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 0 毫秒
1.
An oleaginous hydrocarbon-degrading Rhodococcus opacus strain (PD630) was isolated from a soil sample. The cells were able to grow on a variety of substrates and to produce large amounts of three different types of intracellular inclusions during growth on alkanes, phenylalkanes, or non-hydrocarbon substrates. Electron microscopy revealed large numbers of electron-transparent inclusions with a sphere-like structure. In addition, electron-dense inclusions representing polyphosphate and electron-transparent inclusions with an elongated disc-shaped morphology occurred in small amounts. The electron-transparent inclusions of alkane- or gluconate-grown cells were composed of neutral lipids (98%, w/w), phospholipids (1.2%, w/w), and protein (0.8%, w/w). The major component of the cellular inclusions was triacylglycerols; minor amounts of diacylglycerols and probably also some free fatty acids were also present. Free fatty acids and/or fatty acids in acylglycerols in cells of R. opacus amounted up to 76 or 87% of the cellular dry weight in gluconate- or olive-oil-grown cells, respectively. The fatty acid composition of the inclusions depended on the substrate used for cultivation. In cells cultivated on n-alkanes, the composition of the fatty acids was related to the substrate, and intermediates of the β-oxidation pathway, such as hexadecanoic or pentadecanoic acid, were among the acylglycerols. Hexadecanoic acid was also the major fatty acid (up 36% of total fatty acids) occurring in the lipid inclusions of gluconate-grown cells. This indicated that strain PD630 utilized β-oxidation and de novo fatty acid biosynthesis for the synthesis of storage lipids. Inclusions isolated from phenyldecane-grown cells contained mainly the non-modified substrate and phenylalkanoic acids derived from the hydrocarbon oxidation, such as phenyldecanoic acid, phenyloctanoic acid, and phenylhexanoic acid, and approximately 5% (w/w) of diacylglycerols. The lipid inclusions seemed to have definite structures, probably with membranes at their surfaces, which allow them to maintain their shape, and with some associated proteins, probably involved in the inclusion formation. Received: 22 December 1995 / Accepted: 12 March 1996  相似文献   

2.
The time course of the accumulation of triacylglycerols (TAGs) in Rhodococcus opacus PD630 or of TAGs plus polyhydroxyalkanoates (PHA) in Rhodococcus ruber NCIMB 40126 with gluconate or glucose as carbon source, respectively, was studied. In addition, we examined the mobilization of these storage compounds in the absence of a carbon source. R. opacus accumulated TAGs only after the exhaustion of ammonium in the medium, and, with a fixed concentration of the carbon source, the amounts of TAGs in the cells increased with decreasing concentrations of ammonium in the medium. When these cells were incubated in the absence of an additional carbon source, about 90% of these TAGs were mobilized and used as endogenous carbon source, particularly if ammonium was available. R. ruber accumulated a copolyester consisting of 3-hydroxybutyrate and 3-hydroxyvalerate already during the early exponential growth phase, whereas TAGs were synthesized and accumulated mainly during the late exponential and stationary growth phases. In the stationary growth phase, synthesis of TAGs continued, whereas PHA was partially mobilized. In the absence of an additional carbon source but in the presence of ammonium, mobilization of TAGs started first and was then paralleled by the mobilization of PHA, resulting in an approximately 90% and 80% decrease of these storage compounds, respectively. During the accumulation phase, interesting shifts in the composition of the two storage compounds occurred, indicating that the substrates of the PHA synthase and the TAG synthesizing enzymes were provided to varying extents, depending on whether the cells were in the early or late exponential or in the stationary growth phase. Received: 12 January 2000 / Received revision: 22 February 2000 / Accepted: 25 February 2000  相似文献   

3.
Members of the genus Rhodococcus were investigated for their ability to produce glycogen during cultivation on gluconate or glucose. Strains belonging to Rhodococcus ruber, Rhodococcus opacus, Rhodococcus fascians, Rhodococcus erythropolis and Rhodococcus equi were able to produce glycogen up to 0.2–5.6% of cellular dry weight (CDW). The glycogen content varied from 0.8% to 3.2% of CDW in cells of R. opacus PD630, which is a well-known oleaginous bacterium, during the exponential growth phase, when cultivated on diverse carbon sources. Maltose and pyruvate promoted glycogen accumulation by cells of strain PD630 to a greater extent than glucose, gluconate, lactose, sucrose or acetate. This strain was able to produce triacylglycerols, polyhydroxyalkanoates and glycogen as storage compounds during growth on gluconate, although triacylglycerols were always the main product under the conditions of this study. Cerulenin, an inhibitor of de novo fatty acid synthesis, inhibited the accumulation of triacylglycerols from gluconate and increased the content of polyhydroxyalkanoates (from 2.0% to 4.2%, CDW) and glycogen (from 0.1% to 3.0%, CDW). An increase of the polyhydroxyalkanoates and glycogen content was also observed in two mutants of R. opacus PD630, which produced reduced amounts of triacylglycerols during cultivation of cells on gluconate.  相似文献   

4.
The triacylglycerol (TAG)-accumulating bacterium Rhodococcus opacus strain PD630 was investigated with respect to the fermentative production of TAGs consisting of an unusually high fraction of fatty acids with an odd-number of carbon atoms and unsaturated monoenic fatty acids from sugar beet molasses and sucrose. Fed-batch fermentations were optimized at the 30-1 scale in a stirred tank bioreactor at 30 degrees C using a mineral salts medium, which contained sugar beet molasses and sucrose as sole carbon sources. Approximately 37.5 g cell dry matter (CDM) per liter was the highest cell density that was obtained at that scale with a TAG content in the cells of 52%. This fermentative process was also applied to a 500-1 pilot-plant scale. Cell densities as high as 18.4 g CDM per liter were obtained, and 42% of the sucrose present in the medium was converted into cell mass which consisted of 38.4% TAGs.  相似文献   

5.
6.
  1. Download : Download high-res image (117KB)
  2. Download : Download full-size image
  相似文献   

7.
8.
9.
The in vivo effects of sterculic acid methyl ester on triacylglycerol fatty acid composition in the oleaginous, hydrocarbon-degrading bacterium R. opacus strain PD630 was investigated. Sterculic acid, a cyclopropene fatty acid and an inhibitor of the stearoyl-CoA desaturase system, strongly inhibited the synthesis of monoenic fatty acids, of saturated fatty acids with more than 16 carbon atoms and of odd-numbered fatty acids when added to the culture medium. In addition, chemical mutagenesis and the application of the penicillin enrichment technique provided mutants, which were more or less completely impaired in the desaturation of long-chain fatty acids and exhibited in some cases a similar fatty acid composition like the wild-type in the presence of sterculic acid methyl ester. The implications of these findings for fatty acid metabolism in R. opacus strain PD630 are discussed.  相似文献   

10.
Sequence analysis of a 9-kb genomic fragment of the actinobacterium Rhodococcus opacus 1CP led to identification of an open reading frame encoding a novel fusion protein, StyA2B, with a putative function in styrene metabolism via styrene oxide and phenylacetic acid. Gene cluster analysis indicated that the highly related fusion proteins of Nocardia farcinica IFM10152 and Arthrobacter aurescens TC1 are involved in a similar physiological process. Whereas 413 amino acids of the N terminus of StyA2B are highly similar to those of the oxygenases of two-component styrene monooxygenases (SMOs) from pseudomonads, the residual 160 amino acids of the C terminus show significant homology to the flavin reductases of these systems. Cloning and functional expression of His10-StyA2B revealed for the first time that the fusion protein does in fact catalyze two separate reactions. Strictly NADH-dependent reduction of flavins and highly enantioselective oxygenation of styrene to (S)-styrene oxide were shown. Inhibition studies and photometric analysis of recombinant StyA2B indicated the absence of tightly bound heme and flavin cofactors in this self-sufficient monooxygenase. StyA2B oxygenates a spectrum of aromatic compounds similar to those of two-component SMOs. However, the specific activities of the flavin-reducing and styrene-oxidizing functions of StyA2B are one to two orders of magnitude lower than those of StyA/StyB from Pseudomonas sp. strain VLB120.The incorporation of one atom of oxygen during hydroxylation, epoxidation, sulfoxidation, or Baeyer-Villiger oxidation is a common initial step of the aerobic degradation of aromatic compounds by microorganisms. In bacteria, these reactions are most frequently catalyzed by inducible flavoprotein monooxygenases (EC 1.14.13 [57]). The majority of these enzymes (so-called single-component flavoprotein monooxygenases) utilize electrons from NAD(P)H, which are transferred to a non-covalently bound flavin adenine dinucleotide (FAD) in order to activate molecular oxygen as a flavin (hydro)peroxide. Depending on the protonation of this intermediate and the type of substrate, an oxygen atom is then incorporated by nucleophilic or electrophilic attack. More recently, different two-component flavoprotein monooxygenases have been characterized (57). These systems cover an NAD(P)H-dependent flavin reductase in order to generate reduced flavin and an oxygenase that utilizes this cofactor for the activation of oxygen.The exquisite regio- and stereoselectivities of oxygen insertion by flavoprotein monooxygenases favor these enzymes for biocatalytic applications (23, 24, 33). This is especially true because chemical synthesis approaches by hetero- or homogenic catalysis often do not yield a sufficiently high enantiomeric excess for the production of pharmaceuticals and their chiral building blocks. The use of oxygen as an inexpensive nontoxic oxidant and mild reaction conditions are additional advantages with the potential for increasing the environmental sustainability of oxygenase-catalyzed biotransformations.The necessity for expensive cofactors is perhaps the most striking drawback limiting the industrial application of flavoprotein monooxygenases. Different electrochemical and enzymatic procedures for in vitro cofactor regeneration are available (20, 21, 32, 52, 56), but these systems are currently lacking long-term stability. As a consequence, the practical application of flavoprotein monooxygenases is virtually restricted to in vivo systems in which cofactor regeneration is mediated by the metabolism of the expression host (45, 49). The limitations of whole-cell-mediated biotransformations by substrate and/or product toxicity can be overcome by means of two-phase systems, as was recently shown for the two-component styrene monooxygenase (SMO) from Pseudomonas sp. strain VLB120 (45).Two-component flavoprotein monooxygenases present additional challenges for biocatalytic applications. The need for two separate protein components may hamper attempts at recombinant enzyme expression, the application of immobilized enzymes in cell-free systems, and the detection of novel oxygenases during activity-based metagenome-screening approaches. Moreover, and as was already shown for the two-component SMO from Pseudomonas sp. strain VLB120 (44), the interprotein transfer of reduced FAD is accompanied to a certain extent by the auto-oxidative formation of reactive oxygen species such as hydrogen peroxide (Fig. (Fig.1).1). Auto-oxidation of reduced FAD not only decreases the efficiency of the oxygenation process but also negatively interferes with the physiological conditions of the cell. The extent of oxidative stress is considerably increased when FAD oxidoreductase activity exceeds oxygenase activity and uncoupling becomes dominant.Open in a separate windowFIG. 1.Catalytic mechanism of two-component SMOs and the formation of oxidative stress by uncoupling between FAD oxidoreductase (StyB) and oxygenase (StyA) (adapted from reference 36). FADOX and FADRED, oxidized and reduced forms of FAD, respectively.Presently the details of reduced-FAD transfer between the oxygenase and FAD oxidoreductase components of SMOs are not known. Recent kinetic studies have indicated that reduced FAD is transferred by a mixed mechanism in which direct contact of both proteins and free diffusion of the reduced cofactor play a role (25). This hypothesis is supported by the work of Otto and coworkers in which the formation of hydrogen peroxide was shown to be reciprocally proportional to the concentration of active oxygenase StyA (44). The high level of efficiency of self-sufficient cytochrome P450 enzymes compared to that of multicomponent types is attributed to the closer location between the heme-containing P450 domain (oxygenase) and a reductase domain (FAD/flavin mononucleotide [FMN] and NADH binding site), which should also promote the efficiency of diffusive transfer (38). These self-sufficient P450 systems are of high biocatalytic interest (8, 34, 39), and it is likely that other types of self-sufficient monooxygenases (e.g., flavoenzymes) behave in a similar way.Members of the gram-positive genus Rhodococcus are characterized by their exceptionally high level of metabolic versatility toward a broad range of organic substrates (31). Large genome sizes, the presence of megaplasmids, and a distinct gene redundancy (multiple enzyme homologs) favor these organisms as a promising source for novel enzymes (59). Moreover, several studies have provided evidence of functionally convergent evolution of the catabolic activities of rhodococci and proteobacteria (12, 37). Since most research on bacterial catabolic activities has so far been performed on the latter group, novel enzymes and mechanisms are likely to be identified in gram-positive bacteria.The nocardioform actinobacterium Rhodococcus opacus 1CP was originally isolated from contaminated soil by enrichment with 4-chloro- and 2,4-dichlorophenol as the sole carbon sources (18). To solve questions related to the enzymes involved in chlorophenol mineralization, an R. opacus 1CP clone library was generated, leading to the identification of a 9-kb genomic fragment harboring genes with a presumed function in styrene metabolism. Sequence analysis indicated the presence of an open reading frame (ORF) encoding a fusion protein composed of an oxygenase and a reductase subunit with a high level of similarity to the corresponding subunits of two-component SMOs from pseudomonads. Recombinant expression and biochemical characterization confirmed that it has enantioselective styrene epoxidation ability and showed that StyA2B is the first representative of a new class of NADH- and flavin-dependent single-component flavoprotein monooxygenases.  相似文献   

11.
Rhodococcus opacus 1CP, a potent degrader of (chloro-) aromatic compounds was found to utilise C10–C16 n-alkanes as sole carbon sources. Highest conversion rates were observed with n-tetradecane and n-hexadecane, whereas the utilisation of n-dodecane and n-decane was considerably slower. Thin-layer chromatography of organic extracts of n-alkane-grown 1CP cultures indicated the growth-associated formation of a glycolipid which was characterised as a trehalose dimycolate by 1H-NMR spectroscopy and mass spectrometry. Total chain lengths between 48 and 54 carbons classify the fatty acid residues as nocardiomycolic acids. The presence of two double bonds in each mycolic acid is another feature that distinguishes the corresponding trehalose dinocardiomycolates from trehalose dicorynomycolates reported for Rhodococcus erythropolis DSM43215 and Rhodococcus ruber IEGM231. R. opacus 1CP was not found, even under nitrogen limitation, to produce anionic trehalose tetraesters which have previously been reported for R. erythropolis DSM43215.  相似文献   

12.
Preparative isolation of phosphatidyl serine from brain   总被引:3,自引:0,他引:3  
  相似文献   

13.
Immobilization of microorganisms on/in insoluble carriers is widely used to stabilize functional activity of microbial cells in industrial biotechnology. We immobilized Rhodococcus ruber, an important hydrocarbon degrader, on biosurfactant-coated sawdust. A biosurfactant produced by R. ruber in the presence of liquid hydrocarbons was found to enhance rhodococcal adhesion to solid surfaces, and thus, it was used as a hydrophobizing agent to improve bacterial attachment to a sawdust carrier. Compared to previously used hydrophobizers (drying oil and n-hexadecane) and emulsifiers (methyl- and carboxymethyl cellulose, poly(vinyl alcohol), and Tween 80), Rhodococcus biosurfactant produced more stable and homogenous coatings on wood surfaces, thus resulting in higher sawdust affinity to hydrocarbons, uniform monolayer distribution of immobilized R. ruber cells (immobilization yield 29–30 mg dry cells/g), and twofold increase in hydrocarbon biooxidation rates compared to free rhodococcal cells. Two physical methods, i.e., high-resolution profilometry and infrared thermography, were applied to examine wood surface characteristics and distribution of immobilized R. ruber cells. Sawdust-immobilized R. ruber can be used as an efficient biocatalyst for hydrocarbon transformation and degradation.  相似文献   

14.
Chloromuconolactone dehalogenase ClcF plays a unique role in 3-chlorocatechol degradation by Rhodococcus opacus 1CP by compensating the inability of its chloromuconate cycloisomerase ClcB2 to dechlorinate the chemically stable cycloisomerization product (4R,5S)-5-chloromuconolactone (5CML). High sequence similarities showed relatedness of ClcF to muconolactone isomerases (MLIs, EC 5.3.3.4) of the 3-oxoadipate pathway. Although both enzyme types share the ability to dechlorinate 5CML, comparison of kcat/Km indicated a significant extent of specialization of ClcF for dechlorination. This assumption was substantiated by an almost complete inability of ClcF to convert (4S)-muconolactone and the exclusive formation of cis-dienelactone from 5CML. Mutational analysis of ClcF by means of variants E27D, E27Q, Y50A, N52A, and A89S indicated relevance of some highly conserved residues for substrate binding and catalysis. Based on the putative isomerization mechanism of MLI, evidence was provided for a role of E27 in initial proton abstraction as well as of Y50 and N52 in substrate binding. In case of N52 substrate binding is likely to occur to the carboxylic group of 5CML as indicated by a significant change of product specificity. Expression in Escherichia coli BL21-CP(DE)-RIL followed by a three-step purification procedure with heat treatment is a convenient strategy to obtain recombinant ClcF and variants thereof.  相似文献   

15.
The growth of Rhodococcus opacus GM-14 on mixtures of 2-chloro- and 2-bromophenol, of 4-chloro, 4-bromo-, and 4-iodophenol, and of chloro-, bromo-, and iodobenzenes was accompanied by consumption of the substrates and excretion of halogen ions to the medium. During the growth on monochlorophenols, the substrates were consumed sequentially in the following order: 4-chloro-, 3-chloro-, and then 2-chlorophenol. Chlorine ions were excreted in a two-phase manner in amounts comprising 79% of the theoretical yield. The diauxic growth of R. opacus GM-14 can be explained by the existence in this bacterium of two modified metabolic pathways for the ortho-cleavage of halogenated pyrocatechols. The first pathway included 4-halogeno- or dihalogenopyrocatechols as intermediates, whereas the second pathway included 3-halogenopyrocatechols.  相似文献   

16.
The noncellulolytic actinomycete Rhodococcus opacus strain PD630 is the model oleaginous prokaryote with regard to the accumulation and biosynthesis of lipids, which serve as carbon and energy storage compounds and can account for as much as 87% of the dry mass of the cell in this strain. In order to establish cellulose degradation in R. opacus PD630, we engineered strains that episomally expressed six different cellulase genes from Cellulomonas fimi ATCC 484 (cenABC, cex, cbhA) and Thermobifida fusca DSM43792 (cel6A), thereby enabling R. opacus PD630 to degrade cellulosic substrates to cellobiose. Of all the enzymes tested, five exhibited a cellulase activity toward carboxymethyl cellulose (CMC) and/or microcrystalline cellulose (MCC) as high as 0.313 ± 0.01 U · ml−1, but recombinant strains also hydrolyzed cotton, birch cellulose, copy paper, and wheat straw. Cocultivations of recombinant strains expressing different cellulase genes with MCC as the substrate were carried out to identify an appropriate set of cellulases for efficient hydrolysis of cellulose by R. opacus. Based on these experiments, the multicellulase gene expression plasmid pCellulose was constructed, which enabled R. opacus PD630 to hydrolyze as much as 9.3% ± 0.6% (wt/vol) of the cellulose provided. For the direct production of lipids from birch cellulose, a two-step cocultivation experiment was carried out. In the first step, 20% (wt/vol) of the substrate was hydrolyzed by recombinant strains expressing the whole set of cellulase genes. The second step was performed by a recombinant cellobiose-utilizing strain of R. opacus PD630, which accumulated 15.1% (wt/wt) fatty acids from the cellobiose formed in the first step.  相似文献   

17.
The growth of Rhodococcus opacus 1CP in medium with different concentrations of p-cresol as the sole source of carbon and energy was studied. It was shown that the optimal concentration of p-cresol was 600 mg/L. The ability of this strain to transform practically all amounts of p-cresol to 4-methylcatechol followed by its utilization through ortho-pathway was shown. New enzymes (4-methylcatechol 1,2-dioxygenase, catechol 1,2-dioxygenase, and methylmuconate cycloisomerase) were purified to homogeneity and characterized. Based on the data obtained on p-cresol degradation, formation of intermediates, and the enzymes participating in this pathway, we suggest an ortho-pathway of p-cresol degradation by R. opacus 1CP through 4-methylcatechol and 3-methyl-cis, cis-muconate.  相似文献   

18.
Bacterial strain M213 was isolated from a fuel oil-contaminated soil in Idaho, USA, by growth on naphthalene as a sole source of carbon, and was identified as Rhodococcus opacus M213 by 16S rDNA sequence analysis and growth on substrates characteristic of this species. M213 was screened for growth on a variety of aromatic hydrocarbons, and growth was observed only on simple 1 and 2 ring compounds. No growth or poor growth was observed with chlorinated aromatic compounds such as 2,4-dichlorophenol and chlorobenzoates. No growth was observed by M213 on salicylate, and M213 resting cells grown on naphthalene did not attack salicylate. In addition, no salicylate hydroxylase activity was detected in cell free lysates, suggesting a pathway for naphthalene catabolism that does not pass through salicylate. Enzyme assays indicated induction of catechol 1,2-dioxygenase and catechol 2,3-dioxygenase on different substrates. Total DNA from M213 was screened for hybridization with a variety of genes encoding catechol dioxygenases, but hybridization was observed only with catA (encoding catechol 1,2-dioxygenase) from R. opacus 1CP and edoD (encoding catechol 2,3-dioxygenase) from Rhodococcus sp. I1. Plasmid analysis indicated the presence of two plasmids (pNUO1 and pNUO2). edoD hybridized to pNUO1, a very large (approximately 750 kb) linear plasmid.  相似文献   

19.
The schemes for preparative electrophoretic isolation and purification of major proteins from influenza virus are described. The viral envelope protein, hemagglutinin, two of its subunits, internal M and NP proteins of influenza viruses A/FPV/Rostock (H7N1), A/PR/8/34 (H1N1) and X-31 (H3N2) were obtained in preparative amounts and characterized by amino acid and N-terminus analyses.  相似文献   

20.
We have recently isolated a biofilm-producing strain (C208) of Rhodococcus ruber that degraded polyethylene at a rate of 0.86% per week (r 2=0.98). Strain C208 adheres to polyethylene immediately upon exposure to the polyolefin. This initial biofilm differentiates (in a stepwise process that lasts about 20 h) into cell-aggregation-forming microcolonies. Further organization yields “mushroom-like” three-dimensional structures on the mature biofilm. The ratio between the population densities of the biofilm and the planktonic C208 cells after 10 days of incubation was about 60:1, indicating a high preference for the biofilm mode of growth. Analysis of extracellular polymeric substances (EPS) in the biofilm of C208 revealed that the polysaccharides level was up to 2.5 folds higher than that of the protein. The biofilm showed a high viability even after 60 days of incubation, apparently due to polyethylene biodegradation.  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