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 31 毫秒
1.
Site-directed mutagenesis can be employed to alter activity critical residues in proteins which are susceptible to chemical oxidation. Previous studies have implicated methionine 222 as a primary site for oxidative inactivation of subtilisin (Stauffer, C. E., and Etson, D. (1969) J. Biol. Chem. 244, 5333-5338). Because of uncertainties in predicting which amino acid would be the optimal substitute for methionine 222, we prepared all 19 amino acid substitutions at this site in the cloned subtilisin gene using a cassette mutagenesis method (Wells, J. A., Vasser, M., and Powers, D. P. (1985) Gene (Amst.), in press). Mutant enzymes were expressed in Bacillus subtilis and were found to vary widely in specific activity. Mutants containing nonoxidizable amino acids (i.e. Ser, Ala, and Leu) were resistant to inactivation by 1 M H2O2, whereas methionine and cysteine enzymes were rapidly inactivated. These studies demonstrate the feasibility of improving oxidative stability in proteins by site-directed mutagenesis.  相似文献   

2.
The refined crystal structure of subtilisin Carlsberg at 2.5 A resolution   总被引:4,自引:0,他引:4  
We report here the X-ray crystal structure of native subtilisin Carlsberg, solved at 2.5 A resolution by molecular replacement and refined by restrained least squares to a crystallographic residual (Formula see text): of 0.206. we compare this structure to the crystal structure of subtilisin BPN'. We find that, despite 82 amino acid substitutions and one deletion in subtilisin Carlsberg relative to subtilisin BPN', the structures of these enzymes are remarkably similar. We calculate an r.m.s. difference between equivalent alpha-carbon positions in subtilisin Carlsberg and subtilisin BPN' of only 0.55 A. This confirms previous reports of extensive structural homology between these two subtilisins based on X-ray crystal structures of the complex of eglin-c with subtilisin Carlsberg [McPhalen, C.A., Schnebli, H.P. and James, M.N.G. (1985) FEBS Lett., 188, 55; Bode, W., Papamokos, E. and Musil, D. (1987) Eur. J. Biochem., 166, 673-692]. In addition, we find that the native active sites of subtilisins Carlsberg and BPN' are virtually identical. While conservative substitutions at residues 217 and 156 may have subtle effects on the environments of substrate-binding sites S1' and S1 respectively, we find no obvious structural correlate for reports that subtilisins Carlsberg and BPN' differ in their recognition of model substrates. In particular, we find no evidence that the hydrophobic binding pocket S1 in subtilisin Carlsberg is 'deeper', 'narrower' or 'less polar' than the corresponding binding site in subtilisin BPN'.  相似文献   

3.
The crystal structure of subtilisin BPN' complexed with a proteinaceous inhibitor SSI (Streptomyces subtilisin inhibitor) was refined at 1.8 A resolution to an R-factor of 0.177 with a root-mean-square deviation from ideal bond lengths of 0.014 A. The work finally established that the SSI-subtilisin complex is a Michaelis complex with a distance between the O gamma of active Ser221 and the carbonyl carbon of the scissile peptide bond being an intermediate value between a covalent bond and a van der Waals' contact, 2.7 A. This feature, as well as the geometry of the catalytic triad and the oxyanion hole, is coincident with that found in other highly refined crystal structures of the complex of subtilisin Novo, subtilisin Carlsberg, bovine trypsin or Streptomyces griseus protease B with their proteinaceous inhibitors. The enzyme-inhibitor beta-sheet interaction is composed of two separate parts: that between the P1-P3 residues of SSI and the 125-127 chain segment (the "S1-3 site") of subtilisin and that between the P4-P6 residues of SSI and th 102-104 chain segment (the "S4-6 site") of subtilisin. The latter beta-interaction is unique to subtilisin. In contrast, the beta-sheet interaction previously found in the complex of subtilisin Novo and chymotrypsin inhibitor 2 or in the complex of subtilisin Carlsberg and Eglin C is distinct from the present complex in that the two types of beta-interactions are not separate. As for the flexibility of the molecules comprising the present complex, the following observations were made by comparing the B-factors for free and complexed SSI and comparing those for free and complexed subtilisin BPN'. The rigidification of the component molecules upon complex formation occurs in a very localized region: in SSI, the "primary" and "secondary" contact regions and the flanking region; in subtilisin BPN', the S1-3 and S4-6 sites and the flanking region.  相似文献   

4.
The three methionine residues of subtilisin DY were specifically modified into methionine sulfoxide using increasing amounts of chloramine T. By means of subsequent treatment with cyanogen bromide, gel chromatography, Edman degradation of the obtained peptides and the known structure of subtilisin DY it was established that Met222 is exposed to the surrounding solution, Met124 is partially exposed and Met199 is buried. The data obtained were confirmed on a computer graphics space movable model of subtilisin Carlsberg where Met222 was seen to be on the surface of the molecule and Met199 shielded by Tyr262, Ala179 and Leu196. Upon oxidation of Met222 of subtilisin DY by chloramine T, 25% of its caseinolytic activity was lost. This can be explained by the immediate adjacency to the active-site Ser221. An additional 5% loss of activity was observed at each subsequent methionine modification.  相似文献   

5.
Triclinic crystals of the complex formed by eglin with subtilisin Carlsberg were analyzed by X-ray diffraction. The crystal and molecular structure of this complex was determined with data that extended to 0.12-nm resolution by a combination of Patterson search methods and isomorphous replacement techniques. Its structure was refined to a crystallographic R value of 0.178 (1.0-0.12 nm) using an energy-restraint least-squares procedure. The complete subtilisin molecule could be traced without ambiguity in the refined electron density. The eglin component, from which an amino-terminal segment is cleaved off, is only defined from Lys8I (i.e. the lysine residue 8 of the inhibitor) onwards. Per unit cell, 436 fixed solvent molecules and 2 calcium ions were located. In spite of 84 amino acid replacements and one deletion, subtilisin Carlsberg exhibits a very similar polypeptide fold to subtilisin BPN'. The root-mean-square deviations of all alpha-carbon atoms (excluding those at the deletion site) from models of subtilisin BPN' [Alden, R. A., Birktoft, J. J., Kraut, J., Robertus, J. D. & Wright, C. S. (1971) Biochem. Biophys. Res. Commun. 45, 337-344] and subtilisin Novo [Drenth, J., Hol, W. G. J., Jansonius, J. N. & Kockoek, R. (1972) Eur. J. Biochem. 25, 177-181] are 0.077 nm and 0.103 nm. Most of these deviations result from global shifts rather than changes of the local geometry. The single-residue deletion at position 56 affects only the surrounding conformation. Two sites of high electron density and close distances to surrounding oxygen ligands have been found in the Carlsberg enzyme which are probably occupied by calcium ions. Eglin consists of a twisted four-stranded beta-sheet flanked by an alpha-helix and by an exposed proteinase binding loop on opposite sides. Around the reactive site, Leu45I-Asp46I, this loop is mainly stabilized by electrostatic/hydrogen bond interactions with the side chains of two arginine residues which project from the hydrophobic core [Bode, W., Papamokos, E., Musil, D., Seemüller, W. & Fritz, H. (1986) EMBO J. 5, 813-818]. The reactive site loop conformation resembles that found in other 'small' proteinase inhibitors. The scissile peptide bond is not cleaved but its carbonyl group is slightly distorted from planar geometry. Most of the intermolecular contacts are contributed by the nine residues of the reactive-site loop Gly40I-Arg48I.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
We have studied the structures of adducts formed between subtilisin BPN' and both benzeneboronic acid and 2-phenylethaneboronic acid by x-ray diffraction techniques. Electron density and difference maps at 2.5 A resolution were computed with phases calculated from a partially refined structure of the native enzyme (R = 0.23 at 2.0 A). Both adducts contain a covalent bond between Ogamma of the catalytic Ser-221 and the inhibitor boron atom. The boron atom is coordinated tetrahedrally, with one of the two additional boronic acid oxygen atoms lying in the "oxyanion hole" and the other at the leaving group site identified in previous studies (ROBERTUS, J.D., Kraut, J. ALDEN, R.A., and BIRKTOFT, J.J. (1972) Biochemistry 11, 4293-4303). Moreover, the previously postulated structure of the tetrahedral intermediate for substrate hydrolysis is isosteric with these boronic acid adducts, which can therefore be considered good models for the transition state complex (KOEHLER, K.K., and LIENHARD, G.E. (1972) Biochemistry 10, 2477-2483). These observations further support the suggestion that an important contribution to stabilization of this transition state complex, relative to both the Michaelis complex and the acyl intermediate, occurs as a consequence of hydrogen bond donation to the substrate carbonyl oxygen atom from the side chain amido group of Asn-155 and from the backbone amido group of Ser-221.  相似文献   

7.
Benzeneboronic acid, a transition-state analog for serine proteases, binds to the catalytic center of subtilisin BPN'. The binding mechanism is so-called two-step mechanism; the initial fast association followed by a slow unimolecular process (Nakatani, H., Uehara, Y. and Hiromi, K. (1975) J. Biochem. (Tokyo) 77, 615--616), E + S fast equilibrium ES slow equilibrium ES (E = subtilisin, S = benzenebroonic acid). The structure of the transient complex (ES) at the initial association process was manifested by the substituent effect of benzeneboronic acid on the rate parameters in the elementary processes. The study by the temperature-junp and stopped-flow methods showed that the boron atom in benzeneboronic acid strongly interacts with a nucleophilic site, probably, O gamma of Ser-221 or imidazole of His-64 at the catalytic center, already at the initial fast association.  相似文献   

8.
Subtilisin-type proteinases (thermitase, subtilisin Carlsberg, alkaline proteinase ZIMET 10911, proteinase K) are partially inactivated by hydrogen peroxide in the alkaline pH range only in the presence of boric acid or phenylboronic acid. A model is presented to describe the inactivation mechanism. Both boric acid and perboric acid existing in equilibrium in the presence of hydrogen peroxide bind competitively at the active site of the enzyme. The inactivation, which is known to be caused by sulfoxide formation from the methionine residue in the active site (Stauffer, C.E. and Etson, D. (1969) J. Biol. Chem. 244, 5333-5338), is due to the enzyme-bound perboric acid species. The dissociation constants for the boric acid-thermitase and perboric acid-thermitase complexes are 36 +/- 7 and 4 +/- 1 mM, respectively. The first-order rate constant of inactivation is k = 0.63 +/- 0.14 min-1. The same mechanism of inactivation holds true for phenylboronic acid in alkaline hydrogen peroxide solutions.  相似文献   

9.
The protein Streptomyces subtilisin inhibitor, SSI, efficiently inhibits a bacterial serine protease, subtilisin BPN'. We recently demonstrated that functional change in SSI was possible simply by replacing the amino acid residue at the reactive P1 site (methionine 73) of SSI. The present paper reports the additional effect of replacing methionine 70 at the P4 site of SSI (Lys73) on inhibitory activity toward two types of serine proteases, trypsin (or lysyl endopeptidase) and subtilisin BPN'. Conversion of methionine 70 at the P4 site of SSI(Lys73) to glycine or alanine resulted in increased inhibitory activity toward trypsin and lysyl endopeptidase, while replacement with phenylalanine weakened the inhibitory activity toward trypsin. This suggests that steric hindrance at the P4 site of SSI(Lys73) is an obstacle for its binding with trypsin. In contrast, the same P4 replacements had hardly any effect on inhibitory activity toward subtilisin BPN'. Thus the subsite structure of subtilisin BPN' is tolerant to these replacements. This contrast in the effect of P4 substitution might be due to the differences in the S4 subsite structures between the trypsin-like and the subtilisin-like proteases. These findings demonstrate the importance of considering structural complementarity, not only at the main reactive site but also at subsites of a protease, when designing stronger inhibitors.  相似文献   

10.
The conformational stabilization of proteins by sucrose has been previously attributed to a preferential exclusion mechanism. The present study links this mechanism to stability against a chemical degradation pathway for subtilisin. Oxidation of a methionine residue adjacent to the active site to the sulfoxide form compromises subtilisin's enzymatic activity. In the presence of hydrogen peroxide and borate buffer, a borate-hydrogen peroxide complex binds to subtilisin's active site prior to the formation of methionine sulfoxide. Sucrose decreases the oxidation rate by limiting the accessibility of the complex to the methionine at the partially buried active site. The stabilization mechanism of sucrose is based on shifting the equilibrium of transiently expanding native conformations of subtilisin to favor the most compact states. Enzymatic parameter determination (kcat, KM) and hydrogen-deuterium exchange measurements confirm the limited conformational mobility of the enzyme in the presence of sucrose. Further support for limited mobility as the cause of oxidation inhibition by sucrose comes from the findings that neither viscosity nor possible interactions of sucrose with hydrogen peroxide, hydroxyl radicals, or borate can adequately explain the inhibition. The volume exclusion of sucrose from subtilisin is used to estimate the extent by which the native state of subtilisin must expand in solution to allow oxidation. The surface area of the oxidation-competent state is ca. 3.9% greater than that of the native state.  相似文献   

11.
Subtilisin-type proteinases (thermitase, subtilisin Carlsberg, alkaline proteinase ZIMET 10911, proteinase K) are partially inactivated by hydrogen peroxide in the alkaline pH range only in the presence of boric acid or phenylboronic acid. A model is presented to describe the inactivation mechanism. Both boric acid and perboric acid existing in equilibrium in the presence of hydrogen peroxide bind competitively at the active site of the enzyme. The inactivation, which is known to be caused by sulfoxide formation from the methionine residue in the active site (Stauffer C.E. and Etson D. (1969) J. Biol. Chem. 244, 5333–5338), is due to the enzyme-bound perboric acid species. The dissociation constants for the boric acid-thermitase and perboric acid-thermitase complexes are 36 ± 7 and 4 ± 1 mM, respectively. The first-order rate constant of inactivation is k = 0.63 ± 0.14 min−1. The same mechanism of inactivation holds true for phenylboronic acid in alkaline hydrogen peroxide solutions.  相似文献   

12.
The crystal structure of an oxidatively stable subtilisin-like alkaline serine protease, KP-43 from Bacillus sp. KSM-KP43, with a C-terminal extension domain, was determined by the multiple isomorphous replacements method with anomalous scattering. The native form was refined to a crystallographic R factor of 0.134 (Rfree of 0.169) at 1.30-A resolution. KP-43 consists of two domains, a subtilisin-like alpha/beta domain and a C-terminal jelly roll beta-barrel domain. The topological architecture of the molecule is similar to that of kexin and furin, which belong to the subtilisin-like proprotein convertases, whereas the amino acid sequence and the binding orientation of the C-terminal beta-barrel domain both differ in each case. Since the C-terminal domains of subtilisin-like proprotein convertases are essential for folding themselves, the domain of KP-43 is also thought to play such a role. KP-43 is known to be an oxidation-resistant protease among the general subtilisin-like proteases. To investigate how KP-43 resists oxidizing reagents, the structure of oxidized KP-43 was also determined and refined to a crystallographic R factor of 0.142 (Rfree of 0.212) at 1.73-A resolution. The structure analysis revealed that Met-256, adjacent to catalytic Ser-255, was oxidized similarly to an equivalent residue in subtilisin BPN'. Although KP-43, as well as proteinase K and subtilisin Carlsberg, lose their hydrolyzing activity against synthetic peptides after oxidation treatment, all of them retain 70-80% activity against proteinaceous substrates. These results, as well as the beta-casein digestion pattern analysis, have indicated that the oxidation of the methionine adjacent to the catalytic serine is not a dominant modification but might alter the substrate specificities.  相似文献   

13.
The structural gene for a subtilisin J from Bacillus stearothermophilus NCIMB10278 was cloned in Bacillus subtilis using pZ124 as a vector, and its nucleotide sequence was determined. The nucleotide sequence revealed only one large open reading frame, composed of 1,143 base pairs and 381 amino acid residues. A Shine-Dalgarno sequence was found 8 bp upstream from the translation start site (GTG). The deduced amino acid sequence revealed an N-terminal signal peptide and pro-peptide of 106 residues followed by the mature protein comprised of 275 residues. The productivity of subtilisin in the culture broth of the Bacillus subtilis was about 46-fold higher than that of the Bacillus stearothermophilus. The amino acid sequence of the extracellular alkaline protease subtilisin J is highly homologous to that of subtilisin E and it shows 69% identity with subtilisin Carlsberg, 89% with subtilisin BPN' and 70% with subtilisin DY. Some properties of the subtilisin J that had been purified from the Bacillus subtilis were examined. The subtilisin J has alkaline pH characteristics and a molecular weight of 27,500. It retains about 50% of its activity even after treatment at 60 degrees C for 30 min in the presence of 2 mM calcium chloride.  相似文献   

14.
Functional interaction among catalytic residues in subtilisin BPN'   总被引:4,自引:0,他引:4  
P Carter  J A Wells 《Proteins》1990,7(4):335-342
Variants of the serine protease, subtilisin BPN', in which the catalytic triad residues (Ser-221, His-64, and Asp-32) are replaced singly or in combination by alanine retain activities with the substrate N-succinyl-L-Ala-L-Ala-L-Pro-L-Phe-p-nitroanilide (sAAPF-pna) that are at least 10(3) to 10(4) above the non-enzymatic rate [Carter, P., Wells, J.A. Nature (London) 322:564-568, 1988]. A possible source of the residual activity was the hydrogen bond with the N delta 2 of Asn-155 that helps to stabilize the oxyanion generated in the tetrahedral transition state during amide bond hydrolysis by the wild-type enzyme. Replacing Asn-155 by Gly (N155G) lowers the turnover number (kcat) for sAAPF-pna by 150-fold with virtually no change in the Michaelis constant (KM). However, upon combining the N155G and S221A mutations to give N155G:S221A, kcat is actually 5-fold greater than for the S221A enzyme. Thus, the catalytic role of Asn-155 is dependent upon the presence of Ser-221. The residual activity of the N155G:S221A enzyme (approximately 10(4)-fold above the uncatalyzed rate) is not an artifact because it can be completely inhibited by the third domain of the turkey ovomucoid inhibitor (OMTKY3), which forms a strong 1:1 complex with the active site. The mutations N155G and S221A individually weaken the interaction between subtilisin and OMTKY3 by 1.8 and 2.0 kcal/mol, respectively, and in combination by 2.1 kcal/mol. This is consistent with disruption of stabilizing interactions around the reactive site carbonyl of the OMTKY3 inhibitor.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Crystal structure of thermitase at 1.4 A resolution   总被引:4,自引:0,他引:4  
The crystal structure of thermitase, a subtilisin-type serine proteinase from Thermoactinomyces vulgaris, was determined by X-ray diffraction at 1.4 A resolution. The structure was solved by a combination of molecular and isomorphous replacement. The starting model was that of subtilisin BPN' from the Protein Data Bank, determined at 2.5 A resolution. The high-resolution refinement was based on data collected using synchrotron radiation with a Fuji image plate as detector. The model of thermitase refined to a conventional R factor of 14.9% and contains 1997 protein atoms, 182 water molecules and two Ca ions. The tertiary structure of thermitase is similar to that of the other subtilisins although there are some significant differences in detail. Comparison with subtilisin BPN' revealed two major structural differences. The N-terminal region in thermitase, which is absent in subtilisin BPN', forms a number of contacts with the tight Ca2+ binding site and indeed provides the very tight binding of the Ca ion. In thermitase the loop of residues 60 to 65 forms an additional (10) beta-strand of the central beta-sheet and the second Ca2+ binding site that has no equivalent in the subtilisin BPN' structure. The observed differences in the Ca2+ binding and the increased number of ionic and aromatic interactions in thermitase are likely sources of the enhanced stability of thermitase.  相似文献   

16.
Yang Y  Jiang L  Zhu L  Wu Y  Yang S 《Journal of biotechnology》2000,81(2-3):113-118
A remarkable thermal stable and oxidation-resistant mutant was obtained using the random mutagenesis PCR technique on the mutant M222A gene of subtilisin E. Sequencing analysis revealed an A was replaced by G at nucleotide 671 of the subtilisin E gene, converting the asparagine codon (AAT) to serine codon (AGT) at position 118. The half-life of M222A/N118S enzyme activity, when heated at 65 degrees C, was approximately 80 min while the half-life of M222A and wild-type subtilisin E were 13 min and 15 min, respectively. This suggested the stability of the M222A/N118S mutant was five times greater than that of the wild-type enzyme. The mutant was also as oxidation resistant as the mutant M222A of subtilisin E. These results indicated the M222A/N118S mutant is both an oxidation-resistant and a heat-stable variant of subtilisin E.  相似文献   

17.
The crystal structures of two thermally stabilized subtilisin BPN' variants, S63 and S88, are reported here at 1.8 and 1.9 A resolution, respectively. The micromolar affinity calcium binding site (site A) has been deleted (Delta75-83) in these variants, enabling the activity and thermostability measurements in chelating conditions. Each of the variants includes mutations known previously to increase the thermostability of calcium-independent subtilisin in addition to new stabilizing mutations. S63 has eight amino acid replacements: D41A, M50F, A73L, Q206W, Y217K, N218S, S221C, and Q271E. S63 has 75-fold greater stability than wild type subtilisin in chelating conditions (10 mm EDTA). The other variant, S88, has ten site-specific changes: Q2K, S3C, P5S, K43N, M50F, A73L, Q206C, Y217K, N218S, and Q271E. The two new cysteines form a disulfide bond, and S88 has 1000 times greater stability than wild type subtilisin in chelating conditions. Comparisons of the two new crystal structures (S63 in space group P2(1) with A cell constants 41.2, 78.1, 36.7, and beta = 114.6 degrees and S88 in space group P2(1)2(1)2(1) with cell constants 54.2, 60.4, and 82.7) with previous structures of subtilisin BPN' reveal that the principal changes are in the N-terminal region. The structural bases of the stabilization effects of the new mutations Q2K, S3C, P5S, D41A, Q206C, and Q206W are generally apparent. The effects are attributed to the new disulfide cross-link and to improved hydrophobic packing, new hydrogen bonds, and other rearrangements in the N-terminal region.  相似文献   

18.
Protein engineering techniques were used to construct a derivative of the serine protease subtilisin that ligates peptides efficiently in water. The subtilisin double mutant in which the catalytic Ser221 was converted to Cys (S221C) and Pro225 converted to Ala (P225A) has 10-fold higher peptide ligase activity and at least 100-fold lower amidase activity than the singly mutated thiolsubtilisin (S221C) that was previously shown to have some peptide ligase activity [Nakatsuka, T., Sasaki, T., & Kaiser, E.T. (1987) J. Am. Chem. Soc. 109, 3808-3810]. A 1.5-A X-ray crystal structure of an oxidized derivative of the double mutant (S221C/P225A) supports the protein design strategy in showing that the P225A mutation partly relieves the steric crowding expected from the S221C substitution, thus accounting for its improved catalytic efficiency. Stable and synthetically reasonable alkyl ester peptide substrates were prepared that rapidly acylate the S221C/P225A enzyme, and aminolysis of the resulting thioacyl-enzyme intermediate by various peptides is strongly preferred over hydrolysis. The efficiency of aminolysis is relatively insensitive to the sequence of the first two residues in the acyl acceptor peptide whose alpha-amino group attacks the thioacyl-enzyme. To obtain greater flexibility in the choice of coupling sites, a set of three additional peptide ligases were engineered by introducing mutations into the parent ligase (S221C/P225A) that were previously shown to change the specificity of subtilisin for the residue nearest the acyl bond (the P1 residue). The specificity properties of the parent ligase and derivatives of it paralleled those of wild type and corresponding specificity variants.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
P Carter  L Abrahmsén  J A Wells 《Biochemistry》1991,30(25):6142-6148
A mutant of the serine protease, subtilisin BPN', in which the catalytic His64 is replaced by Ala (H64A), is very specific for substrates containing a histidine, presumably by the substrate-bound histidine assisting in catalysis [Carter, P., & Wells, J.A. (1987) Science (Washington, D.C.) 237, 394-399]. Here we probe the catalytic mechanism of H64A subtilisin for cleaving His and non-His substrates. We show that the ratio of aminolysis to hydrolysis is the same for ester and amide substrates as catalyzed by the H64A subtilisin. This is consistent with formation of a common acyl-enzyme intermediate for H64A subtilisin, analogous to the mechanism of the wild-type enzyme. However, the catalytic efficiencies (kcat/KM) for amidase and esterase activities with His-containing substrates are reduced by 5000-fold and 14-fold, respectively, relative to wild-type subtilisin BPN, suggesting that acylation is more compromised than deacylation in the H64A mutant. High concentrations of imidazole are much less effective than His substrates in promoting hydrolysis by the H64A variant, suggesting that the His residue on the bound (not free) substrate is involved in catalysis. The reduction in catalytic efficiency kcat/KM for hydrolysis of the amide substrate upon replacement of the oxyanion stabilizing asparagine (N155G) is only 7-fold greater for wild-type than H64A subtilisin. In contrast, the reductions in kcat/KM upon replacement of the catalytic serine (S221A) or aspartate (D32A) are about 3000-fold greater for wild-type than H64A subtilisin, suggesting that the functional interactions between the Asp32 and Ser221 with the substrate histidine are more compromised in substrate-assisted catalysis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Engineering subtilisin BPN' for site-specific proteolysis   总被引:6,自引:0,他引:6  
A combination of protein engineering and substrate optimization was used to create variants of the serine protease, subtilisin BPN', which efficiently and specifically cleave a designed target sequence in a fusion protein. The broad substrate specificity of wild-type subtilisin BPN' is greatly restricted by substitution of the catalytic histidine-containing of the catalytic histidine 64 with alanine (H64A) so that certain histidine-containing substrates are preferentially hydrolysed (Carter, P., Wells, J.A. Science 237:394-399, 1987). The catalytic efficiency, (kcat/Km), of this H64A variant was increased almost 20-fold by judicious choice of substrate and by installing three additional mutations which increase the activity of wild-type subtilisin. The most favorable substrate sequence identified was introduced as a linker in a fusion protein between a synthetic IgG binding domain of Staphylococcus aureus protein A and Escherichia coli alkaline phosphatase. The fusion protein (affinity purified on an IgG column) was cleaved by the prototype H64A enzyme and its improved variant, efficiently and exclusively at the target site, to liberate an alkaline phosphatase product of the expected size and N-terminal sequence. Several features of H64A variants of subtilisin make them attractive for site-specific proteolysis of fusion proteins: they have exquisite substrate specificity on the N-terminal side of the cleavage site and yet are broadly specific on the C-terminal side; they can be produced in large quantities and remain highly active even in the presence of detergents, reductants (modest concentrations), protease inhibitors, at high temperatures, or when specifically immobilized on a solid support.  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