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 31 毫秒
1.
At the end of translation in bacteria, ribosome recycling factor (RRF) is used together with elongation factor G to recycle the 30S and 50S ribosomal subunits for the next round of translation. In x-ray crystal structures of RRF with the Escherichia coli 70S ribosome, RRF binds to the large ribosomal subunit in the cleft that contains the peptidyl transferase center. Upon binding of either E. coli or Thermus thermophilus RRF to the E. coli ribosome, the tip of ribosomal RNA helix 69 in the large subunit moves away from the small subunit toward RRF by 8 Å, thereby disrupting a key contact between the small and large ribosomal subunits termed bridge B2a. In the ribosome crystals, the ability of RRF to destabilize bridge B2a is influenced by crystal packing forces. Movement of helix 69 involves an ordered-to-disordered transition upon binding of RRF to the ribosome. The disruption of bridge B2a upon RRF binding to the ribosome seen in the present structures reveals one of the key roles that RRF plays in ribosome recycling, the dissociation of 70S ribosomes into subunits. The structures also reveal contacts between domain II of RRF and protein S12 in the 30S subunit that may also play a role in ribosome recycling.  相似文献   

2.
《Journal of molecular biology》2019,431(12):2343-2353
For Escherichia coli growing rapidly in rich medium at 37 °C, the doubling time can be as short as ~ 20 min and the average rate of translation (ktrl) can be as fast as ~ 20 amino acids/s. For slower growth arising from poor nutrient quality or from higher growth osmolality, ktrl decreases significantly. In earlier work from the Hwa lab, a simplified Michaelis–Menten model suggested that the decrease in ktrl arises from a shortage of ternary complexes (TCs) under nutrient limitation and from slower diffusion of TCs under high growth osmolality. Here we present a single-molecule tracking study of the diffusion of EF-Tu in E. coli growing with doubling times in the range 62–190 min at 37 °C due to nutrient limitation, high growth osmolality, or both. The diffusive properties of EF-Tu remain quantitatively indistinguishable across all growth conditions studied. Dissection of the total population into ribosome-bound and free sub-populations, combined with copy number estimates for EF-Tu and ribosomes, indicates that in all cases ~ 3.7 EF-Tu copies are bound on average to each translating 70S ribosome. Thus, the four L7/L12 binding sites adjacent to the ribosomal A-site in E. coli are essentially saturated with TCs in all conditions, facilitating rapid testing of aminoacyl-tRNAs for a codon match. Evidently, the average translation rate is not limited by either the supply of cognate TCs under nutrient limitation or by the diffusion of free TCs at high osmolality. Some other step or steps must be rate limiting for translation in slow growth.  相似文献   

3.
Summary The ultrastructure of Drosophila melanogaster cytoplasmic ribosomal subunits and monomers have been examined by electron microscopy. The Drosophila ribosomal structures are compared to those determined for other eucaryotes and E. coli. Negatively contrasted images of 60S subunits are seen in the most frequent view to be approximately round particles about 280 Å in diameter. About 35% of the particles present a single prominent projection which we call the 60S peak. The peak emanates from a flattened region of the 60S subunit. The maximum observed length of the 60S peak is approximately 90Å. The Drosophila 60S peak is highly reminiscent of the E. coli L7/L12 stalk. The Drosophila 40S subunit is an elongated, slightly bent particle which measures 280×170×160 Å. It bears a strong resemblance to small ribosomal subunits of other eucaryotes and is strikingly similar to the E. coli 30S subunit. Micrographs of 80S monomeric ribosomes show the long axis of the 40S to be parallel and in apparent contact with the flattened region of 60S subunit. The 60S peak appears to bisect the long axis of the 40S subunit. The 40S subunit seems to be oriented in the monomeric ribosome so that the 40S projection is toward the body of the large subunit. Comparison of our data with similar studies in different organisms indicates that the eucaryotic large ribosomal subunits exhibit morphological heterogeneity while the small subunits remain remarkably similar.  相似文献   

4.
The in vivo assembly of ribosomal subunits requires assistance by auxiliary proteins that are not part of mature ribosomes. More such assembly proteins have been identified for the assembly of the 50S than for the 30S ribosomal subunit. Here, we show that the RimP protein (formerly YhbC or P15a) is important for the maturation of the 30S subunit. A rimP deletion (ΔrimP135) mutant in Escherichia coli showed a temperature-sensitive growth phenotype as demonstrated by a 1.2-, 1.5-, and 2.5-fold lower growth rate at 30, 37, and 44 °C, respectively, compared to a wild-type strain. The mutant had a reduced amount of 70S ribosomes engaged in translation and showed a corresponding increase in the amount of free ribosomal subunits. In addition, the mutant showed a lower ratio of free 30S to 50S subunits as well as an accumulation of immature 16S rRNA compared to a wild-type strain, indicating a deficiency in the maturation of the 30S subunit. All of these effects were more pronounced at higher temperatures. RimP was found to be associated with free 30S subunits but not with free 50S subunits or with 70S ribosomes. The slow growth of the rimP deletion mutant was not suppressed by increased expression of any other known 30S maturation factor.  相似文献   

5.
Eukaryotic translation initiation factor 6 (eIF6) binds to the 60S ribosomal subunit and prevents its association with the 40S ribosomal subunit. The Saccharomyces cerevisiae gene that encodes the 245-amino-acid eIF6 (calculated Mr 25,550), designated TIF6, has been cloned and expressed in Escherichia coli. The purified recombinant protein prevents association between 40S and 60S ribosomal subunits to form 80S ribosomes. TIF6 is a single-copy gene that maps on chromosome XVI and is essential for cell growth. eIF6 expressed in yeast cells associates with free 60S ribosomal subunits but not with 80S monosomes or polysomal ribosomes, indicating that it is not a ribosomal protein. Depletion of eIF6 from yeast cells resulted in a decrease in the rate of protein synthesis, accumulation of half-mer polyribosomes, reduced levels of 60S ribosomal subunits resulting in the stoichiometric imbalance in the 40S/60S subunit ratio, and ultimately cessation of cell growth. Furthermore, lysates of yeast cells depleted of eIF6 remained active in translation of mRNAs in vitro. These results indicate that eIF6 does not act as a true translation initiation factor. Rather, the protein may be involved in the biogenesis and/or stability of 60S ribosomal subunits.  相似文献   

6.
Light activation of photosensitizing dyes in presence of molecular oxygen generates highly cytotoxic reactive oxygen species leading to cell inactivation. Nucleic acids are molecular targets of this photodynamic action but not considered the main cause of cell death. The in vivo effect of the photodynamic process on the intracellular nucleic acid content of Escherichia coli and Staphylococcus warneri was evaluated herein.Two cationic porphyrins (Tetra-Py+-Me and Tri-Py+-Me-PF) were used to photoinactivate E. coli (5.0 μM; 108 cells mL?1) and S. warneri (0.5 μM; 108 cells mL?1) upon white light irradiation at 4.0 mW cm?2 for 270 min and 40 min, respectively. Total nucleic acids were extracted from photosensitized bacteria after different times of irradiation and analyzed by agarose gel electrophoresis. The double-stranded DNA was quantified by fluorimetry and the porphyrin binding to bacteria was determined by spectrofluorimetry.E. coli was completely photoinactivated with both porphyrins (5.0 μM), whereas S. warneri was only completely inactivated by Tri-Py+-Me-PF (0.5 μM). The hierarchy of nucleic acid changes in E. coli was in the order: 23S rRNA > 16S rRNA > genomic DNA. The nucleic acids of S. warneri were extensively reduced after 5 min with Tri-Py+-Me-PF but almost unchanged with Tetra-Py+-Me after 40 min of irradiation. The amount of Tri-Py+-Me-PF bound to E. coli after washing the cells is higher than Tetra-Py+-Me and the opposite was observed for S. warneri. The binding capacity of the photosensitizers is not directly related to the PDI efficiency or nucleic acid reduction and this reduction occurs in parallel with the decrease of surviving cells.  相似文献   

7.
The genes encoding glycerol dehydratase were cloned and characterized by genomic DNA from Klebsiella pneumoniae XJPD-Li, and the assigned accession number EF634063 was available from the GenBank database. The DNA sequence analysis showed that the clone included three ORFs (dhaB, dhaC and dhaE, encoding α, β and γ subunit of glycerol dehydratase, respectively). Among three subunits of glycerol dehydratase, amino acid residues H13, S193, N359, E407, and M515 of α subunit, N47, L150, V189 of β subunit are different with what had been reported. Subsequently, the expression vector was constructed and transformed into E. coli BL21, and the colony carried genes of glycerol dehydratase were selected. SDS-PAGE examination showed that the three subunits were well expressed. The specific activity of recombined glycerol dehydratase reached to 0.299 U mg?1, which was about 3 times comparing with that of the wild strain. The research also displayed that both glycerol and O2 could inactive the glycerol dehydratase expressed in E. coli quickly in 10 min. The inactivated glycerol dehydratase could be effectively reactivated under the system as follows: the concentration of ATP, Mg2+ and coenzyme B12 were 50 mM, 10 mM and 3 μM, respectively, when the ratio (W/W) of glycerol dehydratase to reactivation factor was 4:1. The O2-inactivated and glycerol-inactivated dehydratase could be reactivated to 97.3% and 98.9% of initial activity in 10 min in above-mentioned conditions, respectively. The reactivation factor together with ATP was considered as the “ON/OFF” reactivating condition.  相似文献   

8.
5S rRNA is an integral component of the large ribosomal subunit in virtually all living organisms. Polyamine binding to 5S rRNA was investigated by cross-linking of N1-azidobenzamidino (ABA)-spermine to naked 5S rRNA or 50S ribosomal subunits and whole ribosomes from Escherichia coli cells. ABA-spermine cross-linking sites were kinetically measured and their positions in 5S rRNA were localized by primer extension analysis. Helices III and V, and loops A, C, D and E in naked 5S rRNA were found to be preferred polyamine binding sites. When 50S ribosomal subunits or poly(U)-programmed 70S ribosomes bearing tRNAPhe at the E-site and AcPhe-tRNA at the P-site were targeted, the susceptibility of 5S rRNA to ABA-spermine was greatly reduced. Regardless of 5S rRNA assembly status, binding of spermine induced significant changes in the 5S rRNA conformation; loop A adopted an apparent ‘loosening’ of its structure, while loops C, D, E and helices III and V achieved a more compact folding. Poly(U)-programmed 70S ribosomes possessing 5S rRNA cross-linked with spermine were more efficient than control ribosomes in tRNA binding, peptidyl transferase activity and translocation. Our results support the notion that 5S rRNA serves as a signal transducer between regions of 23S rRNA responsible for principal ribosomal functions.  相似文献   

9.
Acetohydroxyacid synthase (AHAS) is the key enzyme in branched chain amino acid biosynthesis pathway. The enzyme activity and properties of a highly thermostable AHAS from the hyperthermophilic bacterium Thermotoga maritima is being reported. The catalytic and regulatory subunits of AHAS from T. maritima were over-expressed in Escherichia coli. The recombinant subunits were purified using a simplified procedure including a heat-treatment step followed by chromatography. A discontinuous colorimetric assay method was optimized and used to determine the kinetic parameters. AHAS activity was determined to be present in several Thermotogales including T. maritima. The catalytic subunit of T. maritima AHAS was purified approximately 30-fold, with an AHAS activity of approximately 160±27 U/mg and native molecular mass of 156±6 kDa. The regulatory subunit was purified to homogeneity and showed no catalytic activity as expected. The optimum pH and temperature for AHAS activity were 7.0 and 85 °C, respectively. The apparent Km and Vmax for pyruvate were 16.4±2 mM and 246±7 U/mg, respectively. Reconstitution of the catalytic and regulatory subunits led to increased AHAS activity. This is the first report on characterization of an isoleucine, leucine, and valine operon (ilv operon) enzyme from a hyperthermophilic microorganism and may contribute to our understanding of the physiological pathways in Thermotogales. The enzyme represents the most active and thermostable AHAS reported so far.  相似文献   

10.
A heterodimeric xylosidase (E.C. 3.2.1.37) with robust activity is secreted among the plant cell wall degrading enzymes produced by the saprophytic fungus Humicola insolens. The xylosidase has been purified to homogeneity by gel filtration and cation exchange chromatography, and demonstrated to be composed of two protein subunits of 68 and 17 kDa with a molecular mass in solution of approximately 85 kDa based on a combination of SDS-PAGE, size exclusion chromatography and analytical ultracentrifugation. Peptide sequence identities from the subunits indicate the 68 kDa subunit contains a catalytic protein domain and the 17 kDa subunit a carbohydrate binding module. The xylosidase has wide biotechnological potential with maximum activity exhibited at 70 °C and kinetic constants with p-nitrophenol xylopyranoside substrate that suggest it has the highest catalytic efficiency recorded to date (Vmax 22.17 μmoles/min/mg, Km 1.74 mM and Kcat 6787/s).  相似文献   

11.
The effects of live Saccharomyces cerevisiae (CNCM I-1077) supplementation to 72 Saanen dairy goats in early lactation were investigated. Animals were divided into treatment and control groups balanced for parity and milk production in the first 3 weeks of lactation. Treated animals received 0.2 g/day yeast supplement (Levucell SC20) corresponding to 4 × 109 cfu/day S. cerevisiae, beginning week 3 of lactation, and lasting 15 weeks. Dry matter intake (DMI), body condition score (BCS), milk production and composition were recorded to day126 postpartum. Milk samples were analysed weekly for fat, protein, lactose, urea and SCC. Blood samples were analysed monthly for glucose, BHBA, NEFA, GOT and GGT. Intestinal microflora were assessed monthly in faeces, assaying total bacteria, enterobacteria, E. coli, coliform bacteria, clostridia, lactobacilli, molds and yeast. BCS was not affected by the treatment, but treated animals had greater DMI (2.71 kg/day versus 2.35 kg/day, S.E. ±0.12, P < 0.001) and produced more (P < 0.05) milk (2.38 kg/day versus 2.08 kg/day, S.E. ±0.14) of lower fat content (P < 0.05). Plasma metabolites and liver enzymes (GOT, GGT) were similar between groups, although BHBA levels tended to be lower (P = 0.06) in treated animals. Faecal E. coli were reduced (0.72 log10 cfu/g versus 2.39 log10 cfu/g, S.E. ±0.38, P < 0.05) and lactobacilli increased (P < 0.05) by the treatment. Yeast faecal levels were higher (P < 0.001) in treated animals during supplementation. In this study, live yeast supplementation to early lactating dairy goats significantly increased milk production, in relation to increased DMI, without decreasing BCS. Plasma metabolites (glucose, BHBA and NEFA) were consistent with the BCS findings (no differences between groups) and in addition there was a significant reduction of faecal E. coli content together with increase in lactobacilli in the treated animals, suggesting greater stability of the intestinal ecosystem. Results suggest that the inclusion of Saccharomyces cerevisae in the diet of early lactating dairy goats can be recommended under field conditions.  相似文献   

12.
Slowly cooled cells of Streptomyces aureofaciens contained mainly tight-couple ribosomes. Maximum rate of polyphenylalanine synthesis on ribosomes of S. aureofaciens was observed at 40°C, while cultures grew optimally at 28°C. Ribosomes of S. aureofaciens differed from those of E. coli in the amount of poly(U) required for maximum synthetic activity. The polyphenylalanine-synthesizing activity of E. coli ribosomes was about 3-times higher than that of S. aureofaciens ribosomes. The addition of protein S1 of E. coli or the homologous protein from S. aureofaciens had no stimulatory effect on the translation of poly(U). In order to localize alteration(s) of S. aureofaciens ribosomes in the elongation step of polypeptide synthesis we developed an in vitro system derived from purified elongation factors and ribosomal subunits. The enzymatic binding of Phe-tRNA to ribosomes of S. aureofaciens was significantly lower than the binding to ribosomes of E. coli. This alteration was mainly connected with the function of S. aureofaciens 50 S subunits. These subunits were not deficient in their ability to associate with 30 S subunits or with protein SL5 which is homologous to L7/L12 of E. coli.  相似文献   

13.
An efficient purification system for purifying recombinant Bacillus subtilis 168 catalase (KatA) expressed in Escherichia coli was developed. The basic region containing 252–273 amino acids derived from E. coli ribosomal protein L2 was used as an affinity tag while the small ubiquitin-like modifier (SUMO) was introduced as one specific protease cleavage site between the target protein and the purification tags. L2 (252–273)–SUMO fusion protein purification method can be effectively applied to purify the recombinant catalase using cation exchange resin. This purification procedure was used to purify the KatA and achieved a purification fold of 30.5, a specific activity of 48,227.2 U/mg and an activity recovery of 74.5%. The enzyme showed a Soret peak at 407 nm. The enzyme kept its activity between pH 5 and 10 and between 30 °C and 60 °C, with the highest activity at pH 8.0 and 37 °C. The enzyme displayed an apparent Km of 39.08 mM for hydrogen peroxide. These results agree well with the previous reports about B. subtilis catalase. L2 (252–273)–SUMO fusion protein purification technique provides a novel and effective fusion expression system for the production of recombinant proteins.  相似文献   

14.
In this work, two proteins, Z-domains and bovine casein, were autodisplayed on the outer membrane of the same Escherichia coli cells by co-transformation of two different autodisplay vectors. On the basis of SDS-PAGE densitometry, Z-domains and bovine casein were expressed at 3.12 × 105 and 1.55 × 105 proteins/E. coli cell, respectively. The co-autodisplayed Z-domains had antibody-binding activity and the bovine casein had adhesive properties. E. coli with co-autodisplayed proteins were analyzed by fluorescence assisted cell sorting (FACS). E. coli with co-autodisplayed Z-domains and bovine casein aggregated due to hydrophobic interaction. For application to immunoassays, the Z-domain activity was estimated after (1) immobilizing the E. coli and (2) forming an OM layer. E. coli with co-autodisplayed two proteins that were immobilized on a polystyrene microplate had the same antibody-binding activity as did E. coli with autodisplayed Z-domains only. The OM layer from the co-transformed E. coli had Z-domains and bovine casein expressed at a 1:2 ratio from antibody-binding activity measurements.  相似文献   

15.
《Process Biochemistry》2010,45(6):887-891
For efficient production of (R)-(−)-mandelic acid, a nitrilase gene from Alcaligenes sp. ECU0401 was cloned and overexpressed in Escherichia coli. After simple optimization of the culture conditions, the biocatalyst production was greatly increased from 500 to 7000 U/l. The recombinant E. coli whole cells showed strong tolerance against a high substrate concentration of up to 200 mM, and the concentration of (R)-(−)-mandelic acid after only 4 h of transformation reached 197 mM with an enantiomeric excess (eep) of 99%. In a fed-batch reaction with 600 mM mandelonitrile as the substrate, the cumulative production of (R)-(−)-mandelic acid after 17.5 h of conversion reached 520 mM. The recombinant E. coli cells could also be repeatedly used in the biotransformation, retaining 40% of the initial activity after 10 batches of reaction. The highly substrate/product tolerable and enantioselective nature of this recombinant nitrilase suggests that it is of great potential for the practical production of optically pure (R)-(−)-mandelic acid.  相似文献   

16.
In the present work, NAD+-dependent formate dehydrogenase (FDH), encoded by fdh gene from Candida boidinii was successfully displayed on Escherichia coli cell surface using ice nucleation protein (INP) from Pseudomonas borealis DL7 as an anchoring protein. Localization of matlose binding protein (MBP)-INP-FDH fusion protein on the E. coli cell surface was characterized by SDS-PAGE and enzymatic activity assay. FDH activity was monitored through the oxidation of formate catalyzed by cell-surface-displayed FDH with its cofactor NAD+, and the production of NADH can be detected spectrometrically at 340 nm. After induction for 24 h in Luria-Bertani medium containing isopropyl-β-d-thiogalactopyranoside, over 80% of MBP-INP-FDH fusion protein present on the surface of E. coli cells. The cell-surface-displayed FDH showed optimal temperature of 50 °C and optimal pH of 9.0. Additionally, the cell-surface-displayed FDH retained its original enzymatic activity after incubation at 4 °C for one month with the half-life of 17 days at 40 °C and 38 h at 50 °C. The FDH activity could be inhibited to different extents by some transition metal ions and anions. Moreover, the E. coli cells expressing FDH showed different tolerance to solvents. The recombinant whole cell exhibited high formate specificity. Finally, the E. coli cell expressing FDH was used to assay formate with a wide linear range of 5–700 μM and a low limit of detection of 2 μM. It is anticipated that the genetically engineered cells may have a broad application in biosensors, biofuels and cofactor regeneration system.  相似文献   

17.
Escherichia coli O157:H7, a major foodborne pathogen, has been associated with numerous cases of foodborne illnesses. Rapid methods have been developed for the screening of this pathogen in foods in order to circumvent timely plate culture techniques. Unfortunately, many rapid methods are presumptive and do not claim to confirm the presence of E. coli O157:H7. The previously developed method, enzyme-linked immunomagnetic chemiluminescence (ELIMCL), has been improved upon to allow for fewer incidences of false positives when used to detect E. coli O157:H7 in the presence of mixed cultures. The key feature of this assay is that it combines the highly selective synergism of both anti-O157 and anti-H7 antibodies in the sandwich immunoassay format. This work presents application of a newly semi-automated version of ELIMCL to the detection of E. coli O157:H7 in pristine buffered saline yielding detection limits of approximately 1 × 105 to 1 × 106 of live cells/mL. ELIMCL was further demonstrated to detect E. coli O157:H7 inoculated into artificially contaminated ground beef at ca. 400 CFU/g after a 5 h enrichment and about 1.5 h assay time for a total detection time of about 6.5 h. Finally, ELIMCL was compared with USFDA's Bacteriological Analytical Manual method for E. coli O157:H7 in a double-blind study. Using McNemar's treatment, the two methods were determined to be statistically similar for the detection of E. coli O157:H7 in ground beef inoculated with mixed cultures of select bacteria.  相似文献   

18.
Whole 70S ribosomes and 50S and 30S ribosomal subunits of E. coli and T. maritima were studied by atomic force microscopy. Adsorption of the ribosomal subunits on a substrate revealed considerable heterogeneity of their structures. Analysis of the geometric size of the particles demonstrated essential difference between the heights of E. coli and T. maritima ribosomes 9.4 ± 0.01 nm and 10.35 ± 0.02 nm, respectively. Presumably, the difference in size is determined by the difference in organization of the mobile ribosomal domain, the L7/L12 stalk.  相似文献   

19.
《Process Biochemistry》2007,42(5):856-862
N-Acylamino acid racemase (NAAAR) gene of Deinococcus radiodurans BCRC12827 was cloned into expression vector pQE30 to generate pQE-naaar and expressed in recombinant Escherichia coli JM109. The expressed enzyme purified from the crude cell extract of IPTG-induced E. coli JM109 (pQE-naaar) exhibited high racemization activity to N-carbamoyl-l-homophenylalanine (NCa-l-HPA) and N-carbamoyl-d-homophenylalanine (NCa-d-HPA) with specific activities of 1.91 U/mg protein and 1.31 U/mg protein, respectively. To develop a recombinant E. coli whole cell system for the conversion of racemic NCa-HPA to l-homophenylalanine (l-HPA), naaar gene from D. radiodurans and l-N-carbamoylase (LNCA) gene from Bacillus kaustophilus BCRC11223 were cloned and coexpressed in E. coli cells. Recombinant cells treated with 0.5% toluene at 30 °C for 30 min exhibited enhanced NAAAR and LNCA activities, which are about 20- and 60-fold, respectively, higher than those of untreated cells. Using toluene-permeabilized recombinant E. coli cells, a maximal productivity of 7.5 mmol l-HPA/l h with more than 99% yield could be obtained from 150 mmol racemic NCa-HPA. Permeabilized cells also showed considerable stability in the bioconversion process using 10 mmol racemic NCa-HPA as substrate, no significantly decrease in conversion yield for l-HPA was found in the eight cycles.  相似文献   

20.
《Process Biochemistry》2010,45(2):147-152
A highly active recombinant whole-cell biocatalyst, Escherichia coli pETAB2/pG-KJE1, was developed for the efficient production of (S)-styrene oxide from styrene. The recombinant E. coli overexpressed styAB the genes of styrene monooxygenase of Pseudomonas putida SN1 and coexpressed the genes encoding chaperones (i.e., GroEL–GroES and DnaK–DnaJ–GrpE). The styrene monooxygenases were produced to ca. 40% of the total soluble proteins, enabling the whole-cell activity of the recombinant of 180 U/g CDW. The high StyAB activity in turn appeared to direct cofactors and molecular oxygen to styrene epoxidation. The product yield on energy source (i.e., glucose) reached ca. 40%. In addition, biotransformation in an organic/aqueous two-liquid phase system allowed the product to accumulate to 400 mM in the organic phase within 6 h, resulting in an average specific and volumetric productivity of 6.4 mmol/g dry cells/h (106 U/g dry cells) and 67 mM/h (1110 U/Laq), respectively, under mild reaction conditions. These results indicated that the high productivity and the high product yield on energy source were driven by the high enzyme activity. Therefore, it was concluded that oxygenase activity of whole-cell biocatalysts is one of the critical factors to determine their catalytic performance.  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