catechols with pyrocatechase II was observed. 5. Low pK'1 values of substituted catechols resulted in low Michaelis constants. 6. Electron-attracting substituents such as halogen decreased the reaction rates of catechol 1,2-dioxygenation. The correlation of the Vmax. values observed with pyrocatechase II from Pseudomonas sp. B 13 with the substituent constant sigma+ (Okamoto--Brown equation) was distinctly greater than with Hammett's sigma values. The corresponding logVmax. against sigma+ correlation for pyrocatechase I was considerably disturbed by steric influences of the substituents.  相似文献   

9.
Serotype 3 reoviruses inhibit cellular proliferation by inducing a G(2)/M phase cell cycle arrest. Reovirus-induced G(2)/M phase arrest requires the viral S1 gene-encoded sigma1s nonstructural protein. The G(2)-to-M transition represents a cell cycle checkpoint that is regulated by the kinase p34(cdc2). We now report that infection with serotype 3 reovirus strain Abney, but not serotype 1 reovirus strain Lang, is associated with inhibition and hyperphosphorylation of p34(cdc2). The sigma1s protein is necessary and sufficient for inhibitory phosphorylation of p34(cdc2), since a viral mutant lacking sigma1s fails to hyperphosphorylate p34(cdc2) and inducible expression of sigma1s is sufficient for p34(cdc2) hyperphosphorylation. These studies establish a mechanism by which reovirus can perturb cell cycle regulation.  相似文献   

10.
1. The pH-dependence is considered of a reaction between E and S that proceeds through an intermediate ES under "Briggs-Haldane' conditions, i.e. there is a steady state in ES and [S]o greater than [E]T, where [S]o is the initial concentration of S and [E]T is the total concentration of all forms of E. Reactants and intermediates are assumed to interconvert in three protonic states (E equilibrium ES; EH equilibrium EHS; EH2 equilibrium EH2S), but only EHS provides products by an irreversible reaction whose rate constant is kcat. Protonations are assumed to be so fast that they are all at equilibrium. 2. The rate equation for this model is shown to be v = d[P]/dt = (kcat.[E]T[S]o/A)/[(KmBC/DA) + [S]o], where Km is the usual assembly of rate constants around EHS and A-D are functions of the form (1 + [H]/K1 + K2/[H]), in which K1 and K2 are: in A, the molecular ionization constants of ES; in B, the analogous constants of E; in C and D, apparent ionization constants composed of molecular ionization constants (of E or ES) and assemblies of rate constants. 3. As in earlier treatments of this type of reaction which involve either the assumption that the reactants and intermediate are in equilibrium or the assumption of Peller & Alberty [(1959) J. Am. Chem. Soc. 81, 5907-5914] that only EH and EHS interconvert directly, the pH-dependence of kcat. is determined only by A. 4. The pH-dependence of Km is determined in general by B-C/A-D, but when reactants and intermediate are in equilibrium, C identical to D and this expression simplifies to B/A. 5. The pH-dependence of kcat./Km, i.e. of the rate when [S]o less than Km, is not necessarily a simple bell-shaped curve characterized only by the ionization constants of B, but is a complex curve characterized by D/B-C. 6. Various situations are discussed in which the pH-dependence of kcat./Km is determined by assemblies simpler than D/B-C. The special situation in which a kcat./Km-pH profile provides the molecular pKa values of the intermediate ES complex is delineated.  相似文献   

11.
Widely distributed flavohemoglobins (flavoHbs) function as NO dioxygenases and confer upon cells a resistance to NO toxicity. FlavoHbs from Saccharomyces cerevisiae, Alcaligenes eutrophus, and Escherichia coli share similar spectra, O(2), NO, and CO binding kinetics, and steady-state NO dioxygenation kinetics. Turnover numbers (V(max)) for S. cerevisiae, A. eutrophus, and E. coli flavoHbs are 112, 290, and 365 NO heme(-1) s(-1), respectively, at 37 degrees C with 200 microm O(2). The K(M) values for NO are low and range from 0.1 to 0.25 microm. V(max)/K(M)(NO) ratios of 900-2900 microm(-1) s(-1) indicate an extremely efficient dioxygenation mechanism. Approximate K(M) values for O(2) range from 60 to 90 microm. NO inhibits the dioxygenases at NO:O(2) ratios of > or =1:100 and makes true K(M)(O(2)) values difficult to determine. High and roughly equal second order rate constants for O(2) and NO association with the reduced flavoHbs (17-50 microm(-1) s(-1)) and small NO dissociation rate constants suggest that NO inhibits the dioxygenase reaction by forming inactive flavoHbNO complexes. Carbon monoxide also binds reduced flavoHbs with high affinity and competitively inhibits NO dioxygenases with respect to O(2) (K(I)(CO) = approximately 1 microm). These results suggest that flavoHbs and related hemoglobins evolved as NO detoxifying components of nitrogen metabolism capable of discriminating O(2) from inhibitory NO and CO.  相似文献   

12.
The spectral interaction of a homologous series of alkyl-substituted benzenes and related compounds with purified mammalian cytochrome P-450 has been investigated. Each of the 10 hydrocarbons produced a Type I spectral change, indicative of a low to high spin transition of the haem iron of cytochrome P-450. The extent of perturbation of the cytochrome P-450 spin equilibrium varied for each compound and was used to quantify the spin shifts of the haemoprotein and consequently the substrate-bound spin equilibrium constant, K2. Molecular orbital calculations were utilised to determine the electronic structural parameters of the 10 hydrocarbons investigated, including the electrophilic and nucleophilic super-delocalizabilities summed over all atoms (sigma SE and sigma SN, respectively), the sum of the absolute values of net atomic charge (sigma QT) and the energy levels of both the highest occupied and lowest unoccupied molecular orbitals (E (HOMO) and E (LUMO) respectively). Multiple regression analyses were then utilized to generate quantitative structure-activity relationships between the above structural parameters and the substrate-bound spin equilibrium constant, K2. Good correlations were observed between sigma SE, sigma SN and sigma QT, indicating the importance of hydrophobicity and steric factors in the perturbation of the haemoprotein spin equilibrium. In addition, the electron-accepting potential of the hydrocarbons was an important structural feature and exhibited better correlations with K2 than the electron donating parameter. Taken collectively, our data show the importance of the hydrophobic and charge transfer characteristics of hydrocarbon substrates in dictating the position of the cytochrome P-450 spin equilibrium, and as such, provides a rational molecular explanation based on sound chemical principles for the differential interaction of hydrocarbons with cytochrome P-450.  相似文献   

13.
The C-terminal domain of human extracellular superoxide dismutase (hEC-SOD) plays a crucial role in the protein's interaction with heparin. Here we investigated this interaction in more detail by comparing the heparin-binding characteristics of two variants of hEC-SOD: the two fusion proteins containing the hEC-SOD C-terminal domain and a synthetic peptide homologous to the C-terminal. The interaction studies were performed using a surface plasmon resonance based technique on a BIAcore system. It should be emphasized that this is a model system. However, the kinetic constants, as measured, are valid in a comparative sense. Comparison of affinities for size-fractionated heparins revealed that octa- or decasaccharides are the smallest heparin fragments that can efficiently interact with the C-terminal domain of hEC-SOD. At physiological salt concentration, and pH 7.4, the hEC-SOD/heparin interaction was found to be of a high-affinity type, with an equilibrium dissociation constant, K(d), of 0.12 microM, which is 700 and 10-20 times lower than the K(d) values for the synthetic peptide and the fusion proteins, respectively. However, when an alpha-helical structure was induced in the synthetic peptide, by addition of 10% trifluoroethanol, the K(d) decreased to 0.64 microM. The differences in the K(d) values were mainly governed by differences in the association rate constants (k(ass)). The hEC-SOD/heparin interaction itself was found to have a fairly high dissociation rate constant (0.1 s(-)(1)), and a very high association rate constant (8 x 10(5) M(-)(1) s(-)(1)), suggesting that the interaction is mainly controlled by the association. These results together with circular dichroism spectra of the synthetic peptide suggest that an alpha-helical structure in the C-terminal is essential for optimal binding to heparin and that other parts of hEC-SOD moderate the affinity. Our data also demonstrate that the tetramerization itself does not substantially increase the affinity.  相似文献   

14.
Wang H  Vath GM  Gleason KJ  Hanna PE  Wagner CR 《Biochemistry》2004,43(25):8234-8246
Arylamine N-acetyltransferases (NATs) catalyze an acetyl group transfer from acetyl coenzyme A (AcCoA) to arylamines, hydrazines, and their N-hydroxylated arylamine metabolites. The recently determined three-dimensional structures of prokaryotic NATs have revealed a cysteine protease-like Cys-His-Asp catalytic triad, which resides in a deep and hydrophobic pocket. This catalytic triad is strictly conserved across all known NATs, including hamster NAT2 (Cys-68, His-107, and Asp-122). Treatment of NAT2 with either iodoacetamide (IAM) or bromoacetamide (BAM) at neutral pH rapidly inactivated the enzyme with second-order rate constants of 802.7 +/- 4.0 and 426.9 +/- 21.0 M(-1) s(-1), respectively. MALDI-TOF and ESI mass spectral analysis established that Cys-68 is the only site of alkylation by IAM. Unlike the case for cysteine proteases, no significant inactivation was observed with either iodoacetic acid (IAA) or bromoacetic acid (BAA). Pre-steady state and steady state kinetic analysis with p-nitrophenyl acetate (PNPA) and NAT2 revealed a single-exponential curve for the acetylation step with a second-order rate constant of (1.4 +/- 0.05) x 10(5) M(-1) s(-1), followed by a slow linear rate of (7.85 +/- 0.65) x 10(-3) s(-1) for the deacetylation step. Studies of the pH dependence of the rate of inactivation with IAM and the rate of acetylation with PNPA revealed similar pK(a)(1) values of 5.23 +/- 0.09 and 5.16 +/- 0.04, respectively, and pK(a)(2) values of 6.95 +/- 0.27 and 6.79 +/- 0.25, respectively. Both rates reached their maximum values at pH 6.4 and decreased by only 30% at pH 9.0. Kinetic studies in the presence of D(2)O revealed a large inverse solvent isotope effect on both inactivation and acetylation of NAT2 [k(H)(inact)/k(D)(inact) = 0.65 +/- 0.02 and (k(2)/K(m)(acetyl))(H)/(k(2)/K(m)(acetyl))(D) = 0.60 +/- 0.03], which were found to be identical to the fractionation factors (Phi) derived from proton inventory studies of the rate of acetylation at pL 6.4 and 8.0. Substitution of the catalytic triad Asp-122 with either alanine or asparagine resulted in the complete loss of protein structural integrity and catalytic activity. From these results, it can be concluded that the catalytic mechanism of NAT2 depends on the formation of a thiolate-imidazolium ion pair (Cys-S(-)-His-ImH(+)). However, in contrast to the case with cysteine proteases, a pH-dependent protein conformational change is likely responsible for the second pK(a), and not deprotonation of the thiolate-imidazolium ion. In addition, substitutions of the triad aspartate are not tolerated. The enzyme appears, therefore, to be engineered to rapidly form a stable acetylated species poised to react with an arylamine substrate.  相似文献   

15.
16.
17.
The role of hydrophobic and electronic effects on the kinetic constants kcat and Km for the papain hydrolysis of a series of 22 substituted N-benzoylglycine pyridyl esters was investigated. The series studied comprises a wide variety of substituents on the N-benzoyl ring, with about a 300,000-fold range in their hydrophobicities, and 2.1-fold range in their electronic Hammet constants (sigma). It was found that the variation in the log kcat and log 1/Km constants could be explained by the following quantitative-structure activity relationships (QSAR): log 1/Km = 0.40 pi 4 + 4.40 and log 1/kcat = 0.45 sigma + 0.18. The substituent constant, pi 4, is the hydrophobic parameter for the 4-N-benzoyl substituents. QSAR analysis of two smaller sets of glycine phenyl and methyl esters produced similar results. A clear separation of the substituent effects indicates that in the case of these particular esters, acylation appears to be the rate limiting catalytic step.  相似文献   

18.
Badellino KO  Walsh PN 《Biochemistry》2000,39(16):4769-4777
Protease nexin II, a platelet-secreted protein containing a Kunitz-type domain, is a potent inhibitor of factor XIa with an inhibition constant of 250-400 pM. The present study examined the protein interactions responsible for this inhibition. The isolated catalytic domain of factor XIa is inhibited by protease nexin II with an inhibition constant of 437 +/- 62 pM, compared to 229 +/- 40 pM for the intact protein. Factor XIa is inhibited by a recombinant Kunitz domain with an inhibition constant of 344 +/- 37 pM versus 422 +/- 33 pM for the catalytic domain. Kinetic rate constants were determined by progress curve analysis. The association rate constants for inhibition of factor XIa by protease nexin II [(3.35 +/- 0.35) x 10(6) M(-1) s(-1)] and catalytic domain [(2.27 +/- 0. 25) x 10(6) M(-1) s(-1)] are nearly identical. The dissociation rate constants are very similar, (9.17 +/- 0.71) x 10(-4) and (7.97 +/- 1.1) x 10(-4) s(-1), respectively. The rate constants for factor XIa and catalytic domain inhibition by recombinant Kunitz domain are also very similar: association constants of (3.19 +/- 0.29) x 10(6) and (3.25 +/- 0.44) x 10(6) M(-1) s(-1), respectively; dissociation constants of (10.73 +/- 0.84) x 10(-4) and (10.36 +/- 1.3) x 10(-4) s(-1). The inhibition constant (K(i)) values calculated from these kinetic parameters are in close agreement with those measured from equilibrium binding experiments. These results suggest that the major interactions required for factor XIa inhibition by protease nexin II are localized to the catalytic domain of factor XIa and the Kunitz domain of protease nexin II.  相似文献   

19.
In this work, we analyzed the role of the PI3K-p70 S6 kinase (S6K) signaling cascade in the stimulation of endothelial cell proliferation. We found that inhibitors of the p42/p44 MAPK pathway (PD98059) and the PI3K-p70 S6K pathway (wortmannin, Ly294002, and rapamycin) all block thymidine incorporation stimulated by fetal calf serum in the resting mouse endothelial cell line 1G11. The action of rapamycin can be generalized, since it completely inhibits the mitogenic effect of fetal calf serum in primary endothelial cell cultures (human umbilical vein endothelial cells) and another established capillary endothelial cell line (LIBE cells). The inhibitory effect of rapamycin is only observed when the inhibitor is added at the early stages of G(0)-G(1) progression, suggesting an inhibitory action early in G(1). Rapamycin completely inhibits growth factor stimulation of protein synthesis, which perfectly correlates with the inhibition of cell proliferation. In accordance with its inhibitory action on protein synthesis, activation of cyclin D1 and p21 proteins by growth factors is also blocked by preincubation with rapamycin. Expression of a p70 S6K mutant partially resistant to rapamycin reverses the inhibitory effect of the drug on DNA synthesis, indicating that rapamycin action is via p70 S6K. Thus, in vascular endothelial cells, activation of protein synthesis via p70 S6K is an essential step for cell cycle progression in response to growth factors.  相似文献   

20.
NMR studies of the binding of a substrate to an inactive HIV-1 protease construct, containing an active site mutation PR(D25N), are reported. Substrate titration measurements monitored by HSQC spectra and a (15)N-edited NOESY experiment show that the chromogenic substrate analog of the capsid/p2 cleavage site binds to PR(D25N) with an equilibrium dissociation constant, K(D), of 0.27 +/- 0.05 mM, and upper limits of the association and dissociation rate constants, 2 x 10(4) M(-1)s(-1) and 10 s(-1), respectively, at 20 degrees C, pH 5.8. This association rate constant is not in the diffusion limit, suggesting that association is controlled by a rare event, such as opening of the protease flaps. Analysis of (15)N relaxation experiments reveals a slight reduction of S(2) values in the flap region, indicating a small increase in the amplitude of internal motion on the sub-nsec timescale. In addition, several residues in the flap region are mobile on the conformational exchange timescale, msec-microsec. Flap dynamics of the protease-substrate complex are compared with those of protease-inhibitor complexes, and the implications of these results for substrate-binding models are discussed.  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