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.
固定化分子伴侣GroE促进变性溶菌酶复性的研究   总被引:10,自引:0,他引:10  
董晓燕  杨晖  甘一如  白姝  孙彦   《生物工程学报》2000,16(2):169-172
利用重组大肠杆菌表达制备了分子伴侣GroE(GroEL和GroES),研究了GroE以及GroEL辅助变性溶菌酶复性的作用。结果表明,不仅游离GroEL单独作用可使溶菌酶复性收率达到90%以上,而且固定化GroEL亦可有效地促进蛋白质复性,最佳复性温度为37℃,最佳pH值范围为6~8,复性酶的活性收率在85%以上。另外,固定化GroEL可反复回收利用,表明固定化GroEL有可能在实际生物下游过程中得到应用。  相似文献   

2.
Chaperonins cpn60/cpn10 (GroEL/GroES in Escherichia coli) assist folding of nonnative polypeptides. Folding of the chaperonins themselves is distinct in that it entails assembly of a sevenfold symmetrical structure. We have characterized denaturation and renaturation of the recombinant human chaperonin 10 (cpn10), which forms a heptamer. Denaturation induced by chemical denaturants urea and guanidine hydrochloride (GuHCl) as well as by heat was monitored by tyrosine fluorescence, far-ultraviolet circular dichroism, and cross-linking; all denaturation reactions were reversible. GuHCl-induced denaturation was found to be cpn10 concentration dependent, in accord with a native heptamer to denatured monomer transition. In contrast, urea-induced denaturation was not cpn10 concentration dependent, suggesting that under these conditions cpn10 heptamers denature without dissociation. There were no indications of equilibrium intermediates, such as folded monomers, in either denaturant. The different cpn10 denatured states observed in high [GuHCl] and high [urea] were supported by cross-linking experiments. Thermal denaturation revealed that monomer and heptamer reactions display the same enthalpy change (per monomer), whereas the entropy-increase is significantly larger for the heptamer. A thermodynamic cycle for oligomeric cpn10, combining chemical denaturation with the dissociation constant in absence of denaturant, shows that dissociated monomers are only marginally stable (3 kJ/mol). The thermodynamics for co-chaperonin stability appears conserved; therefore, instability of the monomer could be necessary to specify the native heptameric structure.  相似文献   

3.
During refolding and reassociation of chemically denatured non-glycosylated invertase from Saccharomyces cerevisiae, aggregation competes with correct folding, leading to low yields of reactivation (Kern et al. (1992) Protein Sci. 1, 120-131). In the presence of the chaperone GroEL, refolding is completely arrested. This suggests the formation of a stable complex between GroEL and non-native non-glycosylated invertase. Addition of MgATP results in a slow release of active invertase from the chaperone complex. When GroEL/ES and MgATP are present during refolding, the final reactivation yield increases from 14% to 36%. In contrast, refolding of the core-glycosylated and the high-mannose glycosylated forms of invertase is not arrested by GroEL. Only a short lag phase at the beginning of reactivation and a slightly increased reactivation yield (64% to 86% for core-glycosylated and 62% to 76% for external invertase) indicate a weak interaction of the glycosylated forms with the chaperone.  相似文献   

4.
Tyagi NK  Fenton WA  Deniz AA  Horwich AL 《FEBS letters》2011,585(12):1969-1972
Under "permissive" conditions at 25°C, the chaperonin substrate protein DM-MBP refolds 5-10 times more rapidly in the GroEL/GroES folding chamber than in free solution. This has been suggested to indicate that the chaperonin accelerates polypeptide folding by entropic effects of close confinement. Here, using native-purified DM-MBP, we show that the different rates of refolding are due to reversible aggregation of DM-MBP while folding free in solution, slowing its kinetics of renaturation: the protein exhibited concentration-dependent refolding in solution, with aggregation directly observed by dynamic light scattering. When refolded in chloride-free buffer, however, dynamic light scattering was eliminated, refolding became concentration-independent, and the rate of refolding became the same as that in GroEL/GroES. The GroEL/GroES chamber thus appears to function passively toward DM-MBP.  相似文献   

5.
Protein aggregation is commonly observed during protein refolding. To better understand this phenomenon, the intermolecular interactions experienced by a protein during unfolding and refolding are inferred from second virial coefficient (SVC) measurements. It is accepted that a negative SVC is indicative of protein-protein interactions that are attractive, whereas a positive SVC indicates net repulsive interactions. Lysozyme denatured and reduced in guanidinium hydrochloride exhibited a decreasing SVC as the denaturant was diluted, and the SVC approached zero at approximately 3 M GdnHCl. Further dilution of denaturant to renaturation conditions (1.25 M GdnHCl) led to a negative SVC, and significant protein aggregation was observed. The inclusion of 500 mM L-arginine in the renaturation buffer shifted the SVC to positive and suppressed aggregation, thereby increasing refolding yield. The formation of mixed disulfides in the denatured state prior to refolding also increased protein solubility and suppressed aggregation, even without the use of L-arginine. Again, the suppression of aggregation was shown to be caused by a shift from attractive to repulsive intermolecular interactions as reflected in a shift from a negative to a positive SVC value. To the best of our knowledge, this is the first time that SVC data have been reported for renaturation studies. We believe this technique will aid in our understanding of how certain conditions promote renaturation and increase protein solubility, thereby suppressing aggregation. SVC measurements provide a useful link, for protein folding and aggregation, between empirical observation and thermodynamics.  相似文献   

6.
To compare folding/assembly processes of heptameric co-chaperonin proteins 10 (cpn10) from different species and search for the origin of thermostability in hyper-thermostable Aquifex aeolicus cpn10 (Aacpn10), we have studied two bacterial variants-Aacpn10 and Escherichia coli cpn10 (GroES)-and compared the results to data on Homo sapiens cpn10 (hmcpn10). Equilibrium denaturation of GroES by urea, guanidine hydrochloride (GuHCl) and temperature results in coupled heptamer-to-monomer transitions in all cases. This is similar to the behavior of Aacpn10 but differs from hmcpn10 denaturation in urea. Time-resolved experiments reveal that GroES unfolds before heptamer dissociation, whereas refolding/reassembly begins with folding of individual monomers; these assemble in a slower step. The sequential folding/assembly mechanism for GroES is rather similar to that observed for Aacpn10 but contradicts the parallel paths of hmcpn10. We reveal that Aacpn10's stability profile is shifted upwards, broadened, and also moved horizontally to higher temperatures, as compared to that of GroES.  相似文献   

7.
When Bacillus stearothermophilus LDH dimer is incubated with increasing concentrations of the denaturant guanidinium chloride, three distinct unfolded states of the molecule are observed at equilibrium [Smith, C. J., et al. (1991) Biochemistry 30, 1028-1036]. The kinetics of LDH refolding are consistent with an unbranched progression through these states. The Escherichia coli chaperonin, GroEL, binds with high affinity to the completely denatured form and more weakly to the earliest folding intermediate, thus retarding the refolding process. A later structurally defined folding intermediate, corresponding to a molten globule form, is not bound by GroEL; neither is the inactive monomer. The complex between GroEL and denatured LDH is destabilized by the binding of magnesium/ATP (Mg/ATP) or by the nonhydrolyzable analogue adenylyl imidodiphosphate (AMP-PNP). From our initial kinetic data, we propose that GroEL exists in two interconvertible forms, one of which is stabilized by the binding of Mg/ATP but associates weakly with the unfolded protein. The other is destabilized by Mg/ATP and associates strongly with unfolded LDH. The relevance of these findings to the role of GroEL in vivo is discussed.  相似文献   

8.
Recombinant human macrophage colony-stimulating factor (rhM-CSF), a homodimeric, disulfide bonded protein, was expressed in Escherichia coli in the form of inclusion bodies. Reduced and denatured rhM-CSF monomers were refolded in the presence of a thiol mixture (reduced and oxidized glutathione) and a low concentration of denaturing agent (urea or guanidinium chloride). Refolding was monitored by nonreducing gel electrophoresis and recovery of bioactivity. The effects of denaturant type and concentration, protein concentration, concentration of thiol/disulfide reagents, temperature, and presence of impurities on the kinetics of rhM-CSF renaturation were investigated. Low denaturant concentrations (<0.5 M urea) and high protein concentrations (>0.4 mg/ml) in the refolding mixture resulted in increased formation of aggregates, although aggregation was never significant even when refolding was carried out at room temperature. Higher protein concentration resulted in higher rates but did not lead to increased yields, due to the formation of unwanted aggregates. Experiments conducted at room temperature resulted in slightly higher rates than those conducted at 4 degrees C. Although the initial renaturation rate for solubilized inclusion body protein without purification was higher than that of the reversed-phase purified reduced denatured rhM-CSF, the final renaturation yield was much higher for the purified material. A maximum refolding yield of 95% was obtained for the purified material at the following refolding conditions: 0.5 M urea, 50 mM Tris, 1.25 mM DTT, 2 mM GSH, 2 mM GSSG, 22 degrees C, pH 8, [protein] = 0.13 mg/ml.  相似文献   

9.
Chaperonins cpn60 and cpn10 are essential proteins involved in cellular protein folding. Plant chloroplasts contain a unique version of the cpn10 co-chaperonin, cpn20, which consists of two homologous cpn10-like domains (N-cpn20 and C-cpn20) that are connected by a short linker region. Although cpn20 seems to function like other single domain cpn10 oligomers, the structure and specific functions of the domains are not understood. We mutated amino acids in the "mobile loop" regions of N-cpn20, C-cpn20 or both: a highly conserved glycine, which was shown to be important for flexibility of the mobile loop, and a leucine residue shown to be involved in binding of co-chaperonin to chaperonin. The mutant proteins were purified and their oligomeric structure validated by gel filtration, native gel electrophoresis, and circular dichroism. Functional assays of protein refolding and inhibition of GroEL ATPase both showed (i) mutation of the conserved glycine reduced the activity of cpn20, whether in N-cpn20 (G32A) or C-cpn20 (G130A). The same mutation in the bacterial cpn10 (GroES G24A) had no effect on activity. (ii) Mutations in the highly conserved leucine of N-cpn20 (L35A) and in the corresponding L27A of GroES resulted in inactive protein. (iii) In contrast, mutant L133A, in which the conserved leucine of C-cpn20 was altered, retained 55% activity. We conclude that the structure of cpn20 is much more sensitive to alterations in the mobile loop than is the structure of GroES. Moreover, only N-cpn20 is necessary for activity of cpn20. However, full and efficient functioning requires both domains.  相似文献   

10.
Two new 2-D crystal forms of the Escherichia coli chaperone GroEL (cpn60) 2 × 7-mer have been produced using the negative staining-carbon film (NS-CF) technique. These 2-D crystals, which contain the cylindrical GroEL in side-on and end-on orientations, both possess p21 symmetry, with two molecules in the respective unit cells. The crystallographically averaged images correlate well with those obtained by other authors from single particle analysis of GroEL and our own previous crystallographic analysis. 2-D crystallization of the smaller chaperone GroES (cpn10) 7-mer has also been achieved using the NS-CF technique. Crystallographically averaged images of GroES single particle images indicate considerable variation in molecular shape, which is most likely due to varying molecular orientation on the carbon support film. The quaternary structure of GroES does, nevertheless, approximate to a ring-like shape. The complex formed by GroEL and GroES in the presence of ATP at room temperature has been shown to possess a symmetrical hollow ellipsoidal conformation. This symmetrical complex forms in the presence of a 2:1 or greater molar ratio of GroES:GroEL. At lower molar ratios linear chains of GroEL form, apparently linked by GroES in a 1:1 manner, which provide supportive evidence for the ability of both ends of the GroEL cylinder to interact with GroES. The apparent discrepancy between our data and that of other groups who have described an asymmetrical "bullet-shaped" (holo-chaperone) GroEL/ES complex is discussed in detail.  相似文献   

11.
Many bacteria possess 2 or more genes for the chaperonin GroEL and the cochaperonin GroES. In particular, rhizobial species often have multiple groEL and groES genes, with a high degree of amino-acid similarity, in their genomes. The Rhizobium leguminosarum strain A34 has 3 complete groE operons, which we have named cpn.1, cpn.2 and cpn.3. Previously we have shown the cpn. 1 operon to be essential for growth, but the two other cpn operons to be dispensable. Here, we have investigated the extent to which loss of the essential GroEL homologue Cpn60.1 can be compensated for by expression of the other two GroEL homologues (Cnp60.2 and Cpn60.3). Cpn60.2 could not be overexpressed to high levels in R. leguminosarum, and was unable to replace Cpn60.1. A strain that overexpressed Cpn60.3 grew in the absence of Cpn60.1, but the complemented strain displayed a temperature-sensitive phenotype. Cpn60.1 and Cpn60.3, when coexpressed in Escherichia coli, preferentially selfassembled rather than forming mixed heteroligomers. We conclude that, despite their high amino acid similarity, the GroEL homologues of R. leguminosarum are not functionally equivalent in vivo.  相似文献   

12.
Our understanding of mechanisms for GroEL/GroES-assisted protein folding to date has been derived mostly from studies with small proteins. Little is known concerning the interaction of these chaperonins with large multidomain polypeptides during folding. In the present study, we investigated chaperonin-dependent folding of a large 86-kDa fusion polypeptide, in which the mature maltose-binding protein (MBP) sequence was linked to the N terminus of the alpha subunit of the decarboxylase (E1) component of the human mitochondrial branched-chain alpha-ketoacid dehydrogenase complex. The fusion polypeptide, MBP-alpha, when co-expressed with the beta subunit of E1, produced a chimeric protein MBP-E1 with an (MBP-alpha)2beta2 structure, similar to the alpha2 beta2 structure in native E1. Reactivation of MBP-E1 denatured in 8 M urea was absolutely dependent on GroEL/GroES and Mg2+-ATP, and exhibited strikingly slow kinetics with a rate constant of 376 M-1 s-1, analogous to denatured untagged E1. Chaperonin-mediated refolding of the MBP-alpha fusion polypeptide showed that the folding of the MBP moiety was about 7-fold faster than that of the alpha moiety on the same chain with rate constants of 1.9 x 10(-3) s-1 and 2.95 x 10(-4) s-1, respectively. This explained the occurrence of an MBP-alpha. GroEL binary complex that was isolated with amylose resin from the refolding mixture and transformed Escherichia coli lysates. The data support the thesis that distinct functional sequences in a large polypeptide exhibit different folding characteristics on the same GroEL scaffold. Moreover, we show that when the alpha.GroEL complex (molar ratio 1:1) was incubated with GroES, the latter was capable of capping either the very ring that harbored the 48-kDa (His)6-alpha polypeptide (in cis) or the opposite unoccupied cavity (in trans). In contrast, the MBP-alpha.GroEL (1:1) complex was capped by GroES exclusively in the trans configuration. These findings suggest that the productive folding of a large multidomain polypeptide can only occur in the GroEL cavity that is not sequestered by GroES.  相似文献   

13.
In Escherichia coli cells expressing 6-hydroxy-D-nicotine oxidase (6-HDNO), a flavoprotein with covalently bound FAD, approximately 40% of the polypeptide is in its apoform. We investigated whether in vivo holoenzyme formation was influenced by the association of the apoenzyme with cellular chaperones. Immunoprecipitation of apoenzyme-containing cell extract with protein-A-Sepharose-bound 6-HDNO- or GroEL-specific antibodies failed to reveal the formation of complexes between these proteins. The limiting factor in holoenzyme formation in vivo appeared to be the intracellular supply of phosphorylated tricarbon compounds (e.g. glycerol-3-P) acting as allosteric effectors in the flavinylation reaction. When holoenzyme formation from purified apo6-HDNO was investigated in vitro, addition of GroEL and GroES to the reaction assays increased the yield of holoenzyme formation. The observed increase in apoenzyme to holoenzyme transition was ATP independent, and the effect of GroE could be simulated by high concentrations of glycerol (40%). Apparently, a nonspecific protein-protein interaction between the GroE proteins and the apo6-HDNO favored holoenzyme formation. The refolding of guanidinium hydrochloride-unfolded holoenzyme, however, was catalyzed by GroEL and GroES in an ATP-dependent reaction. Recovery of the native, enzymatically active, conformation ranged from 30 to 40%. When apo6-HDNO was denatured and refolded, the same dependence on GroE and ATP was observed in the recovery of a conformation able to incorporate FAD and to holoenzyme. [14C] FAD in the refolding assay yielded radioactively labeled 6-HDNO demonstrating the autocatalytical covalent incorporation of FAD into the polypeptide during the folding process.  相似文献   

14.
Kinetic model of lysozyme renaturation with the molecular chaperone GroEL   总被引:1,自引:0,他引:1  
From the renaturation kinetics of denatured/reduced lysozyme assisted by the molecular chaperone GroEL, a simplified kinetic model was established based on the competition between protein folding and aggregation. In the presence of GroEL and ATP, the aggregate formation was a second order reaction. With 2 mM ATP, a renaturation yield of 90% at a high renaturation rate was obtained when the molar ratio of GroEL to lysozyme was 1:1.  相似文献   

15.
Molecular chaperones are a special class of heat shock proteins (Hsp) that assist the folding and formation of the quaternary structure of other proteins both in vivo and in vitro. However, some chaperones are complex oligomeric proteins, and one of the intriguing questions is how the chaperones fold. The representatives of the Escherichia coli chaperone system GroEL (Hsp60) and GroES (Hsp10) have been studied most intensively. GroEL consists of 14 identical subunits combined into two interacting ring-like structures of seven subunits each, while the co-chaperone GroES interacting with GroEL consists of seven identical subunits combined into a dome-like oligomeric structure. In spite of their complex quaternary structure, GroEL and GroES fold well both in vivo and in vitro. However, the specific oligomerization of GroEL subunits is dependent on ligands and external conditions. This review analyzes the literature and our own data on the study of unfolding (denaturation) and refolding (renaturation) processes of these molecular chaperones and the effect of ligands and solvent composition. Such analysis seems to be useful for understanding the folding mechanism not only of the GroEL/GroES complex, but also of other oligomeric protein complexes.  相似文献   

16.
Bacteriophage T4-encoded Gp31 is a functional ortholog of the Escherichia coli GroES cochaperonin protein. Both of these proteins form transient, productive complexes with the GroEL chaperonin, required for protein folding and other related functions in the cell. However, Gp31 is specifically required, in conjunction with GroEL, for the correct folding of Gp23, the major capsid protein of T4. To better understand the interaction between GroEL and its cochaperonin cognates, we determined whether the so-called "pseudo-T-even bacteriophages" are dependent on host GroEL function and whether they also encode their own cochaperonin. Here, we report the isolation of an allele-specific mutation of bacteriophage RB49, called epsilon22, which permits growth on the E. coli groEL44 mutant but not on the isogenic wild type host. RB49 epsilon22 was used in marker rescue experiments to identify the corresponding wild type gene, which we have named cocO (cochaperonin cognate). CocO has extremely limited identity to GroES but is 34% identical and 55% similar at the protein sequence level to T4 Gp31, sharing all of the structural features of Gp31 that distinguish it from GroES. CocO can substitute for Gp31 in T4 growth and also suppresses the temperature-sensitive phenotype of the E. coli groES42 mutant. CocO's predicted mobile loop is one residue longer than that of Gp31, with the epsilon22 mutation resulting in a Q36R substitution in this extra residue. Both the CocO wild type and epsilon22 proteins have been purified and shown in vitro to assist GroEL in the refolding of denatured citrate synthase.  相似文献   

17.
Paul S  Punam S  Chaudhuri TK 《The FEBS journal》2007,274(22):6000-6010
In vitro refolding of maltodextrin glucosidase, a 69 kDa monomeric Escherichia coli protein, was studied in the presence of glycerol, dimethylsulfoxide, trimethylamine-N-oxide, ethylene glycol, trehalose, proline and chaperonins GroEL and GroES. Different osmolytes, namely proline, glycerol, trimethylamine-N-oxide and dimethylsulfoxide, also known as chemical chaperones, assist in protein folding through effective inhibition of the aggregation process. In the present study, it was observed that a few chemical chaperones effectively reduced the aggregation process of maltodextrin glucosidase and hence the in vitro refolding was substantially enhanced, with ethylene glycol being the exception. Although, the highest recovery of active maltodextrin glucosidase was achieved through the ATP-mediated GroEL/GroES-assisted refolding of denatured protein, the yield of correctly folded protein from glycerol- or proline-assisted spontaneous refolding process was closer to the chaperonin-assisted refolding. It was also observed that the combined application of chemical chaperones and molecular chaperone was more productive than their individual contribution towards the in vitro refolding of maltodextrin glucosidase. The chemical chaperones, except ethylene glycol, were found to provide different degrees of protection to maltodextrin glucosidase from thermal denaturation, whereas proline caused the highest protection. The observations from the present studies conclusively demonstrate that chemical or molecular chaperones, or the combination of both chaperones, could be used in the efficient refolding of recombinant E. coli maltodextrin glucosidase, which enhances the possibility of identifying or designing suitable small molecules that can act as chemical chaperones in the efficient refolding of various aggregate-prone proteins of commercial and medical importance.  相似文献   

18.
Strong aggregation occurred in the refolding route of arginine kinase (AK) denatured with 3 mol GdnHCl/L (GdnHCl, guanidine hydrochloride). The activity recovery of GdnHCl-denatured AK was very low and dependent on the protein concentration in the process of refolding. For denatured AK at 1.2 micromol/L concentration, the recovered activity yield was about 45.2% of the native enzyme, whereas at 5.2 micromol/L the activity recovery yield was only 20% of native activity. The nonionic detergent Triton X-100 and Tween 20 (< or = 100 mmol/L concentration) not only effectively blocked the aggregation but also enabled the denatured AK to recover most of its native activity. The kinetics of aggregate solubilization showed that there was an induction phase dependent on the detergent, but there was no dependency when detergent was absent. The apparent activity recovery had a cooperative relation with detergents in the process of refolding, which suggested the existence of some interaction between the detergent and the refolding intermediate. On the basis of the study results, a scheme of refolding was proposed.  相似文献   

19.
Two proteins belonging to the group I chaperonin family were isolated from an obligate methanotroph, Methylobacillus glycogenes. The two proteins, one a GroEL homologue (cpn60: M. glycogenes 60 kDa chaperonin) and the other a GroES homologue (cpn10: M. glycogenes 10 kDa chaperonin), composed a heteropolymeric complex in the presence of ATP. Both proteins were purified from crude extracts of M. glycogenes by anion-exchange (DEAE-Toyopearl) and gel-filtration (Sephacryl S-400) chromatography. The native molecular weights of each chaperonin protein as determined by high-performance liquid chromatography (HPLC) gel-filtration were 820 000 for cpn60 and 65 000 for cpnl0. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed that the subunit molecular weights of cpn60 and cpnl0 were 58 000 and 10 000, respectively. Both cpn60 and cpnl0 possessed amino acid sequences which were highly homologous to other group I chaperonins. M. glycogenes cpn60 displayed an ATPase activity which was inhibited in the presence of cpn10. The chaperonins also displayed an ability to interact with and facilitate the refolding of Thermus malate dehydrogenase and yeast enolase in a manner similar to that of GroEL/ES. The similarities between the Escherichia coli GroE proteins are discussed.  相似文献   

20.
Molecular chaperones are involved in protein folding both in vivo and in vitro. The Escherichia coli chaperone GroEL interacts with a number of nonnative proteins. A common structural motif of nonnative proteins, which is recognized by GroEL, has not yet been identified. In order to study the role of beta-sheet secondary structure on the interaction of nonnative proteins with GroEL, we used the F(ab) fragment of a monoclonal antibody as a model substrate protein. Here we show that GroEL interacts functionally with this all-beta-protein during reactivation. Antibody fragments refold spontaneously in good yield from the guanidine-denatured state. Functional refolding to the native state is inhibited transiently by GroEL, but there is no complete folding arrest in the absence of Mg-ATP and GroES. The yield of these unspecifically released GroEL-bound F(ab) fragments corresponds to that of the spontaneous reactivation in the absence of chaperones. However, the refolding kinetics in the presence of GroEL are considerably slower. The addition of Mg-ATP to the GroEL.F(ab) complex results in an immediate release of bound substrate protein and a significant increase in the amount of reconstituted antibody fragments compared to spontaneous reactivation. GroES is not essential for functional GroEL-mediated refolding of the F(ab) fragment but affects the reactivation yield to a small extent. Interestingly, stimulation of the GroEL-mediated F(ab) refolding depends primarily on the binding and not on hydrolysis of adenosine triphosphates. Previous results indicate the binding of alpha-helices to GroEL. The results presented in this paper suggest that beta-sheet secondary structural elements are recognized by GroEL. We therefore conclude that the interaction of a nonnative protein with GroEL depends mainly on the nature of the early folding intermediate but not on a specific element of secondary structure.  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