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.
Alcaligenes sp. MTCC 10675 has been isolated from soil sample using enrichment method and has nitrilase catalytic system which is highly specific for the hydrolysis of arylaliphatic nitriles. Optimization of culture conditions using response surface methodology and inducer-mediated approach enhanced arylacetonitrilase production significantly (2.4-fold). Isobutyronitrile acted as an effective inducer for the induction of arylacetonitrilase, and it is highly specific for arylacetonitriles (phenyl acetonitrile and mandelonitrile). Arylacetonitrilase has no effect on its relative velocity (V r) up to 20 mM substrate (mandelonitrile) concentration and at 30 mM mandelonitrile, 23.4 % degree of inhibition (I d) was recorded. Half life of arylacetonitrilase of Alcaligenes sp. MTCC 10675 was 27.5 h at 25 °C. Hg2+, Ag+, Pb3+, and Co2+ were strong inhibitor of arylacetonitrilase activity which resulted into 100 %, 91 %, 84 %, and 83 % inhibition, respectively. Polar protic solvent (dichloromethane, dimethylsulphooxide, and n-butanol) reduce arylacetonitrilase activity up to 80–94 % at 10 % concentration. Alcaligenes sp. MTCC 10675 has higher biocatalytic activity, i.e., 3.9 gg-1 dcw, which is highest in comparison to till reported organism. Arylacetonitrilase-mediated hydrolysis of racemic mandelonitrile resulted into R-(-) mandelic acid with 99.0 % enantiomeric excess (e.e.)  相似文献   

2.
The highly enantioselective arylacetonitrilase of Pseudomonas putida was purified to homogeneity using a combination of (NH4)2SO4 fractionation and different chromatographic techniques. The enzyme has a molecular weight of 412 kDa and consisted of approximately nine to ten identical subunits (43 kDa). The purified enzyme exhibited a pH optimum of 7.0 and temperature optimum of 40°C. The nitrilase was highly susceptible to thiol-specific reagents and metal ions and also required a reducing environment for its activity. These reflected the presence of a catalytically essential thiol group for enzyme activity which is in accordance with the proposed mechanism for nitrilase-catalyzed reaction. The enzyme was highly specific for arylacetonitriles with phenylacetonitrile and its derivatives being the most preferred substrates. Higher specificity constant (k cat/K m) values for phenylacetonitrile compared to mandelonitrile also revealed the same. Faster reaction rate achieved with this nitrilase for mandelonitrile hydrolysis was possibly due to the low activation energy required by the protein. Incorporation of low concentration (<5%) of organic solvent increased the enzyme activity by increasing the availability of the substrate. Higher stability of the enzyme at slightly alkaline pH and ambient temperature provides an excellent opportunity to establish a dynamic kinetic resolution process for the production of (R)-(−)-mandelic acid from readily available mandelonitrile.  相似文献   

3.
Random mutagenesis of malate dehydrogenase from a thermophilic bacterium, Thermus flavus AT-62, had revealed that a Thr190----Ile replacement near the essential catalytic residue His187 caused marked modulation of the catalytic properties. For further exploration of a role of the residue at this position, this residue was substituted with each of the other amino acids by site-directed mutagenesis. Most of the mutations except for substitution with Ser caused increases in Km for oxaloacetate and increases Ki for oxaloacetate of 2-110 times. Substitution with His or Pro was characterized by the complete loss of substrate inhibition, along with a marked increase in Km for oxaloacetate. Kinetic analyses of the native and altered malate dehydrogenases at various pHs revealed that both Km and Ki for oxaloacetate decreased proportionally to the decrease in pH from 8.40 to 5.75, whereas kcat was nearly constant within the pH range. Apparent shifts of the optimum pH values toward acidity observed with most of the altered malate dehydrogenases were attributed to the increase in Ki, which facilitated the release from the substrate inhibition at a lower pH. Replacement of Thr310, a possible counterpart with which Thr190 forms a hydrogen bond, by Ile caused changes in the catalytic properties similar to those of the Thr190-substituted enzymes. These results suggest that not only the loss of the hydrogen bond between Thr190 and Thr310 but also properties of the residues introduced at position 190 cause modulation of the catalytic properties, probably through dislocation of the loop structure that contains the catalytic residue His187.  相似文献   

4.
Mycophenolic acid (MPA) is a potent natural product inhibitor of fungal and other eukaryotic inosine 5′‐monophosphate dehydrogenases (IMPDHs) originally isolated from spoiled corn silage. MPA is produced by the filamentous fungi Penicillium brevicompactum, which contains two IMPDHs, PbIMPDHA and PbIMPDHB, both of which are MPA‐resistant. The MPA binding sites of these enzymes are identical to MPA‐sensitive IMPDHs, so the structural determinants of resistance are unknown. Here we show that a single residue, Ser267, accounts for the MPA resistance of PbIMPDHA. Substitution of Ser267 with Ala, the residue most commonly found in this position in eukaryotic IMPDHs, makes PbIMPDHA sensitive to MPA. Conversely, Aspergillus nidulans IMPDH becomes MPA‐resistant when the analogous Ala residue is substituted with Ser. These substitutions have little effect on the catalytic cycles of either enzyme, suggesting the fitness costs are negligible despite the strong conservation of Ala at this position. Intriguingly, while only 1% of fungal IMPDHs contain Ser or Thr at position 267, these residues are found in the IMPDHs from several Aspergillus species that grow at the low temperatures also favored by Penicillium. Perhaps Ser/Thr267 is an evolutionary signature of MPA exposure.  相似文献   

5.
The nitrilase from Alcaligenes faecalis ECU0401 belongs to the category of arylacetonitrilase, which could hydrolyze 2-chloromandelonitrile, 3,4-dimethoxyphenylacetonitrile, mandelonitrile, and phenylacetonitrile into the corresponding arylacetic acids. To overcome the permeability barrier and prepare whole cell biocatalysts with high activities, permeabilization of Alcaligenes faecalis ECU0401 in relation to nitrilase activity was optimized by using cetyltrimethylammonium bromide (CTAB) as permeabilizing agent. The nitrilase activity from Alcaligenes faecalis ECU0401 increased 4.5-fold when the cells were permeabilized with 0.3% (w/v) CTAB for 20 min at 25°C and pH 6.5. Consequently, almost all the mandelonitrile was consumed and converted to (R)-(−)-mandelic acid with greater than 99.9% enantiomeric excess (e.e.) by the CTAB-permeabilized cells. The permeability barrier has been significantly reduced in the hydrolysis of mandelonitrile by using CTAB-permeabilized cells and a dynamic resolution was successfully achieved, giving a 100% theoretical yield of (R)-(−)-mandelic acid. Efficient biocatalyst recycling was achieved as a result of cell immobilization in calcium alginate, with a product-to-biocatalyst ratio of 3.82 g (R)-(−)-mandelic acid g−1 dry cell weight (dcw) cell after 20 cycles of repeated use.  相似文献   

6.
Abstract

Hydroxynitrile lyases (HNLs) are sought-after, stereo-selective biocatalysts used in the agrochemical, pharmaceutical and fine chemical industries to produce cyanohydrin enantiomers. There are several approaches for the discovery of HNLs, most of which are methodologically demanding and not suitable for high-throughput. Bioprospecting studies to date have also been constrained/limited to commercialised plants or botanical gardens, leaving a vast majority of plant species untested for HNL activity or cyanogenesis. To increase the rate of discovery of HCN liberating plants, we devised a Feigl-Anger microfuge tube that is portable and capable of high throughput detection of naturally cyanogenic plants. A workflow suitable for detecting plant candidates containing extractable, novel HNLs was subsequently applied. In this study, we screened over 600 plants for cyanogenic activity as well as the ability to degrade racemic mandelonitrile. We detected 33 plants able to degrade racemic mandelonitrile, of which, 25 were identified to the species level. Six of these plants were found to be naturally cyanogenic. Protein extracts from 5 of the naturally cyanogenic plants retained the ability to degrade racemic mandelonitrile pointing to five yet undescribed enzymes in the species Achyranthes aspera, Davallia trichomonoides, Morus mesozygia, Polypodium aureum “Mandaianum”, and Thelypteris confluens. In contrast, although Acalypha glabrata was found to be naturally cyanogenic, the protein extract did not break down racemic mandelonitrile. Here, we used racemic mandelonitrile as substrate and detected enzymes with mandelonitrile lyase activity, however, any cyanohydrin could be used as part of the approach taken here to detect novel HNLs specific to the substrate utilised.  相似文献   

7.
Biphenyl dioxygenase (Bph Dox) catalyzes the initial dioxygenation step in the metabolism of biphenyl. The large subunit (BphA1) of Bph Dox plays a crucial role in the determination of the substrate specificity of biphenyl-related compounds including polychlorinated biphenyls (PCBs). Previously, the substitution of Asn at Thr-376 near the active-site iron in the BphA1 of Pseudomonas pseudoalcaligenes KF707 expanded the oxidation range and altered the regiospecificity of Bph Dox for PCBs. In this study, we replaced Thr-376 with Gly, Ser, Gln, Tyr, Val, Phe, Asp, and Lys and expressed these enzymes in Escherichia coli. Bph Dox mutants of Thr376Asn, Thr376Val, Thr376Phe, and Thr376Lys showed novel degradation activity for dibenzofuran, which is a poor substrate for KF707 Bph Dox. All active Bph Dox mutants showed altered regiospecificity with 2,2′-dichlorobiphenyl and 2,5,4′-trichlorobiphenyl. The Thr376Gly, Thr376Val, Thr376Phe, and Thr376Asp Bph Dox mutants introduced molecular oxygen at the 2,3 position of 2,2′-dichlorobiphenyl, forming 2-chloro-2′,3′-dihydroxybiphenyl with concomitant dechlorination. The Bph Dox mutants of Thr376Gly, Thr376Ser, Thr376Asp, and Thr376Lys attacked 2,5,4′-trichlorobiphenyl via both 2′,3′- and 3,4-dioxygenation activities. In particular, the Thr376Phe Bph Dox mutant exhibited enhanced and expanded degradation activities toward all of the compounds tested. Further site-directed mutation was induced to change the oxidizing character of KF707 Bph Dox to that of the Bph Dox of Burkholderia xenovorans LB400 by the substitution of two amino acids, Ile335Phe and Thr376Asn, near the active-site.Electronic supplementary material Supplementary material is available in the online version of this article at .  相似文献   

8.
The Gin residue at amino acid position 102 ofBacillus stearothermophilus lactate dehydrogenase was replaced with Ser, Thr, Tyr, or Phe to investigate the effect on substrate recognition. The Q102S and Q102T mutant enzymes were found to have a broader range of substrate specificity (measured byk cat/K m) than the wild-type enzyme. However, it is evident that either Ser or Thr at position 102 are of a size able to accommodate a wide variety of substrates in the active site and substrate specificity appears to rely largely on size discrimination in these mutants. The Q102F and Q102Y mutant enzymes have low catalytic efficiency and do not show this relaxed substrate specificity. However, their activities are restored by the presence of an aromatic substrate. All of the enzymes have a very low catalytic efficiency with branched chain aliphatic substrates.Abbreviations used BSLDH Bacillus stearothermophilus lactate dehydrogenase - FBP fructose-1,6-bisphosphate - HP hydroxypyruvate - KB ketobutyrate - KC ketocaproate - KV ketovalerate - MDH malate dehydrogenase - PP phenylpyruvate - PYR pyruvate - RBE relative binding energy  相似文献   

9.
The characteristic oxidation or reduction reaction mechanisms of short‐chain oxidoreductase (SCOR) enzymes involve a highly conserved Asp‐Ser‐Tyr‐Lys catalytic tetrad. The SCOR enzyme Q9HYA2 from the pathogenic bacterium Pseudomonas aeruginosa was recognized to possess an atypical catalytic tetrad composed of Lys118‐Ser146‐Thr159‐Arg163. Orthologs of Q9HYA2 containing the unusual catalytic tetrad along with conserved substrate and cofactor recognition residues were identified in 27 additional species, the majority of which are bacterial pathogens. However, this atypical catalytic tetrad was not represented within the Protein Data Bank. The crystal structures of unligated and NADPH‐complexed Q9HYA2 were determined at 2.3 Å resolution. Structural alignment to a polyketide ketoreductase (KR), a typical SCOR, demonstrated that Q9HYA2's Lys118, Ser146, and Arg163 superimposed upon the KR's catalytic Asp114, Ser144, and Lys161, respectively. However, only the backbone of Q9HYA2's Thr159 overlapped KR's catalytic Tyr157. The Thr159 hydroxyl in apo Q9HYA2 is poorly positioned for participating in catalysis. In the Q9HYA2–NADPH complex, the Thr159 side chain was modeled in two alternate rotamers, one of which is positioned to interact with other members of the tetrad and the bound cofactor. A chloride ion is bound at the position normally occupied by the catalytic tyrosine hydroxyl. The putative active site of Q9HYA2 contains a chemical moiety at each catalytically important position of a typical SCOR enzyme. This is the first observation of a SCOR protein with this alternate catalytic center that includes threonine replacing the catalytic tyrosine and an ion replacing the hydroxyl moiety of the catalytic tyrosine.  相似文献   

10.
Combinatorial libraries of the lid domain of Rhizopus oryzae lipase (ROL; Phe88Xaa, Ala91Xaa, Ile92Xaa) were displayed on the yeast cell surface using yeast cell-surface engineering. Among the 40,000 transformants in which ROL mutants were displayed on the yeast cell surface, ten clones showed clear halos on soybean oil-containing plates. Among these, some clones exhibited high activities toward fatty acid esters of fluorescein and contained non-polar amino acid residues in the mutated positions. Computer modeling of the mutants revealed that hydrophobic interactions between the substrates and amino acid residues in the open form of the lid might be critical for ROL activity. Based on these results, Thr93 and Asp94 were further combinatorially mutated. Among 6,000 transformants, the Thr93Thr, Asp94Ser and Thr93Ser, Asp94Ser transformants exhibited a significant shift in substrate specificity toward a short-chain substrate. Computer modeling of these mutants suggested that a unique oxyanion hole, which is composed of Thr85 Oγ and Ser94 Oγ, was formed and thus the substrate specificity was changed. Therefore, coupling combinatorial mutagenesis with the cell surface display of ROL could lead to the production of a unique ROL mutant.  相似文献   

11.
Using directed evolution and site‐directed mutagenesis, we have isolated a highly thermostable variant of Aspergillus niger glucoamylase (GA), designated CR2‐1 . CR2‐1 includes the previously described mutations Asn20Cys and Ala27Cys (forming a new disulfide bond), Ser30Pro, Thr62Ala, Ser119Pro, Gly137Ala, Thr290Ala, His391Tyr and Ser436Pro. In addition, CR2‐1 includes several new putative thermostable mutations, Val59Ala, Val88Ile, Ser211Pro, Asp293Ala, Thr390Ser, Tyr402Phe and Glu408Lys, identified by directed evolution. CR2‐1 GA has a catalytic efficiency (kcat/Km) at 35°C and a specific activity at 50°C similar to that of wild‐type GA. Irreversible inactivation tests indicated that CR2‐1 increases the free energy of thermoinactivation at 80°C by 10 kJ mol?1 compared with that of wild‐type GA. Thus, CR2‐1 is more thermostable (by 5 kJ mol?1 at 80°C) than the most thermostable A. niger GA variant previously described, THS8 . In addition, Val59Ala and Glu408Lys were shown to individually increase the thermostability in GA variants by 1 and 2 kJ mol?1, respectively, at 80°C.  相似文献   

12.
Ke Zhou 《Phytochemistry》2009,70(3):366-42847
Terpene synthases (TPS) require divalent metal ion co-factors, typically magnesium, that are bound by a canonical DDXXD motif, as well as a putative second, seemingly less well conserved and understood (N/D)DXX(S/T)XXXE motif. Given the role of the Ser/Thr side chain hydroxyl group in ligating one of the three catalytically requisite divalent metal ions and the loss of catalytic activity upon substitution with Ala, it is surprising that Gly is frequently found in this ‘middle’ position of the putative second divalent metal binding motif in plant TPS. Herein we report mutational investigation of this discrepancy in a model plant diterpene cyclase, abietadiene synthase from Abies grandis (AgAS). Substitution of the corresponding Thr in AgAS with Ser or Gly decreased catalytic activity much less than substitution with Ala. We speculate that the ability of Gly to partially restore activity relative to Ala substitution for Ser/Thr stems from the associated reduction in steric volume enabling a water molecule to substitute for the hydroxyl group from Ser/Thr, potentially in a divalent metal ion coordination sphere. In any case, our results are consistent with the observed conservation pattern for this putative second divalent metal ion binding motif in plant TPS.  相似文献   

13.
The maximum activity of Yersinia enterocolitica phytase (YeAPPA) occurs at pH 5.0 and 45 °C, and notably, its specific activity (3.28 ± 0.24 U mg−1) is 800-fold less than that of its Yersinia kristeensenii homolog (YkAPPA; 88% amino acid sequence identity). Sequence alignment and molecular modeling show that the arginine at position 79 (Arg79) in YeAPPA corresponding to Gly in YkAPPA as well as other histidine acid phosphatase (HAP) phytases is the only non-conserved residue near the catalytic site. To characterize the effects of the corresponding residue on the specific activities of HAP phytases, Escherichia coli EcAPPA, a well-characterized phytase with a known crystal structure, was selected for mutagenesis—its Gly73 was replaced with Arg, Asp, Glu, Ser, Thr, Leu, or Tyr. The results show that the specific activities of all of the corresponding EcAPPA mutants (17–2,400 U mg−1) were less than that of the wild-type phytase (3,524 U mg−1), and the activity levels were approximately proportional to the molecular volumes of the substituted residues’ side chains. Site-directed replacement of Arg79 in YeAPPA (corresponding to Gly73 of EcAPPA) with Ser, Leu, and Gly largely increased the specific activity, which further verified the key role of the residue at position 79 for determining phytase activity. Thus, a new determinant that influences the catalytic efficiency of HAP phytases has been identified.  相似文献   

14.
Catalysis of the thermostable alcohol dehydrogenase from Bacillus stearothermophilus is performed by a proton release system involving a zinc-bound water molecule, a hydroxyl group of Thr40 (threonine position at 40), and an imidazole ring of His43. Amino acid residues (Thr40 and His43) at the active center were substituted by Ser and Arg, respectively. Thr40Ser had a tendency toward lower activity for primary alcohols than the wild type enzyme. However, the mutant enzyme became more active for substrates with a larger side chain, such as 2-methyl-1-propanol and cyclohexanol. This phenomena might be explained by the fact that the methyl group of Thr40 was eliminated in Ser. His43Arg exhibited higher activity to primary alcohols (except 2-methyl-1-propanol) and acetaldehyde (as a reverse reaction) than the wild type, but little activity for secondary alcohols and ketones. The Km value for ethanol (Km-e) of His43Arg was fifty-fold larger than that of the wild type. The characteristics of these mutant enzymes are also discussed.  相似文献   

15.
Complete dose-response curves for recently identified members of the AKH/RPCH family (four decapeptides and six octapeptides) have been measured inLocusta migratoria monitoring the lipid-mobilishing activityin vivo. In addition, dose-response curves have been produced for two octapeptide analogues which have a combination of amino acids at position 2 and 3 not occurring in naturally found AKH members. In the decapeptide members changes at position 10 from Thr to Ser are well tolerated, but the combination of Ser at position 5 and 7 around the Pro6 residue results in lowered activity, and efficacy of only 70%. In the octapeptides a single Leu/Val exchange at position 2 does not change the potency, however Tyr or Ile at position 2 lead to at least 3-fold loss of activity. The Ser5-Pro6-Ser7 combination in an octapeptide, as in the decapeptide, reduces potency. Octapeptides with 3 aromatic amino acids (Phe2, Tyr4, Trp8) show no typical dose-response curve and have low efficacies. The combination of Val2-Thr3 which has never been found in an octapeptide is tolerated well, but Leu2-Val3 is not. The latter peptide is rather inactive and has a low efficacy; very likely because the hydrophilicity/hydrophobicity pattern at the N-terminus of the peptide is absent.  相似文献   

16.
Nucleotide pyrophosphatases/phosphodiesterases (NPPs, PF01663) release nucleoside 5′-monophosphates from a wide range of nucleotide substrates. Only very recently, the first plant members of the NPP family were characterised (Joye et al. J Cereal Sci 51: 326–336, 2010), and little is known about their substrate-specifying residues. We elucidated the role of six amino acid residues of the recently identified and characterised Triticum aestivum L. NPP (Joye et al. J Cereal Sci 51: 326–336, 2010). Substitution of the highly conserved catalytic Thr132 into Ser or Ala completely abolished enzyme activity. Mutation of a highly conserved His255 residue into an apolar Ala suprisingly increased enzyme activity against most phosphodiester substrates. Four other residues moderately to highly conserved over NPPs of different organisms were studied as well. Mutation of the Asn153, Asn165 and Glu199 into an Arg, Ser and Asp residue, respectively, increased the relative enzyme activity against p-nitrophenyl phosphate. Furthermore, mutation of Phe194 into Ser increased the relative enzyme activity against adenosine 5′-monophosphate-containing substrates, although the overall enzyme activity of this mutant enzyme decreased. We conclude that the structural requirements and the conservation of the amino acids of the catalytic site of TaNPPr and, by extension, probably of all NPPs, are very stringent.  相似文献   

17.
The structures of fully active cyclin-dependent kinase-2 (CDK2) complexed with ATP and peptide substrate, CDK2 after the catalytic reaction, and CDK2 inhibited by phosphorylation at Thr14/Tyr15 were studied using molecular dynamics (MD) simulations. The structural details of the CDK2 catalytic site and CDK2 substrate binding box were described. Comparison of MD simulations of inhibited complexes of CDK2 was used to help understand the role of inhibitory phosphorylation at Thr14/Tyr15. Phosphorylation at Thr14/Tyr15 causes ATP misalignment for the phosphate-group transfer, changes in the Mg2+ coordination sphere, and changes in the H-bond network formed by CDK2 catalytic residues (Asp127, Lys129, Asn132). The inhibitory phosphorylation causes the G-loop to shift from the ATP binding site, which leads to opening of the CDK2 substrate binding box, thus probably weakening substrate binding. All these effects explain the decrease in kinase activity observed after inhibitory phosphorylation at Thr14/Tyr15 in the G-loop. Interaction of the peptide substrate, and the phosphorylated peptide product, with CDK2 was also studied and compared. These results broaden hypotheses drawn from our previous MD studies as to why a basic residue (Arg/Lys) is preferred at the P+2 substrate position. Figure View of the substrate binding site of the fully active cyclin-dependent kinase-2 (CDK2) (pT160-CDK2/cyclin A/ATP). The pThr160 activation site is located in the T-loop (yellow secondary structure). The G-loop, which partly forms the ATP binding site, is shown in blue. The Thr14 and Tyr15 inhibitory phosphorylation sites located in the G-loop are shown in licorice representation  相似文献   

18.
tert-Butyl (3R,5S)-6-chloro-3,5-dihydroxyhexanoate ((3R,5S)-CDHH) is a key chiral intermediate for the side chain synthesis of rosuvastatin. In this study, random mutagenesis, site-saturation mutagenesis and combinatorial mutagenesis methods were applied to improve the activity of a synthesized stereoselective short chain carbonyl reductase (SCR) to prepare (3R,5S)-CDHH. After screened by high-throughput screening method and high-performance liquid chromatography, mut-Phe145Met/Thr152Ser and mut-Phe145Tyr/Thr152Ser, were obtained, and the enzyme activities of mutants were improved by 1.60- and 1.91-fold compared with parent enzyme, respectively. The catalytically efficiencies (kcat/Km) of mut-Phe145Met/Thr152Ser and mut-Phe145Tyr/Thr152Ser exhibited 5.11- and 8.07-fold improvements in initial activity toward (S)-6-chloro-5-hydroxy-3-oxohexanoate ((S)-CHOH), respectively. In the asymmetric reduction, mut-Phe145Tyr/Thr152Ser catalyzed 500 g L−1 of (S)-CHOH to produce (3R,5S)-CDHH with >99% yield and >99% e.e., and the highest space-time yield achieved at 752.76 mmol L−1 h−1 g−1 wet cell weight within 8 h bioconversion. This study provides a foundation for the preparation of (3R,5S)-CDHH by carbonyl reductase.  相似文献   

19.

Background  

Protein Ser/Thr phosphatase 5 (PP5) and its Saccharomyces cerevisiae homolog protein phosphatase T1 (Ppt1p) each contain an N-terminal domain consisting of several tetratricopeptide repeats (TPRs) and a C-terminal catalytic domain that is related to the catalytic subunits of protein phosphatases 1 and 2A, and calcineurin. Analysis of yeast Ppt1p could provide important clues to the function of PP5 and its homologs, however it has not yet been characterized at the biochemical or cellular level.  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