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 46 毫秒
1.
It has been shown that the P1 site (the center of the reactive site) of protease inhibitors corresponds to the specificity of the cognate protease, and consequently specificity of Streptomyces subtilisin inhibitor (SSI) can be altered by substitution of a single amino acid at the P1 site. In this paper, to investigate whether similar correlation between inhibitory activity of mutated SSI and substrate preference of protease is observed for subtilisin BPN', which has broad substrate specificity, a complete set of mutants of SSI at the reaction site P1 (position 73) was constructed by cassette and site-directed mutagenesis and their inhibitory activities toward subtilisin BPN' were measured. Mutated SSIs which have a polar (Ser, Thr, Gln, Asn), basic (Lys, Arg), or aromatic amino acid (Tyr, Phe, Trp, His), or Ala or Leu, at the P1 site showed almost the same strong inhibitory activity toward subtilisin as the wild type (Met) SSI. However, the inhibitory activity of SSI variants with an acidic (Glu, Asp), or a beta-branched aliphatic amino acid (Val, Ile), or Gly or Pro, at P1 was decreased. The values of the inhibitor constant (Ki) of mutated SSIs toward subtilisin BPN' were consistent with the substrate preference of subtilisin BPN'. A linear correlation was observed between log(1/Ki) of mutated SSIs and log(1/Km) of synthetic substrates. These results demonstrate that the inhibitory activities of P1 site mutants of SSI are linearly related to the substrate preference of subtilisin BPN', and indicate that the binding mode of the inhibitors with the protease may be similar to that of substrates, as in the case of trypsin and chymotrypsin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Pleurotus ostreatus proteinase A inhibitor 1 (POIA1), which is composed of 76 residues without disulfide bridges, is a unique inhibitor in that it exhibits sequence similarity to the propeptides of subtilisins. In order to elucidate the inhibitory mechanism of POIA1, we constructed an expression system for a synthetic POIA1 gene. The wild-type POIA1 was found to inhibit subtilisin BPN' with an inhibitor constant (K(i)) of 3.2 x 10(-9) M, but exhibited a time-dependent decrease of inhibitory activity as a consequence of degradation by the protease, showing that the wild-type POIA1 was a temporary inhibitor when subtilisin BPN' was used as a target protease. Since POIA1 shows sequence similarity to the propeptide of subtilisin, which is known to inhibit the protease via its C-terminal region, the C-terminal six residues of POIA1 were replaced with those of the propeptide of subtilisin BPN'. The mutated POIA1 inhibited subtilisin BPN' with a K(i) value of 2.8 x 10(-11) M and did not exhibit time-dependent decrease of inhibitory activity, showing about 100-fold increases in binding affinity for, and resistance to, the protease. These results clearly indicate that the C-terminal region of POIA1 plays an important role in determining the inhibitory activity toward the protease, and that the increase in binding ability to the protease is closely related to resistance to proteolytic degradation. Therefore, the inhibitory properties of POIA1 can be altered by mutation of its C-terminal region.  相似文献   

3.
Yeast proteinase B inhibitor 2 (YIB2), which is composed of 74 amino acid residues, is an unusual serine protease inhibitor, since it lacks disulfide bonds. To identify its reactive site for proteases, we constructed an expression system for a synthetic YIB2 gene and then attempted to change the inhibitory properties of YIB2 by amino acid replacements. The purified wild-type YIB2 inhibited the activity of subtilisin BPN', a protein homologous to yeast proteinase B, although its binding ability was not strong, and a time-dependent decrease in its inhibitory activity was observed, demonstrating that wild-type YIB2 behaves as a temporary inhibitor when subtilisin BPN' is the target protease. Since YIB2 exhibits sequence homology to the propeptide of subtilisin, which inhibits a cognate protease using its C-terminal region, we replaced the six C-termi nal residues of YIB2 with those of the propeptide of subtilisin BPN' to make the mutant YIB2m1. This mutant exhibited markedly increased inhibitory activity toward subtilisin BPN' without a time-dependent decrease in its inhibitory activity. Replacement of only the C-terminal Asn of YIB2 by Tyr, or deletion of the C-terminal Tyr of YIB2m1, inhibited subtilisin, but the ability of these mutants to bind subtilisin and their resistance to proteolytic attack were weaker than those of YIB2m1, indicating that the C-terminal residue contributes to the interaction with the protease to a greater extent than the preceding five residues and that the resistance of YIB2 to proteolyic attack is closely related to its ability to bind a protease. These results demonstrate that YIB2 is a unique protease inhibitor that involves its C-terminal region in the interaction with the protease.  相似文献   

4.
Structure of the complex of Streptomyces subtilisin inhibitor (SSI) with subtilisin BPN' was studied by examining the thermal denaturation and reducibility of disulfide bonds. The denaturation temperature of the complex was significantly higher than that of the enzyme. Two disulfide bonds localized in the inhibitor side were completely reduced in the complex, whereas only one of them was reduced in the free SSI. Gel filtration of the reduced complex solution showed clearly that the main products of reduction of the complex were two peptide fragments of SSI divided at the active site. The resistive disulfide bond in the complexed inhibitor became accessible as a result of a large conformational change due to splitting of the half-reduced inhibitor.  相似文献   

5.
The propeptide of subtilisin BPN', which functions as an intramolecular chaperone and a temporary inhibitor of subtilisin, is unique in that it acquires its three-dimensional structure by formation of a complex with the cognate protease. We previously showed that the successive amino acid replacements Ala47-->Phe, Gly13-->Ile, and Val65-->Ile in the propeptide to increase its hydrophobicity resulted in formation of a tertiary structure, accompanied by increased ability to bind to the protease and increased resistance to proteolysis. In this study, we examined the effects of these tertiary-structure-forming mutations on the intramolecular chaperone activity of the propeptide. The successive amino acid replacements mentioned above were introduced into pro-subtilisin*, possessing a Ser221-->Ala mutation in the catalytic residue. Refolding experiments were started by rapid dilution of the denatured pro-subtilisin*, and formation of tertiary structure in subtilisin was monitored kinetically by increase in tryptophan fluorescence. The wild-type pro-subtilisin* was found to refold with a rate constant of 4.8 x 10(-3) s(-1) in the equation describing an intramolecular process. The Ala47-->Phe replacement in the propeptide resulted in a 1.2-fold increase in the rate constant of subtilisin refolding. When the additional replacement Gly13-->Ile was introduced, refolding of subtilisin was substantially accelerated, and its kinetics could be fitted to a double exponential process composed of a fast phase with a rate constant of 2.1 x 10(-2) s(-1) and a slow phase with a rate constant of 4.5 x 10(-3) s(-1). The rate constant of the fast phase was increased slightly by a further replacement, Val65-->Ile. Since the slow phase is considered to correspond to proline isomerization, we concluded that tertiary-structure-forming mutations in the propeptide produce positive effects on its intramolecular chaperone activity through acceleration of the propeptide-induced formation of the tertiary structure of subtilisin BPN'.  相似文献   

6.
Pleurotus ostreatus proteinase A inhibitor 1 (POIA1) has been shown to be unique among the various serine protease inhibitors in that its C-terminal region appears to be the reactive site responsible for its inhibitory action toward proteases. To investigate in more detail the mechanism of inhibition by POIA1, we have been studying its structural requirements for stable inhibition of proteases. In this study, we focused on hydrophobic Phe residues, which are generally located in the interior of protein molecules. A Phe-->Ala replacement at position 44 or 56 was introduced into a 'parent' mutant of POIA1 that had been converted into a strong and resistant inhibitor of subtilisin BPN' by replacement of its six C-terminal residues with those of the propeptide of subtilisin BPN' and the effects on inhibitory properties and structural stability were examined. Both of the mutated POIA1 molecules not only were found to exhibit decreased ability to bind to subtilisin BPN' (80-fold for the F44A mutant and 13-fold for the F56A mutant), but were also converted to temporary inhibitors that were degraded by the protease. The structural stability of the mutated POIA1 was also lowered, as shown by a 13 degrees C decrease in melting temperature for the F56A mutant. In particular, the F44A mutant was found to lose its tertiary structure, as judged from the circular dichroism spectrum, demonstrating that Phe44 is a strict requirement for structural formation by the POIA1 molecule. These results clearly indicate that stabilization of POIA1 by hydrophobic residues in its molecular interior is required for stable inhibition of the protease. This requirement for a stable tertiary structure is shared with other serine protease inhibitors, but other structural requirements seem to differ, in that strong binding with the protease is required for POIA1 whereas conformational rigidity around the reactive site is essential for many other protease inhibitors.  相似文献   

7.
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   

8.
Protease susceptibility of homologous proteins in their native conformations was studied. This work aims to establish a broad and quantitative basis for the utilization of protease digestion to analyze the local stability of native proteins. Using high-performance liquid chromatography (HPLC) the time course of the proteolytic degradation of intact proteins was quantitatively traced. Rapid separation of peptide fragments with HPLC made possible the elucidation of sequential digestion originating from the cleavage at a very few sites which are locally unstable in the protein structure. Using four serine proteases, chymotrypsin, trypsin, elastase and subtilisin BPN', we found some common trends in proteolysis for a group of proteins of the cytochrome c family. By comparing of the proteolysis and thermal denaturation with ten homologous cytochromes c extracted from horse, beef, Candida krusei, Saccharomyces cerevisiae, chicken, tuna, pigeon, rabbit, dog and rat, protease susceptibility was related to locally unfolding states intrinsic to the native conformation.  相似文献   

9.
Although highly stable toward unfolding, native ribonuclease A is known to be cleaved by unspecific proteases in the flexible loop region near Ala20. With the aim to create a protease-resistant ribonuclease A, Ala20 was substituted for Pro by site-directed mutagenesis. The resulting mutant enzyme was nearly identical to the wild-type enzyme in the near-UV and far-UV circular dichroism spectra, in its activity to 2',3'-cCMP and in its thermodynamic stability. However, the proteolytic resistance to proteinase K and subtilisin Carlsberg was extremely increased. Pseudo-first-order rate constants of proteolysis, determined by densitometric analysis of the bands of intact protein in SDS-PAGE, decreased by two orders of magnitude. In contrast, the rate constant of proteolysis with elastase was similar to that of the wild-type enzyme. These differences can be explained by the analysis of the fragments occurring in proteolysis with elastase. Ser21-Ser22 was identified as the main primary cleavage site in the degradation of the mutant enzyme by elastase. Obviously, this bond is not cleavable by proteinase K or subtilisin Carlsberg. The results demonstrate the high potential of a single mutation in protein stabilization to proteolytic degradation.  相似文献   

10.
Abstract

Protease susceptibility of homologous proteins in their native conformations was studied. This work aims to establish a broad and quantitative basis for the utilization of protease digestion to analyze the local stability of native proteins. Using high-performance liquid chromatography (HPLC) the time course of the proteolytic degradation of intact proteins was quantitatively traced. Rapid separation of peptide fragments with HPLC made possible the elucidation of sequential digestion originating from the cleavage at a very few sites which are locally unstable in the protein structure. Using four serine proteases, chymotrypsin, trypsin, elastase and subtilisin BPN', we found some common trends in proteolysis for a group of proteins of the cytochrome c family. By comparing of the proteolysis and thermal denaturation with ten homologous cytochromes c extracted from horse, beef, Candida krusei, Saccharomyces cerevisiae, chicken, tuna, pigeon, rabbit, dog and rat, protease susceptibility was related to locally unfolding states intrinsic to the native conformation.  相似文献   

11.
Streptomyces griseus metalloendopeptidase II (SGMPII) was shown to form tight complexes with several Streptomyces protein inhibitors which had been believed to be specific to serine proteases, such as Streptomyces subtilisin inhibitor (SSI), plasminostreptin (PS), and alkaline protease inhibitor-2c' (API-2c'), as well as with Streptomyces metalloprotease inhibitor (SMPI). The dissociation constants of complexes between SGMPII and these inhibitors were successfully determined by using a novel fluorogenic bimane-peptide substrate. The values ranged from nM to pM. The results of studies by gel chromatographic and enzymatic analyses indicated that SGMPII is liberated from the complex with SSI by the addition of subtilisin BPN'. SGMPII and subtilisin BPN' proved, therefore, to interact with SSI in a competitive manner, despite the difference in the chemical nature of their active sites.  相似文献   

12.
This report examines structural changes in a highly mutated, clinical multidrug-resistant HIV-1 protease, and the crystal structure has been solved to 1.3 A resolution in the absence of any inhibitor. This protease variant contains codon mutations at positions 10, 36, 46, 54, 62, 63, 71, 82, 84, and 90 that confer resistance to protease inhibitors. Major differences between the wild-type and the variant include a structural change initiated by the M36V mutation and amplified by additional mutations in the flaps of the protease, resulting in a "wide-open" structure that represents an opening that is 8 A wider than the "open" structure of the wild-type protease. A second structural change is triggered by the L90M mutation that results in reshaping the 23-32 segment. A third key structural change of the protease is due to the mutations from longer to shorter amino acid side chains at positions 82 and 84.  相似文献   

13.
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   

14.
15.
A Tamura  K Kimura  H Takahara  K Akasaka 《Biochemistry》1991,30(47):11307-11313
Cold denaturation and heat denaturation of the protein Streptomyces subtilisin inhibitor (SSI) were studied in the pH range 1.84-3.21 and in the temperature range -3-70 degrees C by circular dichroism and scanning microcalorimetry. The native structure of the protein was apparently most stabilized at about 20 degrees C and was denatured upon heating and cooling from this temperature. Each denaturation was reversible and cooperative, proceeding in two-state transitions, that is, from the native state to the cold-denatured state or from the native state to the heat-denatured state. The two denatured states, however, were not perfect random-coiled structures, and they differed from each other, indicating that there exist three states in this temperature range, i.e., cold denatured, native, and heat denatured. The difference between the cold and heat denaturations was indicated first by circular dichroism. The isodichroic point for the transition from the native state to the cold-denatured state was different from that from the native state to the heat-denatured state in the pH range between 3.21 and 2.45. Moreover, molar ellipticity for the cold-denatured state was different from that of the heat-denatured state, and the transition from the former to the latter was observed at pH values below 2. Values of van't Hoff enthalpies from the native state to the heat-denatured state at pH values between 3.21 and 2.45 were obtained by curve fitting of the CD data, and delta Cp = 1.82 (+/- 0.11) [kcal/(mol.K)] was obtained from the linear plot of the enthalpies against temperature. The parameters obtained from the heat denaturation studies gave curves for delta G zero which were not in agreement with the experimental data in the cold denaturation region when extrapolated to the low temperature. Moreover, the value of the apparent delta Cp for the cold denaturation in the pH range 3.03-2.45 was estimated to be different from that for the heat denaturation, indicating that the mechanism of the cold denaturation of SSI is different from a simple cold denaturation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The entire subtilisin structural gene from Bacillus subtilis I168 has been cloned, and its nucleotide sequence has been determined. When expressed on a high-copy-number shuttle vector, a fivefold increase in serine protease activity was observed. The DNA sequence of the gene is 80% homologous to the Bacillus amyloliquefaciens subtilisin structural gene, and the translated mature coding sequence is 85% homologous to the published protein sequence of subtilisin BPN'. The chloramphenicol resistance determinant of a plasmid integrated at the subtilisin locus was mapped by PBS1 transduction and was found to be linked to glyB (83%) and argC (60%), but not with metC or purB . The chromosomal locus containing the wild-type subtilisin allele was replaced with an in vitro-derived allele of the gene (delta apr-684) that contained a 684-base-pair deletion. The technique used for introducing the deletion is a variation of the gene replacement methods used in Saccharomyces cerevisiae and Escherichia coli. When used in B. subtilis, deletion mutants could be directly screened among the transformants. Physiological characterization of the delta apr-684 mutation revealed no discernable effect on the formation of heat-resistant endospores, but strains carrying the mutation produced only 10% of wild-type serine protease activity. A model is presented that outlines the pathway for plasmid integration and deletion formation in B. subtilis.  相似文献   

17.
An extracellular serine protease produced by a mutant, M1, derived from Streptomyces albogriseolus S-3253 that no longer produces a protease inhibitor (Streptomyces subtilisin inhibitor [SSI]) was isolated. A 20-kDa protein was purified by its affinity for SSI and designated SAM-P20. The amino acid sequence of the amino-terminal region of SAM-P20 revealed high homology with the sequences of Streptomyces griseus proteases A and B, and the gene sequence confirmed the relationships. The sequence also revealed a putative amino acid signal sequence for SAM-P20 that apparently functioned to allow secretion of SAM-P20 from Escherichia coli carrying the recombinant gene. SAM-P20 produced by E. coli cells was shown to be sensitive to SSI inhibition.  相似文献   

18.
Previously, we isolated a candidate for an endogenous target enzyme(s) of the Streptomyces subtilisin inhibitor (SSI), termed SAM-P20, from a non-SSI-producing mutant strain (S. Taguchi, A. Odaka, Y. Watanabe, and H. Momose, Appl. Environ. Microbiol. 61:180-186, 1995). In this study, in order to investigate the detailed enzymatic properties of this protease, an overproduction system of recombinant SAM-P20 was established in Streptomyces coelicolor with the SSI gene promoter. The recombinant SAM-P20 was purified by salting out and by two successive ion-exchange chromatographies to give a homogeneous band by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Partial peptide mapping and amino acid composition analysis revealed that the recombinant SAM-P20 was identical to natural SAM-P20. From the results for substrate specificity and inhibitor sensitivity, SAM-P20 could be categorized as a chymotrypsin-like protease with an arginine-cleavable activity, i.e., a serine protease with broad substrate specificity. For proteolytic activity, the optimal pH was 10.0 and the optimal temperature was shifted from 50 to 80 degrees C by the addition of 10 mM calcium ion. The strong stoichiometric inhibition of SAM-P20 activity by SSI dimer protein occurred in a subunit molar ratio of these two proteins of about 1, and an inhibitor constant of SSI toward SAM-P20 was estimated to be 8.0 x 10(-10) M. The complex formation of SAM-P20 and SSI was monitored by analytical gel filtration, and a complex composed of two molecules of SAM-P20 and one dimer molecule of SSI was detected, in addition to a complex of one molecule of SAM-P20 bound to one dimer molecule of SSI. The reactive site of SSI toward SAM-P20 was identified as Met-73-Val-74 by sequence analysis of the modified form of SSI, which was produced by the acidification of the complex of SSI and SAM-P20. This reactive site is the same that toward an exogenous target enzyme, subtilisin BPN'.  相似文献   

19.
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   

20.
With the aim to distinguish between local and global conformational changes induced by trifluoroethanol in RNase A, spectroscopic and activity measurements in combination with proteolysis by unspecific proteases have been exploited for probing structural transitions of RNase A as a function of trifluoroethanol concentration. At > 30% (v/v) trifluoroethanol (pH 8.0; 25 degrees C), circular dichroism and fluorescence spectroscopy indicate a cooperative collapse of the tertiary structure of RNase A coinciding with the loss of its enzymatic activity. In contrast to the denaturation by guanidine hydrochloride, urea or temperature, the breakdown of the tertiary structure in trifluoroethanol is accompanied by an induction of secondary structure as detected by far-UV circular dichroism spectroscopy. Proteolysis with the nonspecific proteases subtilisin Carlsberg or proteinase K, both of which attack native RNase A at the Ala20-Ser21 peptide bond, yields refined information on conformational changes, particularly in the pretransition region. While trifluoroethanol at concentrations > 40% results in a strong increase of the rate of proteolysis and new primary cleavage sites (Tyr76-Ser77, Met79-Ser80) were identified, the rate of proteolysis at trifluoroethanol concentrations < 40% (v/v) is much smaller (up to two orders of magnitude) than that of the native RNase A. The proteolysis data point to a decreased flexibility in the surrounding of the Ala20-Ser21 peptide bond, which we attribute to subtle conformational changes of the ribonuclease A molecule. These changes, however, are too marginal to alter the overall catalytic and spectroscopic properties of ribonuclease A.  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