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.
It is scientifically and industrially important to clarify the stabilizing mechanism of proteases in extraordinary environments. We used subtilisins ALP I and Sendai as models to study the mechanism. Subtilisin ALP I is extremely sensitive to highly alkaline conditions, even though the enzyme is produced by alkalophilic Bacillus, whereas subtilisin Sendai from alkalophilic Bacillus is stable under conditions of high alkalinity. We constructed mutant subtilisin ALP I enzymes by mutating the amino acid residues specific for subtilisin ALP I to the residues at the corresponding positions of amino acid sequence alignment of alkaline subtilisin Sendai. We observed that the two mutations in the C-terminal region were most effective for improving stability against surfactants and heat as well as high alkalinity. We predicted that the mutated residues are located on the surface of the enzyme structures and, on thebasis of three-dimensional modelling, that they are involved in stabilizing the conformation of the C-terminal region. As proteolytic enzymes frequently become inactive due to autocatalysis, stability of these enzymes in an extraordinary environment would depend on the conformational stability of the molecular surface concealing scissile peptide bonds. It appeared that the stabilization of the molecular surface structure was effective to improve the stability of the proteolytic enzymes.  相似文献   

2.
Autodigestion of subtilisin ALP I (ALP I), secreted from the alkalophilic Bacillus sp. NKS-21 and its predicted amino acid sequence having about 60% identity with other alkaline subtilisins, was examined under alkaline conditions. At the alkaline pH of 12, ALP I was rapidly degraded, and almost no breakdown products were detectable. However, by incubating ALP I at 5 degrees C for an extended time, a couple of specific peptides (26.7 kDa and 25.6 kDa) were accumulated. Each of them was purified and amino acid sequences of these fragments were found. Both peptides appeared to start at Gly-19 of the mature sequence of ALP I.  相似文献   

3.
The gene for an alkaline serine protease from alkalophilic Bacillus sp. NKS-21 (subtilisin ALP I) was cloned, and its nucleotide sequence was determined. The gene (aprQ) contained an open reading frame of 1125 bp, encoding a primary product of 374 amino acids. The mature protease, composed of 272 amino acids, was preceded by a putative signal sequence of 37 amino acids and a pro-sequence of 65 amino acids. The mature protease conserved the catalytic triad, Asp, His, and Ser, as subtilisin BPN or other subtilisins, and the subtilisin ALP I might belong to the subtilisin super family. The primary structure of subtilisin ALP I was compared and discussed with those of 13 subtilisins, 5 subtilisins from alkalophilic Bacillus, and 8 from neutrophiles. Low homology was shown between subtilisin ALP I and subtilisins from alkalophiles or subtilisins from neutrophiles. Forty-five amino acid residues of the mature protein of subtilisin ALP I were entirely independent of other subtilisins. According to the homology of ALP I with other subtilisins, subtilisin ALP I might be in the middle point between alkaline subtilisins and neutral ones.  相似文献   

4.
An ultraviolet absorption difference spectrum that is typical of a change in ionization state (pKa 9.7 leads to greater than 11.5) of a tyrosyl residue has been observed on the binding between Streptomyces subtilisin inhibitor (SSI) and subtilisin BPN' [EC 3.4.21.14] at alkaline pH, ionic strength 0.1 M, at 25 degrees C (Inouye, K., Tonomura, B., and Hiromi, K., submitted). When the complex of SSI and subtilisin BPN' is formed at an ionic strength of 0.6 M and pH 9.70, the characteristic features of the protonation of a tyrosyl residue in the difference spectrum are diminished. These results suggest that the pKa-shift of a tyrosyl residue observed at alkaline pH and lower ionic strength results from an electrostatic interaction. Nitration of tyrosyl residues of SSI and of subtilisin BPN' was performed with tetranitromethane (TNM). By measurements of the difference spectra observed on the binding of the tyrosyl-residue-nitrated SSI and the native subtilisin BPN', and on the binding of the native SSI and the tyrosyl-residue-nitrated subtilisin BPN' and alkaline pH, the tyrosyl residue in question was shown to be one out of the five tyrosyl residues of pKa 9.7 of the enzyme. This tyrosyl residue was probably either Tyr 217 or Tyr 104 on the basis of the reactivities of tyrosyl residues of the enzyme with TNM and their locations on the enzyme molecule. Carboxyl groups of SSI were modified by covalently binding glycine methyl ester with the aid of water-soluble carbodiimide, in order to neutralize the negative charges on SSI. In the difference spectrum which was observed on the binding of subtilisin BPN' and the 5.3-carboxyl-group-modified SSI at alkaline pH, the characteristic features of the protonation of a tyrosyl residue were essentially lost, and the difference spectrum is rather similar to that observed on the binding of the native SSI and the enzyme at neutral pH. This phenomenon indicates that the pKa of a tyrosyl residue of the enzyme is shifted upwards by interaction with carboxyl group(s) of SSI on the formation of the enzyme-inhibitor complex.  相似文献   

5.
In this report, we demonstrate an interaction between subtilisin NAT (formerly designated BSP, or nattokinase), a profibrinolytic serine proteinase from Bacillus subtilis, and plasminogen activator inhibitor 1 (PAI-1). Subtilisin NAT was purified to homogeneity (molecular mass, 27.7 kDa) from a saline extract of B. subtilis (natto). Subtilisin NAT appeared to cleave active recombinant prokaryotic PAI-1 (rpPAI-1) into low molecular weight fragments. Matrix-assisted laser desorption/ionization in combination with time-of-flight mass spectroscopy and peptide sequence analysis revealed that rpPAI-1 was cleaved at its reactive site (P1-P1': Arg(346)-Met(347)). rpPAI-1 lost its specific activity after subtilisin NAT treatment in a dose-dependent manner (0.02-1.0 nm; half-maximal effect at approximately 0.1 nm). Subtilisin NAT dose dependently (0.06-1 nm) enhanced tissue-type plasminogen activator-induced fibrin clot lysis both in the absence of rpPAI-1 (48 +/- 1.4% at 1 nm) and especially in the presence of rpPAI-1 (78 +/- 2.0% at 1 nm). The enhancement observed in the absence of PAI-1 seems to be induced through direct fibrin dissolution by subtilisin NAT. The stronger enhancement by subtilisin NAT of rpPAI-1-enriched fibrin clot lysis seems to involve the cleavage and inactivation of active rpPAI-1. This mechanism is suggested to be important for subtilisin NAT to potentiate fibrinolysis.  相似文献   

6.
Deng A  Wu J  Zhang G  Wen T 《Biochimie》2011,93(4):783-791
High-alkaline proteases are of great importance because of their proteolytic activity and stability under high-alkaline condition. We have previously isolated a new protease (AprB) which has potential industrial applications based on its high-alkaline adaptation. However, the molecular and structural basis for alkaline adaptation of this enzyme has not been fully elucidated. In the present study, AprB gene was cloned and expressed in the Bacillus subtilis WB600. This gene codes for a protein of 375 amino acids comprised with a 28-residual signal peptide, a 78-residual pro-peptide, and a 269-residual mature protein. The deduced amino acid sequence has the highest homology of 63.2% with that of the high-alkaline proteases. Recombinant AprB was purified and determined to be monomeric with molecular mass of 26.755 kDa. The NH2-terminal sequence of the purified AprB was A-Q-S-I-P-W-G-I-E-R. This enzyme exhibited high catalytic efficiencies (Kcat/Km) towards natural, modified, and synthesis substrates with optimal activity at 60 °C and pH 10. AprB was stable over a wide range of pH 5 to 11 and various surfactants, and could be activated by Mg2+, Ca2+ and Ba2+. The structural properties of AprB, like a higher ratio of R/(R + K), a larger area of hydrophobic surface, increased number of ion pairs formed by Arg residue, and the exposure of Asp active residue on the surface, might be responsible for its alkaline adaptation. In contrast with members of subtilisin family, such as M-protease and subtilisin BPN′, AprB harbored a high content of Glu and Asp residues, and a low content of Arg and Lys residues on the surface. Interestingly, these structural characters were similar with that of psychrophilic proteases, which suggested that these molecular factors were not restricted in the psychrophilic proteases, and therefore were not solely responsible for their cold-adaptation. Our results reveal a novel structural feature of AprB unique to subtilisin family and provide clues for its alkaline adaptation.  相似文献   

7.
Bacillus subtilis IBTC-3 subtilisin was purified by gel filtration on Sephadex G 75 and affinity chromatography on bacitracin-CNBr-Sepharose 4B and characterized. Its molecular mass of 27 kDa was determined by SDS-PAGE, and isoelectric pH of 8.4 by chromatofocusing. FT-Raman and FT-IR spectroscopy studies revealed fragments with alpha-helix and irregular secondary structures within the polypeptide chain. The beta-sheet conformation was observed only in second-derivatives of FT-RS and FT-IR spectra, in the range of the amide II, III, and I bands. Tyr residues were shown to be hydrogen bonded and CSCH(3) groups adopted two conformations (P(H)-T and P(C)-G conformers). Kinetic properties of B. subtilis IBTC-3 subtilisin in hydrolysis of ethyl esters of amino acid derivatives were compared with that of alkaline peptidase from Bacillus alcalophilus PB92. The first enzyme displayed the highest affinity for NAc-Phe-OEt, both in hydrolysis (K(m) of 0.22 mM) and in synthesis (K(m) of 0.85 mM), whereas PB92 peptidase preferred Tyr derivatives (NAc-Tyr-OEt, K(m) of 0.043 and 0.75 mM, respectively). In contrast to the latter enzyme, B. subtilis IBTC-3 subtilisin catalyzed hydrolysis and synthesis of Bz-Arg-OEt.  相似文献   

8.
Alkaline elastase YaB is an extracellular serine protease of the alkalophilic Bacillus strain YaB. We cloned the structural gene, ale, and determined the nucleotide sequence. The mature enzyme (268 amino acids) was preceded by a putative signal sequence and a prosequence (27 and 83 amino acids, respectively). The mature enzyme was 55% homologous to subtilisin BPN'. Almost all the positively charged residues are predicted to be on the surface of the molecule, which would facilitate binding to elastin. The P1 substrate site-related sequences differed between alkaline elastase YaB and subtilisin BPN'.  相似文献   

9.
Interaction of human plasma alpha 1-proteinase inhibitor (alpha 1PI) with subtilisin BPN' was assessed by spectrophotometric determination of the inhibitory capacity and by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). During the course of incubation of the enzyme and the inhibitor (E : I = 1 : 7.5) at pH 8.0 about 17% of the enzyme activity which had been inhibited initially was regenerated, indicating a temporary type of inhibition. The results of the titration experiments indicate that 9.8 mol of the inhibitor is required to inhibit 1 mol of the enzyme completely. However, patterns of 5% disc SDS-PAGE under non-reducing conditions revealed only an equimolar complex (Mr80K) of alpha 1PI with the enzyme and no other higher Mr component than the native inhibitor (Mr 56K). On the other hand, complete dissociation of the complex occurred under reducing conditions, producing an enzymatically modified inhibitor. When 5 21% gradient slab SDS-PAGE was employed, no complex formation was observed under either reducing or non-reducing conditions. With the gradient gel system, dissociation of the equimolar complex produced different forms of the inhibitor, that is, regeneration of an intact alpha 1PI under non-reducing conditions and an enzymatically modified form under reducing conditions. All these results indicate that the complex formed between subtilisin BPN' and human alpha 1PI is not so stable as that of the inhibitor with bovine chymotrypsin and that no covalent bond may be involved in the complex formation. The results also indicate that human alpha 1PI is not an effective inhibitor of subtilisin BPN' and behaves like a substrate for the enzyme.  相似文献   

10.
Analogous soluble and insoluble derivatives of subtilisin Novo (EC 3.4.21.14) were prepared by coupling the enzyme to CNBr-activated DEAE-dextran and DEAE-Sephadex, respectively. The DEAE-dextran-subtilisin displayed pH optima and Km values for ester hydrolysis similar to subtilisin, whereas the pH versus activity profiles obtained with DEAE-Sephadex-subtilisin were shifter towards the alkaline pH region and the Km values were increased. Compared with subtilisin, DEAE-dextran-subtilisin showed a 40-65% reduction of kcat for hydrolysis of N-acetyl-L-tyrosine ethyl ester, p-tosyl-L-arginine methyl ester and benzyloxycarbonyl-glycyl-L-tyrosinamide and its maximum velocities for digestion of casein and clupein also amounted to 40-60% of the subtilisin values. With Deae-sephadex-subtilisin, in contrast, the maximum velocity of hydrolysis decreased to a greater extent for polypeptide substrates compared to ester substrates. The present results indicate that the chemical nature of a support can effect intrinsic properties of a matrix-bound enzyme in addition to the steric and diffusional effects usually observed with polymer-attached enzymes.  相似文献   

11.
Using molecular dynamics simulations, we have obtained an important insight into the structural and dynamical changes exerted by a nonaqueous solvent on the serine protease subtilisin Carlsberg. Our findings show that the structural properties of the subtilisin–acetonitrile (MeCN) system were sensitive to the amount of water present at the protein surface. A decrease or lack of water promoted the enzyme–MeCN interaction, which increased structural changes of the enzyme primarily at the surface loops. This effect caused variations on the secondary and tertiary structure of the protein and induced the opening of a pathway for the solvent to the protein core. Also, disturbance of the oxyanion hole was observed due to changes in the orientation in the Asn-155 side chain. The disruption of the oxyanion hole and the changes of the tertiary structure should affect the optimal activity of the enzyme.  相似文献   

12.
The activity of immobilized subtilisin BPN' on pure cellulose-based membrane support was investigated using site-directed and random immobilization approaches. The catalytic activity of site-directed immobilized subtilisin on pure cellulose fiber-based materials was found to be 81% of that in homogeneous solution, while that of randomly immobilized subtilisin was 27%. Pure cellulose membrane supports provided large surface areas for high enzyme loading without diffusional limitations. The activity of immobilized subtilisin on pure cellulose support was more than twice that on a modified polyether sulfone (MPS) membrane, which was attributed to the higher hydrophilicity of cellulose. Immobilized subtilisin maintained its initial activity for 14 days at 4 degrees C and 7 days at 24 degrees C. The immobilized enzyme could resist higher temperature and operate over a wider range of pH without loss of activity. This study showed that pure cellulose fiber-based membranes are well suited for enzyme immobilization and biocatalysis.  相似文献   

13.
m-Nitrobenzeneboronic acid as a possible transition-state analog for serine proteases was found to cause absorption spectral change from 250 nm 350 nm upon binding with subtilisin BPN' (EC 3.4.21.14) at pH 6.5. Similar difference spectral changes of m-nitrobenzeneboronic acid were also observed at alkaline pH or upon addition of N-methylimidazole at pH 6.5. A characteristic circular dichroism spectrum of m-nitrobenezeneboronic acid was induced upon binding with subtilisin BPN' not only at pH 6.5, but also at alkaline pH. Circular dichroism spectral titration confirmed the stoichiometry of 1 : 1 for the m-nitrobenzeneboronic acid - subtilisin complex. m-Nitrobenzeneboronic acid was shown to be useful as a reversible chromophoric probe for the catalytic site of serine proteases.  相似文献   

14.
We employed random mutagenesis to determine the region of the initial unfolding of hyper-alkaline-sensitive subtilisin, ALP I, that precedes the denaturation of the entire protein under highly alkaline conditions. This region comprises two α-helices and a calcium-binding loop. Stabilization of the region caused the stabilization of the entire protein at a high alkaline pH 12. The alkaline stability of this region was most effectively improved by hydrophobic interactions, followed by ionic interactions with Arg residues. The effect of mutations on the improvement was different with regard to the alkaline stability and thermostability. This indicated that different strategies were necessary to improve the alkaline stability and thermostability of the protein.  相似文献   

15.
Subtilisin BPN′ was immobilized to porous glass via isothiocyanate coupling. The pH optimum of the enzyme was shifted to the alkaline side on binding. This effect was more pronounced with ethyl lactate than with N-tosyl arginine methyl ester (TAME). Presumably, the shift is a reflection of the negative charge on the surface of the glass. The Michaelis constant and Vmax of soluble subtilisin BPN′ with TAME were two and one orders of magnitude, respectively, lower than with ethyl lactate. Vmax, calculated per g of active enzyme, with TAME as the substrate was not affected by immobilization, while Vmax with ethyl lactate decreased greater than tenfold. The apparent KM decreased on immobilization with ethyl lactate as substrate and increased with TAME. Results are explained in terms of diffusional resistance and a possible attraction of ethyl lactate to the glass surface. Active site titration indicated that about 25%, of the immobilized enzyme was active.  相似文献   

16.
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   

17.
Circular dichroic spectroscopy has been used to study the effect of pH, guanidinium hydrochloride concentration and temperature on the conformation of the fungal subtilisin-like proteinase K and the bacterial DY. The ellipticity of the bands in the far ultraviolet region remains almost unchanged in the pH range 3.0-11.0 (PMS-proteinase K) and 5.0-10.0 (PMS-subtilisin DY). The same ranges of pH stability were determined from the pH dependence of the near ultraviolet dichroic spectra. Hence the changes in the tertiary and secondary structure occur in parallel. Proteinase K is considerably more stable at acidic and somewhat more stable at alkaline pH than subtilisin DY. At neutral pH proteinase K is more resistant to denaturation by guanidinium hydrochloride than is subtilisin DY. The midpoints of the denaturation curves were 6.2 M and 3.2 M guanidinium, respectively. The thermal unfolding of proteinase K occurred at a higher temperature than for subtilisin DY, the transition midpoints being 65 degrees and 48 degrees, respectively. Thus proteinase K is overall a much more robust molecule than subtilisin DY, showing greater resistance to all three forms of denaturation. The differences in the stability of the two proteinases can be partly explained by differences in their calcium binding sites.  相似文献   

18.
At pH 6.3 both the native and subtilisin-digested fructose-1,6-bisphosphatase (Fru-P2-ase) molecules exhibit four fast-reacting thiol groups. The kinetic analysis shows that the pK value for the reaction of these thiols is 8.1. The increase of pH from 6.3 to 9.3 results in an uncovering of the remaining 20 thiol groups. In subtilisin-cleaved enzyme the rate of reaction of SH groups is considerably higher than in the native enzyme at pH 9.3, indicating changes in the microenvironments around thiols upon modification. A fluorescent label inserted on a fast-reacting SH group and neighboring NH2 group shifts the pH optimum of the enzyme to alkaline region and decreases its sensitivity toward AMP. Spectral analysis of labeled enzyme indicates that the labeled region of protein is more hydrophilic upon proteolytic digestion. It is concluded that a molecule of subtilisin-digested enzyme has a more relaxed structure than the native enzyme. The relaxation of the enzyme to a new conformation is reflected by urea addition, which mimics the effect of subtilisin digestion. Correlation of enzyme activity versus its sensitivity toward AMP (I 0.5), shows that at low concentrations of urea the active-site region at pH 6.3 is more affected than the region of AMP binding.  相似文献   

19.
The structural and functional properties of arginine kinase (AK) in alkaline conditions in the absence or presence of salt have been investigated. The conformational changes of AK during alkaline unfolding and salt-induced folding at alkaline pH were monitored using intrinsic fluorescence emission, binding of the fluorescence probe 1-anilino-8-naphthalenesulfonate and circular dichroism. The results for the alkaline unfolded enzyme showed that much lower pH (11.0) was required to cause the complete loss of AK activity than was required to cause an obvious conformational change of the enzyme. Compared with the completely unfolded state in 5 M urea, the high pH denatured enzyme had some residual secondary and tertiary structure even at pH 13.0. Increasing the ionic strength by adding salt at pH 12.75 resulted in the formation of a relatively compact tertiary structure and a little new secondary structure with hydrophobic surface enhancement. These results indicate that the partially folded state formed under alkaline conditions may have similarities to the molten globule state which is compact, but it has a poorly defined tertiary structure and a native-like secondary structure.  相似文献   

20.
M Philipp  I H Tsai  M L Bender 《Biochemistry》1979,18(17):3769-3773
The p-nitrophenyl esters of straight-chain fatty acids were used as substrates of the enzyme subtilisin Novo (EC 3.4.4.16) and its chemically produced artificial enzyme thiolsubtilisin. Subtilisin and thiolsubtilisin pH--activity profiles were determined, and kinetic effects of the active site O-S substitution were observed. Among the substrates tested, both enzymes show highest specificity with p-nitrophenyl butyrate. It was also found that subtilisin is more sensitive to changes in substrate chain length than is thiolsubtilisin. Second-order acylation rate constants (k2/Ks) are remarkably similar for both enzymes. However, thiolsubtilisin deacylation rate constants and Km values are lower than analogous subtilisin constants. While thiolsubtilisin deacylation rate constants give a pH profile identical with that of subtilisin, the pH profile of thiolsubtilisin acylation rate constants shows an active site pK value lowered from the subtilisin pK of 7.15 and exhibits an inflection point at pH 8.45, which is absent in subtilisin.  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