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.
Cosgrove S  Rogers L  Hewage CM  Malthouse JP 《Biochemistry》2007,46(39):11205-11215
Z-Ala-Ala-Phe-glyoxal (where Z is benzyloxycarbonyl) has been shown to be a competitive inhibitor of pepsin with a Ki = 89 +/- 24 nM at pH 2.0 and 25 degrees C. Both the ketone carbon (R13COCHO) and the aldehyde carbon (RCO13CHO) of the glyoxal group of Z-Ala-Ala-Phe-glyoxal have been 13C-enriched. Using 13C NMR, it has been shown that when the inhibitor is bound to pepsin, the glyoxal keto and aldehyde carbons give signals at 98.8 and 90.9 ppm, respectively. This demonstrates that pepsin binds and preferentially stabilizes the fully hydrated form of the glyoxal inhibitor Z-Ala-Ala-Phe-glyoxal. From 13C NMR pH studies with glyoxal inhibitor, we obtain no evidence for its hemiketal or hemiacetal hydroxyl groups ionizing to give oxyanions. We conclude that if an oxyanion is formed its pKa must be >8.0. Using 1H NMR, we observe four hydrogen bonds in free pepsin and in pepsin/Z-Ala-Ala-Phe-glyoxal complexes. In the pepsin/pepstatin complex an additional hydrogen bond is formed. We examine the effect of pH on hydrogen bond formation, but we do not find any evidence for low-barrier hydrogen bond formation in the inhibitor complexes. We conclude that the primary role of hydrogen bonding to catalytic tetrahedral intermediates in the aspartyl proteases is to correctly orientate the tetrahedral intermediate for catalysis.  相似文献   

2.
Benzyloxycarbonyl (Z)-Ala-Pro-Phe-glyoxal and Z-Ala-Ala-Phe-glyoxal have both been shown to be inhibitors of alpha-chymotrypsin with minimal Ki values of 19 and 344 nM, respectively, at neutral pH. These Ki values increased at low and high pH with pKa values of approximately 4.0 and approximately 10.5, respectively. By using surface plasmon resonance, we show that the apparent association rate constant for Z-Ala-Pro-Phe-glyoxal is much lower than the value expected for a diffusion-controlled reaction. 13C NMR has been used to show that at low pH the glyoxal keto carbon is sp3-hybridized with a chemical shift of approximately 100.7 ppm and that the aldehyde carbon is hydrated with a chemical shift of approximately 91.6 ppm. The signal at approximately 100.7 ppm is assigned to the hemiketal formed between the hydroxy group of serine 195 and the keto carbon of the glyoxal. In a slow exchange process controlled by a pKa of approximately 4.5, the aldehyde carbon dehydrates to give a signal at approximately 205.5 ppm and the hemiketal forms an oxyanion at approximately 107.0 ppm. At higher pH, the re-hydration of the glyoxal aldehyde carbon leads to the signal at 107 ppm being replaced by a signal at 104 ppm (pKa approximately 9.2). On binding either Z-Ala-Pro-Phe-glyoxal or Z-Ala-Ala-Phe-glyoxal to alpha-chymotrypsin at 4 and 25 degrees C, 1H NMR is used to show that the binding of these glyoxal inhibitors raises the pKa value of the imidazolium ion of histidine 57 to a value of >11 at both 4 and 25 degrees C. We discuss the mechanistic significance of these results, and we propose that it is ligand binding that raises the pKa value of the imidazolium ring of histidine 57 allowing it to enhance the nucleophilicity of the hydroxy group of the active site serine 195 and lower the pKa value of the oxyanion forming a zwitterionic tetrahedral intermediate during catalysis.  相似文献   

3.
New trypsin inhibitors Z-Lys-COCHO and Z-Lys-H have been synthesised. Ki values for Z-Lys-COCHO, Z-Lys-COOH, Z-Lys-H and Z-Arg-COOH have been determined. The glyoxal group (–COCHO) of Z-Lys-COCHO increases binding ~300 fold compared to Z-Lys-H. The α-carboxylate of Z-Lys-COOH has no significant effect on inhibitor binding. Z-Arg-COOH is shown to bind ~2 times more tightly than Z-Lys-COOH. Both Z-Lys-13COCHO and Z-Lys-CO13CHO have been synthesized. Using Z-Lys-13COCHO we have observed a signal at 107.4 ppm by 13C NMR which is assigned to a terahedral adduct formed between the hydroxyl group of the catalytic serine residue and the 13C-enriched keto-carbon of the inhibitor glyoxal group. Z-Lys-CO13CHO has been used to show that in this tetrahedral adduct the glyoxal aldehyde carbon is not hydrated and has a chemical shift of 205.3 ppm. Hemiketal stabilization is similar for trypsin, chymotrypsin and subtilisin Carlsberg. For trypsin hemiketal formation is optimal at pH 7.2 but decreases at pHs 5.0 and 10.3. The effective molarity of the active site serine hydroxyl group of trypsin is shown to be 25300 M. At pH 10.3 the free glyoxal inhibitor rapidly (t1/2=0.15 h) forms a Schiff base while at pH 7 Schiff base formation is much slower (t1/2=23 h). Subsequently a free enol species is formed which breaks down to form an alcohol product. These reactions are prevented in the presence of trypsin and when the inhibitor is bound to trypsin it undergoes an internal Cannizzaro reaction via a C2 to C1 alkyl shift producing an α-hydroxycarboxylic acid.  相似文献   

4.
The stromelysin-1 catalytic domain(83-247) (SCD) is stable for at least 16 h at pHs 6.0-8.4. At pHs 5.0 and 9.0 there is exponential irreversible denaturation with half lives of 38 and 68 min respectively. At pHs 4.5 and 10.0 irreversible denaturation is biphasic. At 25°C, C-terminal truncation of stromelysin-1 decreases the stability of the stromelysin-1 catalytic domain at pH values >8.4 and <6.0. We describe the conversion of the carboxylate group of (βR)-β-[[[(1S)-1-[[[(1S)-2-Methoxy-1-phenylethyl]amino]carbonyl]-2,2-dimethylpropyl]amino]carbonyl]-2-methyl-[1,1'-biphenyl]-4-hexanoic acid (UK-370106-COOH) a potent inhibitor of the metalloprotease stromelysin-1 to a glyoxal group (UK-370106-CO(13)CHO). At pH 5.5-6.5 the glyoxal inhibitor is a potent inhibitor of stromelysin-1 (K(i)=~1μM). The aldehyde carbon of the glyoxal inhibitor was enriched with carbon-13 and using carbon-13 NMR we show that the glyoxal aldehyde carbon is fully hydrated when it is in aqueous solutions (90.4ppm) and also when it is bound to SCD (~92.0ppm). We conclude that the hemiacetal hydroxyl groups of the glyoxal inhibitor are not ionised when the glyoxal inhibitor is bound to SCD. The free enzyme pK(a) values associated with inhibitor binding were 5.9 and 6.2. The formation and breakdown of the signal at ~92ppm due to the bound UK-370106-CO(13)CHO inhibitor depends on pK(a) values of 5.8 and 7.8 respectively. No strong hydrogen bonds are present in free SCD or in SCD-inhibitor complexes. We conclude that the inhibitor glyoxal group is not directly coordinated to the catalytic zinc atom of SCD.  相似文献   

5.
The aldehyde inhibitor Z-Ala-Ala-Phe-CHO has been synthesized and shown by 13C-NMR to react with the active site serine hydroxyl group of alpha-chymotrypsin to form two diastereomeric hemiacetals. For both hemiacetals oxyanion formation occurs with a pKa value of ~ 7 showing that chymotrypsin reduces the oxyanion pKa values by ~ 5.6 pKa units and stabilizes the oxyanions of both diastereoisomers by ~ 32 kJ mol− 1. As pH has only a small effect on binding we conclude that oxyanion formation does not have a significant effect on binding the aldehyde inhibitor. By comparing the binding of Z-Ala-Ala-Phe-CHO with that of Z-Ala-Ala-Phe-H we estimate that the aldehyde group increases binding ~ 100 fold. At pH 7.2 the effective molarity of the active site serine hydroxy group is ~ 6000 which is ~ 7 × less effective than with the corresponding glyoxal inhibitor. Using 1H-NMR we have shown that at both 4 and 25 °C the histidine pKa is ~ 7.3 in free chymotrypsin and it is raised to ~ 8 when Z-Ala-Ala-Phe-CHO is bound. We conclude that oxyanion formation only has a minor role in raising the histidine pKa and that the aldehyde hydrogen must be replaced by a larger group to raise the histidine pKa > 10 and give stereospecific formation of tetrahedral intermediates. The results show that a large increase in the pKa of the active site histidine is not needed for the active site serine hydroxyl group to have an effective molarity of 6000.  相似文献   

6.
The chemical shift of the carboxylate carbon of Z-tryptophan is increased from 179.85 to 182.82 ppm and 182.87 ppm on binding to thermolysin and stromelysin-1 respectively. The chemical shift of Z-phenylalanine is also increased from 179.5 ppm to 182.9 ppm on binding to thermolysin. From pH studies we conclude that the pK(a) of the inhibitor carboxylate group is lowered by at least 1.5 pK(a) units when it binds to either enzyme. The signal at ~183 ppm is no longer observed when the active site zinc atom of thermolysin or stromelysin-1 is replaced by cobalt. We estimate that the distance of the carboxylate carbon of Z-[1-(13)C]-L-tryptophan is ≤3.71? from the active site cobalt atom of thermolysin. We conclude that the side chain of Z-[1-(13)C]-L-tryptophan is not bound in the S(2)' subsite of thermolysin. As the chemical shifts of the carboxylate carbons of the bound inhibitors are all ~183 ppm we conclude that they are all bound in a similar way most probably with the inhibitor carboxylate group directly coordinated to the active site zinc atom. Our spectrophotometric results confirm that the active site zinc atom is tetrahedrally coordinated when the inhibitors Z-tryptophan or Z-phenylalanine are bound to thermolysin.  相似文献   

7.
T C Liang  R H Abeles 《Biochemistry》1987,26(24):7603-7608
A dipeptidyl trifluoromethyl ketone, N-acetyl-L-leucyl-L-[1-13C]phenylalanyl trifluoromethyl ketone, was synthesized. This compound inhibits chymotrypsin with Ki = 1.2 microM [Imperiali B., & Abeles, R.H. (1986) Biochemistry 25, 3760-3767]. The complex formed between this inhibitor and alpha-chymotrypsin was examined with 1H, 13C, and 19F NMR spectroscopy to establish its structure in solution. The keto group of the trifluoro ketone is present as an ionized hemiketal group as deduced from the comparison of its 13C chemical shift with those of model hemiketals. The pKa of the hemiketal hydroxyl in the complex is approximately 4.9, which is about 4.2 units lower than the pKa of model hemiketals. This observation provides direct evidence that serine proteases are able to stabilize the oxyanions of tetrahedral adducts. Evidence is also presented for the presence of an Asp-His H bond and protonation of the imidazole group of His-57 in the tetrahedral adduct. The pKa of His-57 is higher than 10. This observation directly indicates that the pKa of His-57 is elevated in a complex containing a tetrahedral adduct.  相似文献   

8.
The chemical shift of the carboxylate carbon of Z-tryptophan is increased from 179.85 to 182.82 ppm and 182.87 ppm on binding to thermolysin and stromelysin-1 respectively. The chemical shift of Z-phenylalanine is also increased from 179.5 ppm to 182.9 ppm on binding to thermolysin. From pH studies we conclude that the pKa of the inhibitor carboxylate group is lowered by at least 1.5 pKa units when it binds to either enzyme. The signal at ~ 183 ppm is no longer observed when the active site zinc atom of thermolysin or stromelysin-1 is replaced by cobalt. We estimate that the distance of the carboxylate carbon of Z-[1-13C]-L-tryptophan is ≤ 3.71 Å from the active site cobalt atom of thermolysin. We conclude that the side chain of Z-[1-13C]-L-tryptophan is not bound in the S2′ subsite of thermolysin. As the chemical shifts of the carboxylate carbons of the bound inhibitors are all ~ 183 ppm we conclude that they are all bound in a similar way most probably with the inhibitor carboxylate group directly coordinated to the active site zinc atom. Our spectrophotometric results confirm that the active site zinc atom is tetrahedrally coordinated when the inhibitors Z-tryptophan or Z-phenylalanine are bound to thermolysin.  相似文献   

9.
The synthetic ketone peptide analogue of pepstatin, isovaleryl-L-valyl-[3-13C]-(3-oxo-4S)-amino-6-methylheptanoyl-L-al anyl-isoamylamide is a strong inhibitor of aspartyl proteases. When the peptide is added to porcine pepsin in H2O at pH 5.1, the 13C NMR chemical shift of the ketone carbon moves from 208 ppm for the inhibitor in solution to 99.07 ppm when bound to the enzyme active site. In 2H2O the bound shift is 98.71 ppm, 0.36 ppm upfield. For the analogous experiment contrasting H216O and H218O, the 13C chemical shift was 0.05 ppm to higher field for the heavier isotope. These data show that water, and not an enzyme nucleophile, adds to the peptide carbonyl to yield a tetrahedral diol adduct in the enzyme-catalyzed reaction, and provide a method for differentiating between covalent and non-covalent mechanisms.  相似文献   

10.
Phenylalanine chloromethyl ketone covalently attached to porous glass beads was synthesized to serve as a solid-phase active site directed inhibitor of chymotrypsin-like proteolytic enzymes. The solid-phase reagent inhibited 20 nmol of bovine chymotrypsin per gram of glass and covalently bound 30 nmol of protein per gram of glass. Sepharose-bound lysine chloromethyl ketones were synthesized to serve as inhibitors of trypsin-like enzymes. Sepharose-MethionylLysyl chloromethyl ketone inactivated and bound about 6.8 nmol of enzyme per ml of settled gel. In a preliminary experiment, a cyanogen bromide cleavage of the methionine residues showed that it should be possible to release all peptides but the peptide containing the active-site histidine. The immobilized trypsin was also reduced, carboxymethylated and digested with chymotrypsin. The potential of the solid-phase approach is in the isolation of a specific serine proteinase and in the sequence determination of residues surrounding the active-site histidine.  相似文献   

11.
Summary Hydroxyl groups of serine and threonine, and to some extent also tyrosine are usually located on or near the surface of proteins. NMR observations of the hydroxyl protons is therefore of interest to support investigations of the protein surface in solution, and knowledge of the hydroxyl NMR lines is indispensable as a reference for studies of protein hydration in solution. In this paper, solvent suppression schemes recently developed for observation of hydration water resonances were used to observe hydroxyl protons of serine, threonine and tyrosine in aqueous solutions of small model peptides and the protein basic pancreatic trypsin inhibitor (BPTI). The chemical shifts of the hydroxyl protons of serine and threonine were found to be between 5.4 and 6.2 ppm, with random-coil shifts at 4°C of 5.92 ppm and 5.88 ppm, respectively, and those of tyrosine between 9.6 and 10.1 ppm, with a random-coil shift of 9.78 ppm. Since these spectral regions are virtually free of other polypeptide1H NMR signals, cross peaks with the hydroxyl protons are usually well separated even in homonuclear two-dimensional1H NMR spectra. To illustrate the practical use of hydroxyl proton NMR in polypeptides, the conformations of the side-chain hydroxyl groups in BPTI were characterized by measurements of nuclear Overhauser effects and scalar coupling constants involving the hydroxyl protons. In addition, hydroxyl proton exchange rates were measured as a function of pH, where simple first-order rate processes were observed for both acid- and base-catalysed exchange of all but one of the hydroxyl-bearing residues in BPTI. For the conformations of the individual Ser, Thr and Tyr side chains characterized in the solution structure with the use of hydroxyl proton NMR, both exact coincidence and significant differences relative to the corresponding BPTI crystal structure data were observed.[/p]  相似文献   

12.
The compounds 3-hydroxy-4-nitrobenzaldehyde and 3-hydroxy-4-nitrobenzyl alcohol are introduced as new chromophoric substrates for probing the catalytic mechanism of horse liver alcohol dehydrogenase (LADH). Ionization of the phenolic hydroxyl group shifts the spectrum of the aldehyde from 360 to 433 nm (pKa = 6.0), whereas the spectrum of the alcohol shifts from 350 to 417 nm (pKa = 6.9). Rapid-scanning, stopped-flow (RSSF) studies at alkaline pH show that the LADH-catalyzed interconversion of these compounds occurs via the formation of an enzyme-bound intermediate with a blue-shifted spectrum. When reaction is limited to a single turnover of enzyme sites, the formation and decay of the intermediate when aldehyde reacts with enzyme-bound reduced nicotinamide adenine dinucleotide E(NADH) are characterized by two relaxations (lambda f approximately equal to 3 lambda s). Detailed stopped-flow kinetic studies were carried out to investigate the disappearance of aldehyde and NADH, the formation and decay of the intermediate, the displacement of Auramine O by substrate, and 2H kinetic isotope effects. It was found that NADH oxidation takes place at the rate of the slower relaxation (lambda s); when NADD is substituted for NADH, lambda s is subject to a small primary isotope effect (lambda Hs/lambda Ds = 2.0); and the events that occur in lambda s precede lambda f. These findings identify the intermediate as a ternary complex containing bound oxidized nicotinamide adenine dinucleotide (NAD+) and some form of 3-hydroxy-4-nitrobenzyl alcohol. The blue-shifted spectrum of the intermediate strongly implies a structure wherein the phenolic hydroxyl is neutral. When constrained to a mechanism that assumes only the neutral phenolic form of the substrate binds and reacts and that the intermediate is an E(NAD+, product) complex, computer simulations yield RSSF and single-wavelength time courses that are qualitatively and semiquantitatively consistent with the experimental data. We conclude that the LADH substrate site can be divided into two subsites: a highly polar, electropositive subsite in the vicinity of the active-site zinc and, just a few angstroms away, a rather nonpolar region. The polar subsite promotes formation of the two interconverting reactive ternary complexes. The nonpolar region is the binding site for the hydrocarbon-like side chains of substrates and in the case of 3-hydroxy-4-nitrobenzaldehyde conveys specificity for the neutral form of the phenolic group.  相似文献   

13.
Neidhart D  Wei Y  Cassidy C  Lin J  Cleland WW  Frey PA 《Biochemistry》2001,40(8):2439-2447
The structures of the hemiketal adducts of Ser 195 in chymotrypsin with N-acetyl-L-leucyl-L-phenylalanyl trifluoromethyl ketone (AcLF-CF3) and N-acetyl-L-phenylalanyl trifluoromethyl ketone (AcF-CF3) were determined to 1.4-1.5 A by X-ray crystallography. The structures confirm those previously reported at 1.8-2.1 A [Brady, K., Wei, A., Ringe, D., and Abeles, R. H. (1990) Biochemistry 29, 7600-7607]. The 2.6 A spacings between Ndelta1 of His 57 and Odelta1 of Asp 102 are confirmed at 1.3 A resolution, consistent with the low-barrier hydrogen bonds (LBHBs) between His 57 and Asp 102 postulated on the basis of spectroscopy and deuterium isotope effects. The X-ray crystal structure of the hemiacetal adduct between Ser 195 of chymotrypsin and N-acetyl-L-leucyl-L-phenylalanal (AcLF-CHO) has also been determined at pH 7.0. The structure is similar to the AcLF-CF3 adduct, except for the presence of two epimeric adducts in the R- and S-configurations at the hemiacetal carbons. In the (R)-hemiacetal, oxygen is hydrogen bonded to His 57, not the oxyanion site. On the basis of the downfield 1H NMR spectrum in solution, His 57 is not protonated at Nepsilon2, and there is no LBHB at pH >7.0. Because addition of AcLF-CHO to chymotrypsin neither releases nor takes up a proton from solution, it is concluded that the hemiacetal oxygen of the chymotrypsin-AcLF-CHO complex is a hydroxyl group and not attracted to the oxyanion site. The protonation states of the hemiacetal and His 57 are explained by the high basicity of the hemiacetal oxygen (pK(a) > 13.5) relative to that of His 57. The 13C NMR signal for the adduct of AcLF-13CHO with chymotrypsin is consistent with a neutral hemiacetal between pH 7 and 13. At pH <7.0, His 57 in the AcLF-CHO-hemiacetal complex of chymotrypsin undergoes protonation at Nepsilon2 of His 57, leading to a transition of the 15.1 ppm downfield signal to 17.8 ppm. The pK(a)s in the active sites of the AcLF-CF3 and AcLF-CHO adducts suggest an energy barrier of 6-7 kcal x mol(-1) against ionizations that change the electrostatic charge at the active site. However, ionizations of neutral His 57 in the AcLF-CHO-chymotrypsin adduct, or in free chymotrypsin, proceed with no apparent barrier. Protonation of His 57 is accompanied by LBHB formation, suggesting that stabilization by the LBHB overcomes the barrier to ionization. On the basis of the hydration constant for AcLF-13CHO and its inhibition constant, its K(d) is 16 microM, 8000-fold larger than the comparable value for AcLF-CF3.  相似文献   

14.
We report the X-ray analysis at 2.0 A resolution for crystals of the aspartic proteinase endothiapepsin (EC 3.4.23.6) complexed with a potent difluorostatone-containing tripeptide renin inhibitor (CP-81,282). The scissile bond surrogate, an electrophilic ketone, is hydrated in the complex. The pro-(R) (statine-like) hydroxyl of the tetrahedral carbonyl hydrate is hydrogen-bonded to both active-site aspartates 32 and 215 in the position occupied by a water in the native enzyme. The second hydroxyl oxygen of the hydrate is hydrogen-bonded only to the outer oxygen of Asp 32. These experimental data provide a basis for a model of the tetrahedral intermediate in aspartic proteinase-mediated cleavage of the amide bond. This indicates a mechanism in which Asp 32 is the proton donor and Asp 215 carboxylate polarizes a bound water for nucleophilic attack. The mechanism involves a carboxylate (Asp 32) that is stabilized by extensive hydrogen bonding, rather than an oxyanion derivative of the peptide as in serine proteinase catalysis.  相似文献   

15.
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   

16.
A range of substrate-derived chloromethane inhibitors have been synthesized which have one to four amino acid residues. These have been used to inhibit both subtilisin and chymotrypsin. Using 13C NMR, we have shown that all except one of these inhibitors forms a tetrahedral adduct with chymotrypsin, subtilisin, and trypsin. From the pH-dependent changes in the chemical shift of the hemiketal carbon of the tetrahedral adduct, we are able to determine the oxyanion pKa in the different inhibitor derivatives. Our results suggest that in both the subtilisin and chymotrypsin chloromethane derivatives the oxyanion pKa is largely determined by the type of amino acid residue occupying the S1, subsite while binding in the S2-S4 subsites only has minor effects on oxyanion pKa values. Using free energy relationships, we determine that the different R groups of the amino acid residues binding in the S1 subsite only have minor effects on the oxyanion pKa values. We propose that the lower polarity of the chymotrypsin active site relative to that of the subtilisin active site explains why the oxyanion pKa is higher and more sensitive to the type of chloromethane inhibitor used in the chymotrypsin derivatives than in the subtilisin derivatives.  相似文献   

17.
X-ray crystallographic data to 2.57 A resolution (1 A = 0.1 nm) have been measured for the complex of a peptidyl trifluoromethylketone inhibitor with porcine pancreatic elastase (PPE); R = 0.14. The inhibitor forms a stable complex with the enzyme by means of a covalent attachment to active site Ser195O gamma, resulting in a hemiketal moiety with tetrahedral geometry. The tripeptide protion binds as an antiparallel beta-sheet, with four hydrogen bonds augmenting the active-site covalent linkage, Ki = 9.5 microM. His57 exhibits a bifurcated H-bond to both Ser195O gamma and an F atom of the inhibitor. This study is one of a series which explores the binding geometry of a variety of small substrates and inhibitors to PPE. This peptidyl-PPE complex affords insight into the binding geometry of a novel trifluoromethylketone moiety to a serine proteinase.  相似文献   

18.
K N Allen  R H Abeles 《Biochemistry》1989,28(21):8466-8473
A series of trifluoromethyl ketones that reversibly inhibit acetylcholinesterase and pseudocholinesterase were synthesized. By analogy to chymotrypsin and on the basis of data reported here, we propose that the active-site serine adds to the ketone to form an ionized hemiketal. The compound (5,5,5-trifluoro-4-oxopentyl)trimethylammonium bicarbonate (1) inhibits acetylcholinesterase with Ki = 0.06 X 10(-9)M and pseudocholinesterase with Ki = 70 X 10(-9)M. Replacement of the nitrogen of 1 by carbon (compound 2) increases Ki for 1 200-fold for acetylcholinesterase but does not significantly alter Ki for pseudocholinesterase. The Ki for the methyl ketone corresponding to 2 is 2 X 10(-4)M for both enzymes, as compared with 12 X 10(-9)M for the trifluoromethyl ketone (acetylcholinesterase). For both enzymes, a linear decrease in log Ki with decreasing pK of the inhibitor hydrate was observed with ketones containing from 0 to 3 fluorines. We attribute this effect to the stabilization of the hemiketal oxyanion. The reduction of the pK of the hemiketal by the trifluoromethyl group is an important contributing factor to the low Ki of trifluoromethyl ketones. The inhibition of acetylcholinesterase by tetramethylammonium chloride and trifluoroacetone was compared to the inhibition by 1, which is a composite of the two smaller inhibitors. The entropic advantage of combining the smaller inhibitors into one molecule is 1.1 X 10(3)M. Inhibitors with Ki less than or equal to 70 X 10(-9) M are slow binding (Morrison, 1982; Morrison & Walsh, 1988). The kinetic data do not require formation of a noncovalent complex prior to formation of the ketal, although such a complex(es) cannot be excluded.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Kinetics of inhibition have been determined for the interaction of human leukocyte elastase (HLE) with two series of peptide trifluoromethyl ketones (TFMKs): X-Val-CF3,X-Pro-Val-CF3,X-Val-Pro-Val-CF3, and X-Lys(Z)-Val-Pro-Val-CF3, where X is MeOSuc or Z. These compounds are "slow-binding" inhibitors of HLE and, thus, allow the determination of Ki, the dissociation constant for the stable complex of inhibitor and enzyme, as well as kon and koff, the rate constants for formation and decomposition of this complex. Maximal potency is reached with Z-Lys(Z)-Val-Pro-Val-CF3, which displays a Ki less than 0.1 nM. Upon binding to HLE, these compounds undergo addition by the hydroxyl of the active site serine to form a hemiketal. The evidence supporting a hemiketal intermediate includes Ki values of 1.6 and 80,000 nM for Z-Val-Pro-Val-CF3 and its alcohol analogue, linear free energy correlations between inhibitory potency and catalytic efficiency for structurally related TFMKs and substrates, and the pH dependence of kon for the inhibition of HLE by Z-Val-Pro-Val-CF3, which is sigmoidal and displays a pKa of 6.9. Hemiketal formation is probably not rate limiting, however. Kinetic solvent isotope effects of unity suggest that kon cannot be rate limited by a reaction step, like hemiketal formation, that is subject to protolytic catalysis. A general mechanism that is consistent with these results is one in which formation of the hemiketal is rapid and is followed or preceded by a slow step that rate limits kon.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
A R Rendina  W W Cleland 《Biochemistry》1984,23(22):5157-5168
Aldehyde analogues of the normal alcohol substrates induce ATPase activities by glycerokinase (D-glyceraldehyde), fructose-6-phosphate kinase (2,5-anhydromannose 6-phosphate), fructokinase (2,5-anhydromannose or 2,5-anhydrotalose), hexokinase (D-gluco-hexodialdose), choline kinase (betaine aldehyde), and pyruvate kinase (glyoxylate). Since purified deuterated aldehydes give V and V/K isotope effects near 1.0 for glycerokinase, fructokinase with 2,5-anhydro[1-2H]talose, hexokinase, choline kinase, and pyruvate kinase, the hydrates of these almost fully hydrated aldehydes are the activators of the ATPase reactions. Fructose-6-phosphate kinase and fructokinase with 2,5-anhydro[1-2H]mannose show V/K deuterium isotope effects of 1.10 and 1.22, respectively, suggesting either that both hydrate and free aldehyde may be activators (predicted values are 1.37 if only the free aldehyde activates the ATPase) or, more likely, that the phosphorylated hydrate breaks down in a rate-limiting step on the enzyme while MgADP is still present and the back-reaction to yield free hydrate in solution is still possible. 18O was transferred from the aldehyde hydrate to phosphate during the ATPase reactions of glycerokinase, fructose-6-phosphate kinase, fructokinase, and hexokinase but not with choline kinase or pyruvate kinase. Thus, direct phosphorylation of the hydrates by the first four enzymes gives the phosphate adduct of the aldehyde, which decomposes nonenzymatically, while with choline kinase and pyruvate kinase the hydrates induce transfer to water (metal-bound hydroxide or water with pyruvate kinase on the basis of pH profiles). Observation of a lag in the release of phosphate from the glycerokinase ATPase reaction at 15 degrees C supports the existence of a phosphorylated hydrate intermediate with a rate constant for breakdown of 0.035-0.043 s-1 at this temperature. Kinases that phosphorylate creatine, 3-phosphoglycerate, and acetate did not exhibit ATPase activities in the presence of keto or aldehyde analogues (N-methylhydantoic acid, D-glyceraldehyde 3-phosphate, and acetaldehyde, respectively), possibly because of the absence of an acid-base catalytic group in the latter two cases. These analogues were competitive inhibitors vs. the normal substrates, and in the latter case, the hydrate of acetaldehyde was shown to be the inhibitory species on the basis of the deuterium isotope effect on the inhibition constant.  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