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.
Citrate synthase complexes with the transition-state analog inhibitor, carboxymethyl-CoA (CM-CoA), are believed to mimic those with the activated form of acetyl-CoA. The X-ray structure [Karpusas, M., Branchaud, B., & Remington, S.J. (1990) Biochemistry 29, 2213] of the ternary complex of the enzyme, oxaloacetate, and CMCoA has been used as the basis for a proposal that a neutral enol of acetyl-CoA is that activated form. Since the inhibitor carboxyl has a pKa of 3.90, analogy with an enolic acetyl-CoA intermediate leads to the prediction that a proton should be taken up from solution upon formation of the analog complex so that the transition-state analog carboxyl is protonated when bound. We have obtained evidence in solution for this proposal by comparing the isoelectric points and the pH dependence of the dissociation constants of the ternary complexes of the pig heart enzyme with the neutral ground-state analog inhibitor, acetonyl-CoA (KCoA), and the anionic transition-state analog inhibitor (CMCoA) and by studying the NMR spectra of the transition-state analog complexes of allosteric (Escherichia coli) and nonallosteric (pig heart) enzymes. The pH dependence of the dissociation constant of the ground-state analog indicates no proton uptake, while that for the transition-state analog indicates that 0.55 +/- 0.04 proton is taken up when the analog binds to the citrate synthase-oxaloacetate binary complex. The overall charges of ternary complexes of the pig heart enzyme with the transition-state and ground-state analog inhibitors are the same, as monitored by their isoelectric points.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The two preceding papers [Powers, V. M., Koo, C. W., Kenyon, G. L., Gerlt, J. A., & Kozarich, J. W. (1991) Biochemistry (first paper of three in this issue); Neidhart, D. J., Howell, P. L., Petsko, G. A., Powers, V. M., Li, R., Kenyon, G. L., & Gerlt, J. A. (1991) Biochemistry (second paper of three in this issue)] suggest that the active site of mandelate racemase (MR) contains two distinct general acid/base catalysts: Lys 166, which abstracts the alpha-proton from (S)-mandelate, and His 297, which abstracts the alpha-proton from (R)-mandelate. In this paper we report on the properties of the mutant of MR in which His 297 has been converted to asparagine by site-directed mutagenesis (H297N). The structure of H297N, solved by molecular replacement at 2.2-A resolution, reveals that no conformational alterations accompany the substitution. As expected, H297N has no detectable MR activity. However, H297N catalyzes the stereospecific elimination of bromide ion from racemic p-(bromomethyl)mandelate to give p-(methyl)-benzoylformate in 45% yield at a rate equal to that measured for wild-type enzyme; the unreacted p-(bromomethyl)mandelate is recovered as (R)-p-(hydroxymethyl)mandelate. At pD 7.5, H297N catalyzes the stereospecific exchange of the alpha-proton of (S)- but not (R)-mandelate with D2O solvent at a rate 3.3-fold less than that observed for incorporation of solvent deuterium into (S)-mandelate catalyzed by wild-type enzyme. The pD dependence of the rate of the exchange reaction catalyzed by H297N reveals a pKa of 6.4 in D2O, which is assigned to Lys 166.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
L B Dugad  J T Gerig 《Biochemistry》1988,27(12):4310-4316
Binding of 4-fluorobenzenesulfonamide to human carbonic anhydrases I and II has been studied by proton, fluorine, and nitrogen-15 nuclear magnetic resonance spectroscopy. All three types of experiments provide evidence that the stoichiometry of the interaction of this inhibitor with both enzymes is 2 mol of inhibitor bound per mole of enzyme. Observations which suggest that the bound forms are involved in an exchange process that is rapid at room temperature but slower at 2 degrees C are described. Nitrogen-15 shift data show that the bound inhibitors are present at the active site as anions. The proton experiments indicate appreciable reorganization of the tertiary structure of the protein upon binding. Saturation-transfer experiments to determine the rate of dissociation of the inhibitor-enzyme complex lead to the conclusion that the dissociation process is more complicated than a simple free-bound equilibrium.  相似文献   

4.
As part of a structure-based drug design program directed against enzyme targets in the human immunodeficiency virus (HIV), we have determined the three-dimensional structures of the HIV type 1 protease complexed with two hydroxyethylene-based inhibitors. The inhibitors (SKF 107457 and SKF 108738) are hexapeptide substrate analogues with the scissile bond being replaced by a hydroxyethylene isostere. The structures were determined using x-ray diffraction data to 2.2 A measured at the Cornell High Energy Synchrotron Source on hexagonal crystals of each of the complexes. The structures have been extensively refined using a reciprocal space least-squares method to conventional crystallographic R factors of 0.186 and 0.159, respectively. The protein structure differs from that in the unliganded state of the enzyme and is most similar to that of the structure of the other reported (Jaskolski, M., Tomasselli, A. G., Sawyer, T. K., Staples, D. G., Heinrikson, R. L., Schneider, J., Kent, S. B. H., and Wlodawer, A. (1990) Biochemistry 29, 5889-5907) hydroxyethylene-based inhibitor complex. Unlike in that structure, however, the inhibitors are observed, in the present crystal structures, in two equally abundant orientations that are a consequence of the homodimeric nature of the enzyme coupled with the asymmetric structures of the inhibitors. Although the differences between the two inhibitors used in the present study are confined to the P1' site, the van der Waals interactions made by the inhibitor atoms with the amino acid residues in the protein differ throughout the structures of the inhibitors.  相似文献   

5.
Janc JW  Clark JM  Warne RL  Elrod KC  Katz BA  Moore WR 《Biochemistry》2000,39(16):4792-4800
Serine proteases play a role in a variety of disease states and thus are attractive targets for therapeutic intervention. We report the kinetic characterization of a class of serine protease inhibitors whose potencies and selectivities are dramatically enhanced in the presence of Zn(II). The structural basis for Zn(II)-mediated inhibition of trypsin-like proteases has recently been reported [Katz, B. A., Clark, J. M., Finer-Moore, J. S., Jenkins, T. E., Johnson, C. R., Ross, M. J., Luong, C., Moore, W. R., and Stroud, R. M. (1998) Nature 391, 608-612]. A case study of the kinetic behavior of human tryptase inhibitors is provided to illustrate the general phenomenon of Zn(II)-mediated inhibition. Tryptase, Zn(II), and the inhibitor form a ternary complex which exhibits classic tight-binding inhibition. The half-life for release of inhibitor from the tryptase-Zn(II)-inhibitor complex has been measured for a number of inhibitors. Consistent with tight-binding behavior, potent tryptase inhibitors are characterized by extremely slow rates of dissociation from the ternary complex with half-lives on the order of hours. A model of human serum, designed to reproduce physiological levels of Zn(II), has been employed to evaluate the performance of Zn(II)-potentiated tryptase inhibitors under physiological conditions. We demonstrate that Zn(II)-mediated inhibition can be achieved at physiological Zn(II) levels.  相似文献   

6.
NMR studies of carbonic anhydrase-fluorinated benzenesulfonamide complexes   总被引:1,自引:0,他引:1  
L B Dugad  C R Cooley  J T Gerig 《Biochemistry》1989,28(9):3955-3960
Fluorine NMR has been used to examine complexes formed by 2-fluoro-, 3-fluoro-, and 2,5-difluorobenzenesulfonamide and human carbonic anhydrases I and II. The results show that all three inhibitors form complexes with both isozymes that have 2:1 inhibitor/enzyme stoichiometry. The fluorine spectra observed for all inhibitor-isozyme combinations are consistent either with rapid rotation of the aromatic ring of the inhibitor in the complexes or with preferential binding of only one of the two possible conformations of the inhibitors that are isomeric by virtue of rotation about the C1-C4 bond of the fluoro aromatic ring. Because ring rotation is slow in the case of the pentafluorobenzenesulfonamide-CA I complex, selective binding of rotamers is the explanation of these observations presently favored. A computer graphics study shows that formation of 2:1 complexes of CA I is feasible without appreciable distortion of the protein tertiary structure found in the crystalline state.  相似文献   

7.
The effects of pH upon the C-2 resonances of the 5 histidine residues of Escherichia coli MB 1428 dihydrofolate reductase in binary complexes with methotrexate, aminopterin, folate, methopterin, and trimethoprim were studied by 300-MHz 1H nmr spectroscopy. Three of the five histidine residues, labeled 1, 2, and 3, exhibited similar pK' values and chemical shifts for their C-2 protons in the five binary complexes. One histidine, 4, was quite different in the folate complex and the last histidine, 5 was quite different in the trimethoprim complex. For all five binary complexes, each histidine had a pK' which was significantly different from the other 4 histidines of that complex. Titration of the binary methotrexate complex of a 5,5'-dithiobis(2-nitrobenzoate)-modified enzyme showed that 2 histidines were not perturbed by this modification of Cys 152, and that the alkaline form of histidine 2, the acid form of histidine 4, and, to a lesser extent, the acid form of histidine 3 were slightly perturbed. Titration of the binary methotrexate complex of a N-bromosuccinimide-modified enzyme demonstrated that this modification slightly affected all of the histidines and drastically affected histidine 5. Histidines 3 and 5 of the binary methotrexate complex reacted rapidly with the histidine-specific reagent, ethoxyformic anhydride, while histidines 2 and 4 reacted at a moderate rate and histidine 1 reacted slowly if at all. The local electrostatic environments of the 5 histidine residues as deduced from the crystal structure of the binary complex of the enzyme with methotrexate (Matthews, D.A., Alden, R.A., Bolin, J.T., Freer, S.T., Hamlin, R., Xuong, N., Kraut, J., Poe, M., Williams, M.N., and Hoogsteen, K. (1977) Science 197, 594-597) were used as the basis for proposed assignments of the five histidine C-2 nmr resonances. The assignments were: 1, pK' 7.9 to 8.2, His 124; 2, pK' 7.2 to 7.4, His 141; 3, pK' 6.5 to 6.7, His 149; 4, pK' 5.7 to 6.3, His 114; and 5, pK' 5.2 to 5.9, His 45. The effect of the chemical modifications upon the enzyme's histidine residues were consistent with the assignments, but no direct chemical evidence in support of the assignments was obtained. It was proposed that, since the crystallographic data provided consistent assignments of the histidine nmr data for both native and chemically modified enzyme, the local environment of each of the 5 histidine residues was similar in the crystal and in solution.  相似文献   

8.
The binding of indirubin-5-sulphonate (E226), a potential anti-tumour agent and a potent inhibitor (IC(50) = 35 nm) of cyclin-dependent kinase 2 (CDK2) and glycogen phosphorylase (GP) has been studied by kinetic and crystallographic methods. Kinetic analysis revealed that E226 is a moderate inhibitor of GPb (K(i) = 13.8 +/- 0.2 micro m) and GPa (K(i) = 57.8 +/- 7.1 micro m) and acts synergistically with glucose. To explore the molecular basis of E226 binding we have determined the crystal structure of the GPb/E226 complex at 2.3 A resolution. Structure analysis shows clearly that E226 binds at the purine inhibitor site, where caffeine and flavopiridol also bind [Oikonomakos, N.G., Schnier, J.B., Zographos, S.E., Skamnaki, V.T., Tsitsanou, K.E. & Johnson, L.N. (2000) J. Biol. Chem.275, 34566-34573], by intercalating between the two aromatic rings of Phe285 and Tyr613. The mode of binding of E226 to GPb is similar, but not identical, to that of caffeine and flavopiridol. Comparative structural analyses of the GPb-E226, GPb-caffeine and GPb-flavopiridol complex structures reveal the structural basis of the differences in the potencies of the three inhibitors and indicate binding residues in the inhibitor site that can be exploited to obtain more potent inhibitors. Structural comparison of the GPb-E226 complex structure with the active pCDK2-cyclin A-E226 complex structure clearly shows the different binding modes of the ligand to GPb and CDK2; the more extensive interactions of E226 with the active site of CDK2 may explain its higher affinity towards the latter enzyme.  相似文献   

9.
The individual zinc endoproteinases of the tissue degrading matrix metalloproteinase (MMP) family share a common catalytic architecture but are differentiated with respect to substrate specificity, localization, and activation. Variation in domain structure and more subtle structural differences control their characteristic specificity profiles for substrates from among four distinct classes (Nagase, H., and Woessner, J. F. J. (1999) J. Biol. Chem. 274, 21491-21494). Exploitation of these differences may be decisive for the design of anticancer or other drugs, which should be highly selective for their particular MMP targets. Based on the 1.8-A crystal structure of human neutrophil collagenase (MMP-8) in complex with an active site-directed inhibitor (RO200-1770), we identify and describe new structural determinants for substrate and inhibitor recognition in addition to the primary substrate recognition sites. RO200-1770 induces a major rearrangement at a position relevant to substrate recognition near the MMP-8 active site (Ala206-Asn218). In stromelysin (MMP-3), competing stabilizing interactions at the analogous segment hinder a similar rearrangement, consistent with kinetic profiling of several MMPs. Despite the apparent dissimilarity of the inhibitors, the central 2-hydroxypyrimidine-4,6-dione (barbiturate) ring of the inhibitor RO200-1770 mimics the interactions of the hydroxamate-derived inhibitor batimastat (Grams, F., Reinemer, P., Powers, J. C., Kleine, T., Pieper, M., Tschesche, H., Huber, R., and Bode, W. (1995) Eur. J. Biochem. 228, 830-841) for binding to MMP-8. The two additional phenyl and piperidyl ring substituents of the inhibitor bind into the S1' and S2' pockets of MMP-8, respectively. The crystal lattice contains a hydrogen bond between the O(gamma) group of Ser209 and N(delta)1 of His207 of a symmetry related molecule; this interaction suggests a model for recognition of hydroxyprolines present in physiological substrates. We also identify a collagenase-characteristic cis-peptide bond, Asn188-Tyr189, on a loop essential for collagenolytic activity. The sequence conservation pattern at this position marks this cis-peptide bond as a determinant for triple-helical collagen recognition and processing.  相似文献   

10.
Wang X  Minasov G  Shoichet BK 《Proteins》2002,47(1):86-96
The class A beta-lactamase TEM-1 is a key bacterial resistance enzyme against beta-lactam antibiotics, but little is known about the energetic bases for complementarity between TEM-1 and its inhibitors. Most inhibitors form a covalent adduct with the catalytic Ser70, making the measurement of equilibrium constants, and hence interaction energies, technically difficult. This study evaluates noncovalent interactions within covalent complexes by examining the differential stability of TEM-1 and its inhibitor adducts. The thermal denaturation of TEM-1 follows a two-state, reversible model with a melting temperature (T(m)) of 51.6C and a van't Hoff enthalpy of unfolding (DeltaH(VH)) of 146.2 kcal/mol at pH 7.0. The stability of the enzyme changes on forming an inhibitor adduct. As expected, some inhibitors stabilize TEM-1; transition-state analogues increase the T(m) by up to 3.7C (1.7 kcal/mol). Surprisingly, all beta-lactam covalent acyl--enzyme complexes tested destabilize TEM-1 significantly relative to the apo-enzyme. For instance, the clinically used inhibitor clavulanic acid and the beta-lactamase-resistant beta-lactams moxalactam and imipenem destabilize TEM-1 by over 2.6C (1.2 kcal/mol) in their covalent adducts. Based on the structure of the TEM-1/imipenem complex (Maveyraud et al., J Am Chem Soc 1998;120:9748--52), destabilization by moxalactam and imipenem is thought to be caused by a steric clash between the side-chain of Asn132 and the 6(7)-alpha group of these beta-lactams. To test this hypothesis, the mutant enzyme N132A was made. In contrast with wild-type, the covalent complexes between N132A and both imipenem and moxalactam stabilize the enzyme, consistent with the hypothesis. To investigate the structural bases of this dramatic change in stability, the structure of N132A/imipenem was determined by X-ray crystallography. In the complex with N132A, imipenem adopts a very different conformation from that observed in the wild-type complex, and the putative destabilizing interaction with residue 132 is relieved. Studies of several enzymes suggest that beta-lactams, and covalent inhibitors in general, can have either net favorable or net unfavorable noncovalent interaction energies within the covalent complex. In the case of TEM-1, such unfavorable interactions convert substrate analogues into very effective inhibitors.  相似文献   

11.
Proteinase inhibitors in the serpin family form complexes with serine proteinases by interactions between the gamma-OH group at serine 195 of the enzyme and a specific peptide bond within the reactive site loop of the inhibitor. However, the type of complex formed (i.e. Michaelis, acyl, or tetrahedral) is unknown. Until now, 13C NMR spectroscopy studies have only been useful in examining complexes formed with either peptide-related or small protein inhibitors, where 13C-labeled amino acids can be inserted semi-synthetically. Recombinant DNA technology has, however, made it possible to specifically enrich larger proteins with 13C. In the case of serpins we have examined the structure of the complex formed between human alpha 1-proteinase inhibitor uniformally labeled with [13C]methionine and porcine pancreatic elastase. 13C NMR spectroscopic studies revealed a large upfield chemical shift of the carbonyl signal of Met-358 upon complex formation suggesting for the first time that a tetrahedral adduct is formed between a serpin inhibitor and a serine proteinase.  相似文献   

12.
Transition state analogs pepstatin methylester (PME) and L685458 have been shown to inhibit gamma-secretase non-competitively (Tian, G., Sobotka-Briner, C., Zysk, J., Liu, X., Birr, C., Sylvester, M. A., Edwards, P. D., Scott, C. W., and Greenberg, B. D. (2002) J. Biol. Chem. 277, 31499-31505). This unusual kinetics suggests physical separation of the sites for substrate binding and catalysis with binding of the transition state analogs to the catalytic site and not to the substrate binding site. Methods of inhibitor cross-competition kinetics and competition ligand binding were utilized to address whether non-transition state small molecule inhibitors, which also display non-competitive inhibition of gamma-secretase, inhibit the enzyme by binding to the catalytic site as well. Inhibitor cross-competition kinetics indicated competitive binding between the transition state analogs PME and L685458 and between small molecules arylsulfonamides and benzodiazepines, but non-competitive binding between the transition state analogs and the small molecule inhibitors. These results were indicative of two inhibitor binding sites, one for transition state analogs and the other for non-transition state small molecule inhibitors. The presence of two inhibitor binding sites for two different classes of inhibitors was corroborated by results from competition ligand binding using [3H]L685458 as the radioligand. Although L685458 and PME displaced the radioligand at the same concentrations as for enzyme inhibition, arylsulfonamides and benzodiazepines did not displace the radioligand at their Ki values, a result consistent with the presence of two inhibitor binding sites. These findings provide useful insights into the catalytic and regulatory mechanisms of gamma-secretase that may facilitate the design of novel gamma-secretase inhibitors.  相似文献   

13.
Proton and phosphorus nmr have been used to investigate the double-helical structures of polyriboadenylic acid [poly(A)] formed in acidic solutions (pH < 6). The results obtained at low pH (~4.5) are consistent with the model for the acid poly(A) double helix proposed by Rich [Rich, A., Davies, D. R., Crick, F. H. C. & Watson, J. D. (1961) J. Mol. Biol. 3 , 71–86]. Other models that have been proposed are inconsistent with the nmr data. The nmr measurements have also been used to examine the conformation of poly(A) helix in the half-protonated state. Although the base-stacking arrangement of this state is similar to that observed in the more extensively protonated low-pH state, the phosphate backbone conformation is different from that found in either the neutral or low-pH structures.  相似文献   

14.
L Powers  B M Kincaid 《Biochemistry》1989,28(10):4461-4468
Differences in the methods of analysis of X-ray absorption data used by Powers et al. [Powers, L., Blumberg, W. E., Chance, B., Barlow, C., Leigh, J., Jr., Smith, J., Yonetani, T., Vik, S., & Peisach, J. (1979) Biochim. Biophys. Acta 547, 520-538; Powers, L., Chance, B., Ching, Y., & Angiolillo, P. (1981) Biophys. J. 34, 465-498] and Scott et al. [Scott, R., Schwartz, J., & Cramer S. (1986) Biochemistry 25, 5546-5555] are clarified. In addition, we compare the X-ray absorption data and results for resting cytochrome c oxidase reported by both groups using the same analysis method and conclude apart from any assumptions that the data are not identical.  相似文献   

15.
Prediction of fluorine chemical shifts in proteins   总被引:3,自引:0,他引:3  
D H Gregory  J T Gerig 《Biopolymers》1991,31(7):845-858
Molecular dynamics calculations have been used in an effort to estimate the change in fluorine nmr shielding when a fluorine nucleus enters the tertiary structure of a protein. Considerations of the possible interactions that can define the shift parameter change suggest that van der Waals interactions are the leading determinant of fluorine shifts in proteins, although aromatic ring currents, other magnetic anisotropies, and electrostatic field effects could result in shift distinctions of 1 ppm or smaller. Results of our studies of a fluorine-containing analogue of the ribonuclease A S-protein/S-peptide complex indicate that static structures such as those implied by crystallographic data lead to overestimates of the magnitude of the van der Waals shielding term; molecular dynamics simulations provide indications of the effects of conformational averaging in defining this term. The treatment used predicts the correct direction of the shift change when the fluorine enters this protein environment from aqueous solution and, with an experimentally supported choice of adjustable parameters, gives agreement with the magnitude of the shift.  相似文献   

16.
Nitric oxide (NO), an intercellular messenger and an immuno-cytotoxic agent, is synthesized by the family of nitric oxide synthases (NOS), which are thiolate-ligated, heme-containing monooxygenases that convert L-Arg to L-citrulline and NO in a tetrahydrobiopterin (BH4)-dependent manner, using NADPH as the electron donor. The dioxygen complex of the ferrous enzyme has been proposed to be a key intermediate in the NOS catalytic cycle. In this study, we have generated a stable ferrous-O2 complex of the oxygenase domain of rat neuronal NOS (nNOS) by bubbling O2 through a solution of the dithionite-reduced enzyme at -30 degrees C in a cryogenic solvent containing 50% ethylene glycol. The most stable dioxygen complex is obtained using the oxygenase domain which has been preincubated for an extended length of time at 4 degrees C with BH4/dithiothreitol and NG-methyl-L-arginine, a substrate analogue inhibitor. The O2 complex of the nNOS oxygenase domain thus prepared exhibits UV-visible absorption (maxima at 419 and 553 nm, shoulder at approximately 585 nm) and magnetic circular dichroism spectra that are nearly identical to those of ferrous-O2 cytochrome P450-CAM. Our spectral data are noticeably blue-shifted from those seen at 10 degrees C for a short-lived transient species (lambdamax = 427 nm) for the nNOS oxygenase domain using stopped-flow rapid-scanning spectroscopy [Abu-Soud, H. M., Gachhui, R., Raushel, F. M., and Stuehr, D. J. (1997) J. Biol. Chem. 272, 17349], but somewhat similar to those of a relatively stable O2 adduct of L-Arg-free full-length nNOS (lambdamax = 415-416.5 nm) generated at -30 degrees C [Bec, N., Gorren, A. C. F., Voelder, C., Mayer, B., and Lange, R. (1998) J. Biol. Chem. 273, 13502]. Compared with ferrous-O2 P450-CAM, however, the ferrous-O2 adduct of the nNOS oxygenase domain is considerably more autoxidizable and the O2-CO exchange reaction is noticeably slower. The generation of a stable ferrous-O2 adduct of the nNOS oxygenase domain, as described herein, will facilitate further mechanistic and spectroscopic investigations of this important intermediate.  相似文献   

17.
Synthetic peptides corresponding to the active domain of the heat-stable inhibitor protein of cAMP-dependent protein kinase (Cheng, H.-C., Kemp, B. E., Pearson, R. B., Smith, A. J., Misconi, L., Van Patten, S. M., and Walsh, D. A. (1986) J. Biol. Chem. 261, 989-992) were tested as inhibitors of cGMP-dependent protein kinase. The peptides themselves were not substrates. cGMP-dependent protein kinase activity was assayed using histone H2B and two synthetic peptide substrates. Consistent with previous observations of other peptide inhibitors of this enzyme (Glass, D. B. (1983) Biochem. J. 213, 159-164), the inhibitory peptides had no effect on the phosphorylation of histone H2B, but they competitively inhibited cGMP-dependent phosphorylation of the two peptide substrates. The parent inhibitor peptide, PKI(5-24)amide, and a series of analogs had Ki (or IC50) values for cGMP-dependent protein kinase in the range of 15-190 microM. In contrast to their effects on the cAMP-dependent protein kinase, the inhibitory peptides were substantially less potent with cGMP-dependent protein kinase, and potency was reduced by the presence of the NH2-terminal residues (residues 5-13). We conclude that the two protein kinases share a recognition of the basic amino acid cluster within the pseudosubstrate region of the peptide, but that the cGMP-dependent protein kinase does not recognize additional NH2-terminal determinants that make the inhibitor protein extremely potent toward the cAMP-dependent enzyme. Even- when tested at high concentrations and with peptide substrates, the native inhibitor protein did not inhibit cGMP-dependent protein kinase under assay conditions in which the peptides derived from it were inhibitory. Thus, the native inhibitor protein appears to have structural features which block interaction with the cGMP-dependent enzyme and enhance its selectivity for cAMP-dependent protein kinase.  相似文献   

18.
The mode of binding of the irreversible thermolysin inhibitor ClCH2CO-DL-(N-OH)Leu-OCH3 [Rasnick, D., & Powers, J.C. (1978) Biochemistry 17, 4363-4369] has been determined by X-ray crystallography at a resolution of 2.3 A and the structure of the covalent complex refined to give a crystallographic residual of 17.0%. This is the first such structural study of an active-site-directed covalent complex of a zinc protease. As anticipated by Rasnick and Powers, the inhibitor alkylates Glu-143 in the thermolysin active site, and the hydroxamic acid moiety coordinates the zinc ion. The formation of the covalent complex is associated with a significant shift in a segment of the polypeptide backbone in the vicinity of the active site. This conformational adjustment appears to be necessary to relieve steric hindrance which would otherwise prevent alkylation of Glu-143. It is suggested that this steric hindrance, which occurs for thermolysin but would not be expected for carboxypeptidase A, accounts for the previously inexplicable difference in reactivity of these two metalloproteases toward N-haloacetyl amino acids. The relevance of this steric hindrance to the mechanism of catalysis is discussed. In agreement with previous results [Kester, W. R., & Matthews, B. W. (1977) Biochemistry 16, 2506-2516], it appears that steric hindrance prevents the direct attack of Glu-143 on the carbonyl carbon of an extended substrate, therefore ruling out the anhydride pathway in thermolysin-catalyzed hydrolysis of polypeptide substrates and their ester analogues.  相似文献   

19.
Molecular dynamics (MD) simulations on the complexes of glucoamylase II (471) from Aspergillus awamori var. X100 with two powerful inhibitors, 1-deoxynojirimycin and (+)-lentiginosine, have been performed, in order to build a model for these complexes in solution and to clarify the structure-activity relationship. MD calculations were carried out for 105 ps, over a 15 Å sphere centered on the inhibitors. A 8 Å residue-based cut-off was used, and the calculations were performed with explicit inclusion of solvent molecules. The MD structure of the complex 1-deoxynojirimycin-glucoamylase shows only minor deviations from the available X-ray structure. The MD structure of the complex of (+)-lentiginosine-glucoamylase, obtained by docking the inhibitor into the active site, suggests us a suitable orientation for the molecule into the enzyme cavity, which can rationalize the high inhibition activity found for (+)-lentiginosine towards amyloglucosidase from A. niger.  相似文献   

20.
K Brady  A Z Wei  D Ringe  R H Abeles 《Biochemistry》1990,29(33):7600-7607
The peptidyl trifluoromethyl ketones Ac-Phe-CF3 (1) and Ac-Leu-Phe-CF3 (2) are inhibitors of chymotrypsin. They differ in Ki (20 and 2 microM, respectively) as well as in their kinetics of association with chymotrypsin in that 1 is rapidly equilibrating, with an association rate too fast to be observed by steady-state techniques, while 2 is "slow binding", as defined by Morrison and Walsh [Morrison, J. F., & Walsh, C. T. (1988) Adv. Enzymol. Relat. Areas Mol. Biol. 61, 202], with a second-order association rate constant of 750 M-1 s-1 at pH 7.0 [Imperiali, B., & Abeles, R. (1986) Biochemistry 25, 3760]. The crystallographic structures of the complexes of gamma-chymotrypsin with inhibitors 1 and 2 have been determined in order to establish whether structural or conformational differences can be found which account for different kinetic and thermodynamic properties of the two inhibitors. In both complexes, the active-site Ser 195 hydroxyl forms a covalent hemiketal adduct with the trifluoromethyl ketone moiety of the inhibitor. In both complexes, the trifluoromethyl group is partially immobilized, but differences are observed in the degree of interaction of fluorine atoms with the active-site His 57 imidazole ring, with amide nitrogen NH 193, and with other portions of the inhibitor molecule. The enhanced potency of Ac-Leu-Phe-CF3 relative to Ac-Phe-CF3 is accounted for by van der Waals interactions of the leucine side chain of the inhibitor with His 57 and Ile 99 side chains and by a hydrogen bond of the acetyl terminus with amide NH 216 of the enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号