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.
B Ruan  J Hoskins  P N Bryan 《Biochemistry》1999,38(26):8562-8571
In vitro folding of mature subtilisin is extremely slow. The isolated pro-domain greatly accelerates in vitro folding of subtilisin in a bimolecular reaction whose product is a tight complex between folded subtilisin and folded pro-domain. In our studies of subtilisin, we are trying to answer two basic questions: why does subtilisin fold slowly without the pro-domain and what does the pro-domain do to accelerate the folding rate? To address these general questions, we are trying to characterize all the rate constants governing individual steps in the bimolecular folding reaction of pro-domain with subtilisin. Here, we report the results of a series of in vitro folding experiments using an engineered pro-domain mutant which is independently stable (proR9) and two calcium-free subtilisin mutants. The bimolecular folding reaction of subtilisin and proR9 occurs in two steps: an initial binding of proR9 to unfolded subtilisin, followed by isomerization of the initial complex into the native complex. The central findings are as follows. First, the independently stable proR9 folds subtilisin much faster than the predominantly unfolded wild-type pro-domain. Second, at micromolar concentrations of proR9, the subtilisin folding reaction becomes limited by the rate at which prolines in the unfolded state can isomerize to their native conformation. The simpliest mechanism which closely describes the data includes two denatured forms of subtilisin, which form the initial complex with proR9 at the same rate but which isomerize to the fully folded complex at much different rates. In this model, 77% of the subtilisin isomerizes to the native form slowly and the remaining 23% isomerizes more rapidly (1.5 s-1). The slow-folding population may be unfolded subtilisin with the trans form of proline 168, which must isomerize to the cis form during refolding. Third, in the absence of proline isomerization, the rate of subtilisin folding is rapid and at [proR9] 3 s-1. The implications of these results concerning why subtilisin folds slowly without the pro-domain are discussed.  相似文献   

2.
Fourier-transform infrared (FTIR) spectroscopy has been used to quantify the alpha-helix and beta-sheet contents of subtilisin Carlsberg dissolved in several nonaqueous, as well as aqueous, solvents. Independently, the catalytic activity of the enzyme has been measured in the same solvents. While our previous FTIR studies revealed no connection between the secondary structure and enzymatic activity for subtilisin suspended in various organic solvents, a very different situation is observed herein for the dissolved enzyme. Specifically, if either the alpha-helix or beta-sheet content in a given solvent is higher or lower than in water, no appreciable enzymatic catalysis is observed. Conversely, when the secondary structure of subtilisin dissolved in a given nonaqueous solvent is similar to that in water, so is the enzymatic activity. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 56: 485-491, 1997.  相似文献   

3.
We studied a model transesterification reaction catalyzed by subtilisin Carlsberg suspended in carbon dioxide, propane, and mixtures of these solvents under pressure. To account for solvent effects due to differences in water partitioning between the enzyme and the bulk solvents, we measured water sorption isotherms for the enzyme in each solvent. We measured catalytic activity as a function of enzyme hydration and obtained bell-shaped curves with maxima at the same enzyme hydration (12%) in all the solvents. However, the activity maxima were different in all media, being much higher in propane than in either CO(2) or the mixtures with 50 and 10% CO(2). Considerations based on the solvation ability of the solvents did not offer an explanation for the differences in catalytic activity observed. Our results suggest that CO(2) has a direct adverse effect on the catalytic activity of subtilisin. (c) 1996 John Wiley & Sons, Inc.  相似文献   

4.
We studied a model transesterification reaction catalyzed by subtilisin Carlsberg suspended in toluene, n-hexane, diisopropyl ether, and mixtures of these solvents. To account for solvent effects due to differences in water partitioning between the enzyme and the bulk solvents, we measured water sorption isotherms for the enzyme in each solvent. We measured catalytic activity as a function of enzyme hydration and obtained bell-shaped curves with maxima at the same enzyme hydration in all the solvents. However, the activity maxima were different in all the media, being the lowest in toluene. Differences in the partitioning of substrates and product between the bulk solvent phase and the enzyme active site were accounted for but could not explain the lower catalytic activity observed in toluene. The fact that toluene is very similar to one of the substrates suggested the possibility of competitive inhibition by this solvent. We derived a model allowing for differences in solvation of the substrates, by using thermodynamic activities instead of concentrations, as well as for competitive inhibition by toluene. The model fit the experimental data well, confirming that toluene had a direct adverse effect on the catalytic activity of the enzyme. (c) 1996 John Wiley & Sons, Inc.  相似文献   

5.
We examined the effect of a novel disulfide bond engineered in subtilisin E from Bacillus subtilis based on the structure of a thermophilic subtilisin-type serine protease aqualysin I. Four sites (Ser163/Ser194, Lys170/Ser194, Lys170/Glu195, and Pro172/Glu195) in subtilisin E were chosen as candidates for Cys substitutions by site-directed mutagenesis. The Cys170/Cys195 mutant subtilisin formed a disulfide bond in B. subtilis, and showed a 5-10-fold increase in specific activity for an authentic peptide substrate for subtilisin, N-succinyl-L-Ala-L-Ala-L-Pro-L-Phe-p-nitroanilide, compared with the single-Cys mutants. However, the disulfide mutant had a 50% decrease in catalytic efficiency due to a smaller k(cat) and was thermolabile relative to the wild-type enzyme, whereas it was greatly stabilized relative to its reduced form. These results suggest that an electrostatic interaction between Lys170 and Glu195 is important for catalysis and stability in subtilisin E. Interestingly, the disulfide mutant was found to be more stable in polar organic solvents, such as dimethylformamide and ethanol, than the wild-type enzyme, even under reducing conditions; this is probably due to the substitution of uncharged Cys by charged surface residues (Lys170 and Glu195). Further, the amino-terminal engineered disulfide bond (Gly61Cys/Ser98Cys) and the mutation Ile31Leu were introduced to enhance the stability and catalytic activity. A prominent 3-4-fold increase in the catalytic efficiency occurred in the quintet mutant enzyme over the range of dimethylformamide concentration (up to 40%).  相似文献   

6.
Alexander PA  Ruan B  Bryan PN 《Biochemistry》2001,40(35):10634-10639
Subtilisin BPN' contains two cation binding sites. One specifically binds calcium (site A), and the other can bind both divalent and monovalvent metals (site B). By binding at specific sites in the tertiary structure of subtilisin, cations contribute their binding energy to the stability of the native state and increase the activation energy of unfolding. Deconvoluting the influence of binding sites A and B on the inactivation rate of subtilisin is complicated, however. This paper examines the stabilizing effects of cation binding at site B by using a mutant of subtilisin BPN' which lacks calcium site A. Using this mutant, we show that calcium binding at site B has relatively little effect on stability in the presence of moderate concentrations of monovalent cations. At [NaCl] =100 mM, site B is >or=98% occupied with sodium, and therefore its net occupancy with a cation varies little as subtilisin is titrated with calcium. Exchanging sodium for calcium results in a 5-fold decrease in the rate of inactivation. In contrast, because of the high selectivity of site A for calcium, its occupancy changes dramatically as calcium concentration is varied, and consequently the inactivation rate of subtilisin decreases approximately 200-fold as site A becomes saturated with calcium, irrespective of the concentration of monovalent cations.  相似文献   

7.
Alexander PA  Ruan B  Strausberg SL  Bryan PN 《Biochemistry》2001,40(35):10640-10644
Stability is a property of subtilisin which has proven particularly amenable to enhancement via random mutagenesis and screening, yet the effects of most stabilizing mutations are not understood in structural and energetic detail. This paper seeks to explain the longstanding observation that stabilizing mutations are usually calcium-dependent in their stabilizing effect, irrespective of their proximity to the calcium binding sites. Stabilizing mutations in subtilisin fall into one of three classes. The largest class of mutations stabilize only in the presence of excess calcium. A smaller number of mutations stabilize independently of [calcium], and a few mutations stabilize only in the presence of chelating agents, such as EDTA. This study compares the effects of mutations from each class when introduced into subtilisin BPN' and two calcium-free versions of subtilisin. The calcium-dependent effects of mutations can be explained by considering subtilisin to be in conformational equilibrium between two structurally similar but energetically distinct states: N and N*. The equilibrium from the N* to the N state can be altered either by calcium binding to site A or by mutation. Mutations which stabilize only in the presence of calcium stabilize the N state relative to N*. Mutations which stabilize only in the presence of chelants stabilize the N* state relative to N. As a byproduct of this analysis, we have developed a hyperstable variant of subtilisin whose inactivation at high temperature in the presence of EDTA is 10(5) times slower than wild-type subtilisin.  相似文献   

8.
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   

9.
10.
11.
We used directed evolution to convert Bacillus subtilis subtilisin E into an enzyme functionally equivalent to its thermophilic homolog thermitase from Thermoactinomyces vulgaris. Five generations of random mutagenesis, recombination and screening created subtilisin E 5-3H5, whose half-life at 83 degrees C (3.5 min) and temperature optimum for activity (Topt, 76 degrees C) are identical with those of thermitase. The Topt of the evolved enzyme is 17 degrees C higher and its half-life at 65 degrees C is >200 times that of wild-type subtilisin E. In addition, 5-3H5 is more active towards the hydrolysis of succinyl-Ala-Ala-Pro-Phe-p-nitroanilide than wild-type at all temperatures from 10 to 90 degrees C. Thermitase differs from subtilisin E at 157 amino acid positions. However, only eight amino acid substitutions were sufficient to convert subtilisin E into an enzyme equally thermostable. The eight substitutions, which include known stabilizing mutations (N218S, N76D) and also several not previously reported, are distributed over the surface of the enzyme. Only two (N218S, N181D) are found in thermitase. Directed evolution provides a powerful tool to unveil mechanisms of thermal adaptation and is an effective and efficient approach to increasing thermostability without compromising enzyme activity.  相似文献   

12.
13.
Directed evolution of enzyme stability   总被引:7,自引:0,他引:7  
Modern enzyme development relies to an increasing extent on strategies based on diversity generation followed by screening for variants with optimised properties. In principle, these directed evolution strategies might be used for optimising any enzyme property, which can be screened for in an economically feasible way, even if the molecular basis of that property is not known. Stability is an interesting property of enzymes because (1) it is of great industrial importance, (2) it is relatively easy to screen for, and (3) the molecular basis of stability relates closely to contemporary issues in protein science such as the protein folding problem and protein folding diseases. Thus, engineering enzyme stability is of both commercial and scientific interest. Here, we review how directed evolution has contributed to the development of stable enzymes and to new insight into the principles of protein stability. Several recent examples are described. These examples show that directed evolution is an effective strategy to obtain stable enzymes, especially when used in combination with rational or semi-rational engineering strategies. With respect to the principles of protein stability, some important lessons to learn from recent efforts in directed evolution are (1) that there are many structural ways to stabilize a protein, which are not always easy to rationalize, (2) that proteins may very well be stabilized by optimizing their surfaces, and (3) that high thermal stability may be obtained without forfeiture of catalytic performance at low temperatures.  相似文献   

14.
DNAzymes represent a new generation of catalytic nucleic acids for specific RNA targeting in order to inhibit protein translation from the specifically cleaved mRNA. The 10-23 DNAzyme was found to hydrolyze RNA in a sequence-specific manner both in vitro and in vivo. Although single-stranded DNAzymes may represent the most effective nucleic acid drug to date, they are nevertheless sensitive to nuclease degradation and require modifications for in vivo application. However, previously used stabilization of DNAzymes by site-specific phosphorothioate (PT) modifications reduces the catalytic activity, and the PTO displays toxic side effects when applied in vivo. Thus, improving the stability of DNAzymes without reducing their catalytic activity is essential if the potential of these compounds should be realized in vivo. RESULTS: The Circozyme was tested targeting the mRNA of the most common genetic rearrangement in pediatric acute lymphoblastic leukemia TEL/AML1 (ETV6/RUNX1). The Circozyme exhibits a stability comparable to PTO-modified DNAzymes without reduction of catalytic activity and specificity and may represent a promising tool for DNAzyme in vivo applications. CONCLUSION: The inclusion of the catalytic site and the specific mRNA binding sequence of the DNAzyme into a circular loop-stem-loop structure (Circozyme) of approximately 70 bases presented here represents a new effective possibility of DNAzyme stabilization.  相似文献   

15.
In vivo formation and stability of engineered disulfide bonds in subtilisin   总被引:9,自引:0,他引:9  
Computer modeling suggested that a disulfide bond could be built into Bacillus amyloliquefaciens subtilisin between positions 22 (wild-type, Thr) and 87 (Ser) or between positions 24 (Ser) and 87 (Ser). Single cysteines were introduced into this cysteine-free protease at positions 22, 24, or 87 by site-directed mutagenesis of the cloned subtilisin gene. The corresponding double-cysteine mutants were constructed, and recombinant plasmids were expressed in Bacillus subtilis. Double-cysteine mutant enzymes were secreted as efficiently as wild-type, and disulfide bonds were formed quantitatively in vivo. These disulfide bonds were introduced approximately 24 A away from the catalytic site and had no detectable effect on either the specific activities or the pH optima of the mutant enzymes. The equilibrium constants for the reduction of the mutant disulfide bonds by dithiothreitol were determined to be 82 +/- 22 and 20 +/- 5 for Cys22/Cys87 and Cys24/Cys87, respectively. Studies of autoproteolytic inactivation of wild-type subtilisin support a relationship between autolytic stability and conformational stability of the protein. The stabilities of Cys24/Cys87 and wild-type enzymes to autolysis were essentially the same; however, Cys22/Cys87 was actually less stable to autolysis. Reduction of the disulfide cross-bridge lowered the autolytic stability of both double-cysteine mutants relative to their disulfide forms. This correlates with a lowered autolytic stability for the Cys22 and Cys87 single-cysteine mutants, and the fact that an intramolecular hydrogen bond between the hydroxyl groups of Thr22 and Ser87 is likely to be disrupted in the Cys22 and Cys87 single-cysteine mutant proteins.  相似文献   

16.
Remarkable increases in enzyme catalytic stability resulting from addition of charged water-soluble polymers have recently been reported, suggesting that use of these polymers may be an attractive general strategy for enzyme stabilization. To test the proposed hypothesis that coulombic forces between water-soluble polymers and enzymes are primarily responsible for enzyme stabilization, we examined the catalytic stability and activity of two enzymes in the presence of polymers differing in net charge. All polymers tested increased enzyme lifetimes, regardless of their net charge, suggesting that stabilization of these enzymes by water-soluble polymers is not solely dependent on simple electrostatic interactions between the polymers and enzymes.  相似文献   

17.
Papain was modified with the anhydrides of various monocarboxylic (acetic or propionic) and dicarboxylic (citraconic, maleic or succinic) acids. 7–10 of the 11 primary amino groups of the enzyme were modified. The organic solvent tolerances of the modified enzyme forms were increased (especially in the concentration range of 10–60%) in comparison with the unmodified enzyme. Acylation enhanced the catalytic activity and stability of papain both in buffer and in aqueous organic solvents (ethanol and acetonitrile). Decrease of the positive charges on the surface of papain resulted in a higher enzyme stability than when they were replaced by negative charges. The kinetic parameters revealed that in aqueous ethanol the maximum rates (Vmax) and Michaelis constants (KM) of the modified papain forms were increased, and higher catalytic efficiencies (kcat/KM) were detected as compared with the native enzyme. The results of near-UV circular dichroism and tryptophan fluorescence spectroscopic studies suggested that the modifications caused only local changes around the aromatic residues. The modified enzyme forms led to higher N-acetyl-l-tyrosine ethyl ester synthesis conversions in aqueous ethanol; acetyl and propionyl papain furnishing the highest productivity.  相似文献   

18.
Colyophilization with methyl-beta-cyclodextrin activates subtilisin Carlsberg by more than 200-fold in organic solvents, though this is a short-lived effect. About 93% of the enzyme's high initial activity observed in THF (at 45 degrees C) decreases exponentially with a t(1/2) of 1.8 h, until it reaches a residual activity (of 7%) that remains constant throughout the 4 days duration of the experiment. A further study of this enzyme reveals a general trend: the initial activities of the lyophilized powder and the cross-linked enzyme crystals are also greatly reduced upon incubation in this solvent, although these preparations retain 50% of their activity after about 20 h of incubation. All of the preparations studied retained some residual activity (which persisted throughout the duration of the experiments) after the initial exponential decay. The data here presented suggest that the mode of enzyme preparation is an important issue to consider when planning lengthy reactions.  相似文献   

19.
Colyophilization or codrying of subtilisin Carlsberg with the crown ethers 18-crown-6, 15-crown-5, and 12-crown-4 substantially improved enzyme activity in THF, acetonitrile, and 1,4-dioxane in the transesterification reactions of N-acetyl-L-phenylalanine ethylester and 1-propanol and that of (+/-)-1-phenylethanol and vinylbutyrate. The acceleration of the initial rate, V(0), ranged from less than 10-fold to more than 100-fold. All crown ethers activated subtilisin substantially, which excludes a specific macrocyclic effect from being responsible. The secondary structure of subtilisin was studied by Fourier-transform infrared (FTIR) spectroscopy. 18-Crown-6 and 15-crown-5 led to a more nativelike structure of subtilisin in the organic solvents employed when compared with that of the dehydrated enzyme obtained from buffer alone. However, the high level of activation with 12-crown-4 where this effect was not observed excluded overall structural preservation from being the primary cause of the observed enzyme activation. The conformational mobility of subtilisin was investigated by performing thermal denaturation experiments in 1,4-dioxane. Although only a small effect of temperature on subtilisin structure was observed for the samples prepared with or without 12-crown-4, both 18-crown-6 and 15-crown-5 caused the enzyme to denature at quite low temperatures (38 degrees C and 56 degrees C, respectively). No relationship between this property and V(0) was evident, but increased conformational mobility of the protein decreased its storage stability. The possibility of a "molecular imprinting" effect was also tested by removing 18-crown-6 from the subtilisin-18-crown-6 colyophilizate by washing. V(0) was only halved as a result of this procedure, an effect insignificant compared with the ca. 80-fold rate enhancement observed prior to washing in THF. This suggests that molecular imprinting is likely the primary cause of subtilisin activation by crown ethers, as recently suggested.  相似文献   

20.
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