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 343 毫秒
1.
Genomic and proteomic approaches were used to investigate phthalate and benzoate catabolism in Rhodococcus sp. strain RHA1, a polychlorinated biphenyl-degrading actinomycete. Sequence analyses identified genes involved in the catabolism of benzoate (ben) and phthalate (pad), the uptake of phthalate (pat), and two branches of the beta-ketoadipate pathway (catRABC and pcaJIHGBLFR). The regulatory and structural ben genes are separated by genes encoding a cytochrome P450. The pad and pat genes are contained on a catabolic island that is duplicated on plasmids pRHL1 and pRHL2 and includes predicted terephthalate catabolic genes (tpa). Proteomic analyses demonstrated that the beta-ketoadipate pathway is functionally convergent. Specifically, the pad and pat gene products were only detected in phthalate-grown cells. Similarly, the ben and cat gene products were only detected in benzoate-grown cells. However, pca-encoded enzymes were present under both growth conditions. Activity assays for key enzymes confirmed these results. Disruption of pcaL, which encodes a fusion enzyme, abolished growth on phthalate. In contrast, after a lag phase, growth of the mutant on benzoate was similar to that of the wild type. Proteomic analyses revealed 20 proteins in the mutant that were not detected in wild-type cells during growth on benzoate, including a CatD homolog that apparently compensated for loss of PcaL. Analysis of completed bacterial genomes indicates that the convergent beta-ketoadipate pathway and some aspects of its genetic organization are characteristic of rhodococci and related actinomycetes. In contrast, the high redundancy of catabolic pathways and enzymes appears to be unique to RHA1 and may increase its potential to adapt to new carbon sources.  相似文献   

2.
Benzoate catabolism is thought to play a key role in aerobic bacterial degradation of biphenyl and polychlorinated biphenyls (PCBs). Benzoate catabolic genes were cloned from a PCB degrader, Rhodococcus sp. strain RHA1, by using PCR amplification and temporal temperature gradient electrophoresis separation. A nucleotide sequence determination revealed that the deduced amino acid sequences encoded by the RHA1 benzoate catabolic genes, benABCDK, exhibit 33 to 65% identity with those of Acinetobacter sp. strain ADP1. The gene organization of the RHA1 benABCDK genes differs from that of ADP1. The RHA1 benABCDK region was localized on the chromosome, in contrast to the biphenyl catabolic genes, which are located on linear plasmids. Escherichia coli cells containing RHA1 benABCD transformed benzoate to catechol via 2-hydro-1,2-dihydroxybenzoate. They transformed neither 2- nor 4-chlorobenzoates but did transform 3-chlorobenzoate. The RHA1 benA gene was inactivated by insertion of a thiostrepton resistance gene. The resultant mutant strain, RBD169, neither grew on benzoate nor transformed benzoate, and it did not transform 3-chlorobenzoate. It did, however, exhibit diminished growth on biphenyl and growth repression in the presence of a high concentration of biphenyl (13 mM). These results indicate that the cloned benABCD genes could play an essential role not only in benzoate catabolism but also in biphenyl catabolism in RHA1. Six rhodococcal benzoate degraders were found to have homologs of RHA1 benABC. In contrast, two rhodococcal strains that cannot transform benzoate were found not to have RHA1 benABC homologs, suggesting that many Rhodococcus strains contain benzoate catabolic genes similar to RHA1 benABC.  相似文献   

3.
Recent microarray experiments suggested that Burkholderia xenovorans LB400, a potent polychlorinated biphenyl (PCB)-degrading bacterium, utilizes up to three apparently redundant benzoate pathways and a C(1) metabolic pathway during biphenyl and benzoate metabolism. To better characterize the roles of these pathways, we performed quantitative proteome profiling of cells grown on succinate, benzoate, or biphenyl and harvested during either mid-logarithmic growth or the transition between the logarithmic and stationary growth phases. The Bph enzymes, catabolizing biphenyl, were approximately 16-fold more abundant in biphenyl- versus succinate-grown cells. Moreover, the upper and lower bph pathways were independently regulated. Expression of each benzoate pathway depended on growth substrate and phase. Proteins specifying catabolism via benzoate dihydroxylation and catechol ortho-cleavage (ben-cat pathway) were approximately an order of magnitude more abundant in benzoate- versus biphenyl-grown cells at the same growth phase. The chromosomal copy of the benzoyl-coenzyme A (CoA) (box(C)) pathway was also expressed during growth on biphenyl: Box(C) proteins were approximately twice as abundant as Ben and Cat proteins under these conditions. By contrast, proteins of the megaplasmid copy of the benzoyl-CoA (box(M)) pathway were only detected in transition-phase benzoate-grown cells. Other proteins detected at increased levels in benzoate- and biphenyl-grown cells included general stress response proteins potentially induced by reactive oxygen species formed during aerobic aromatic catabolism. Finally, C(1) metabolic enzymes were present in biphenyl-grown cells during transition phase. This study provides insights into the physiological roles and integration of apparently redundant catabolic pathways in large-genome bacteria and establishes a basis for investigating the PCB-degrading abilities of this strain.  相似文献   

4.
5.
Since starvation for carbon sources is a common condition for bacteria in nature and it can also occur in industrial fermentation processes due to mixing zones, knowledge about the response of cells to carbon starvation is beneficial. The preferred carbon source for bacilli is glucose. The response of Bacillus pumilus cells to glucose starvation using metabolic labeling and quantitative proteomics was analyzed. Glucose starvation led to an extensive reprogramming of the protein expression pattern in B. pumilus. The amounts of proteins of the central carbon metabolic pathways (glycolysis and TCC) remained stable in starving cells. Proteins for gluconeogenesis were found in higher amounts during starvation. Furthermore, many proteins involved in acquisition and usage of alternative carbon sources were present in elevated amounts in starving cells. Enzymes for fatty acid degradation and proteases and peptidases were also found in higher abundance when cells entered stationary phase. Among the proteins found in lower amounts were many enzymes involved in amino acid and nucleotide synthesis and several NRPS and PKS proteins.  相似文献   

6.
7.
The ability of microorganisms to assimilate aromatic substances as alternative carbon sources is the basis of biodegradation of natural as well as industrial aromatic compounds. In this study, Corynebacterium glutamicum was grown on benzoate as sole carbon and energy source. To extend the scarce knowledge about physiological adaptation processes occurring in this cell compartment, the membrane proteome was investigated under quantitative and qualitative aspects by applying shotgun proteomics to reach a comprehensive survey. Membrane proteins were relatively quantified using an internal standard metabolically labeled with 15N. Altogether, 40 proteins were found to change their abundance during growth on benzoate in comparison to glucose. A global adaptation was observed in the membrane of benzoate‐grown cells, characterized by increased abundance of proteins of the respiratory chain, by a starvation response, and by changes in sulfur metabolism involving the regulator McbR. Additional to the relative quantification, stable isotope‐labeled synthetic peptides were used for the absolute quantification of the two benzoate transporters of C. glutamicum, BenK and BenE. It was found that both transporters were expressed during growth on benzoate, suggesting that both contribute substantially to benzoate uptake.  相似文献   

8.
The degree and temporal context of variations in ribosome content during nutrient starvation of two copiotrophic marine bacteria, Vibrio alginolyticus and Vibrio furnissii, have been examined. The organisms were starved either by nutritional shift-down or by consumption of limiting nutrients resulting from growth into stationary phase. Measurements of the amount of hybridization to 16S rRNA-specific probes revealed that the cells retained between 10 and 26% of their original rRNA content after 15 days of starvation. In V. alginolyticus, losses in stationary-phase cells occurred rapidly (1 to 2 days), whereas cells shifted into starvation remained larger and retained considerably more rRNA. The ability of V. alginolyticus to recover from starvation was assessed after cells were maintained for 2, 8, and 15 days in nutrient-depleted medium. The pattern of recovery at the level of rRNA accumulation depended upon the duration of nutrient deprivation and the manner in which it was imposed. Stationary-phase cells starved for 2 days had only slight relative increases in rRNA levels after excess nutrients were added. As the duration of starvation lengthened to 8 and 15 days, increasingly greater amounts of rRNA (30 and 70 times preenrichment values, respectively) were transcribed after nutrient enrichment. Shift-down cells recovered from 2 and 8 days of starvation without extensive rRNA production. After 15 days, nutrient enrichment caused 16S rRNA levels to increase 30-fold. The results indicate that the mechanisms controlling starvation-survival in these marine bacterial species are linked to the physiological state at the onset of starvation and that the subsequent pattern of recovery will depend upon how starvation was initiated.  相似文献   

9.
The degree and temporal context of variations in ribosome content during nutrient starvation of two copiotrophic marine bacteria, Vibrio alginolyticus and Vibrio furnissii, have been examined. The organisms were starved either by nutritional shift-down or by consumption of limiting nutrients resulting from growth into stationary phase. Measurements of the amount of hybridization to 16S rRNA-specific probes revealed that the cells retained between 10 and 26% of their original rRNA content after 15 days of starvation. In V. alginolyticus, losses in stationary-phase cells occurred rapidly (1 to 2 days), whereas cells shifted into starvation remained larger and retained considerably more rRNA. The ability of V. alginolyticus to recover from starvation was assessed after cells were maintained for 2, 8, and 15 days in nutrient-depleted medium. The pattern of recovery at the level of rRNA accumulation depended upon the duration of nutrient deprivation and the manner in which it was imposed. Stationary-phase cells starved for 2 days had only slight relative increases in rRNA levels after excess nutrients were added. As the duration of starvation lengthened to 8 and 15 days, increasingly greater amounts of rRNA (30 and 70 times preenrichment values, respectively) were transcribed after nutrient enrichment. Shift-down cells recovered from 2 and 8 days of starvation without extensive rRNA production. After 15 days, nutrient enrichment caused 16S rRNA levels to increase 30-fold. The results indicate that the mechanisms controlling starvation-survival in these marine bacterial species are linked to the physiological state at the onset of starvation and that the subsequent pattern of recovery will depend upon how starvation was initiated.  相似文献   

10.
Heo J  Staples CR  Halbleib CM  Ludden PW 《Biochemistry》2000,39(27):7956-7963
Radiolabeling studies support the existence of a nonsubstrate CO ligand (CO(L)) to the Fe atom of the proposed [FeNi] cluster of carbon monoxide dehydrogenase (CODH) from Rhodospirillum rubrum. Purified CODH has variable amounts of CO(L) dissociated depending on the extent of handling of the proteins. This dissociated CO(L) can be restored by incubation of CODH with CO, resulting in a 30-40% increase in initial activity relative to as-isolated purified CODH. A similar amount of CO(L) binding is observed when as-isolated purified CODH is incubated with (14)CO: approximately 0.33 mol of CO binds per 1 mol of CODH. Approximately 1 mol of CO was released from CO-preincubated CODH upon denaturation of the protein. No CO could be detected upon denaturation of CODH that had been incubated with cyanide. CO(L) binds to both Ni-containing and Ni-deficient CODH, indicating that CO(L) is liganded to the Fe atom of the proposed [FeNi] center. Furthermore, the Ni in the CO(L)-deficient CODH can be removed by treatment with a Ni-specific chelator, dimethylglyoxime. CO preincubation protects the dimethylglyoxime-labile Ni, indicating that CO(L) is also involved in the stability of Ni in the proposed [FeNi] center.  相似文献   

11.
Methanosarcina acetivorans, a member of the methanogenic archaea, can grow with carbon monoxide (CO) as the sole energy source and generates, unlike other methanogens, substantial amounts of acetate and formate in addition to methane. Phenotypic analyses of mutant strains lacking the cooS1F operon and the cooS2 gene suggest that the monofunctional carbon monoxide dehydrogenase (CODH) system contributes to, but is not required for, carboxidotrophic growth of M. acetivorans. Further, qualitative proteomic analyses confirm a recent report (Lessner et al., Proc Natl Acad Sci USA, 103:17921–17926, 2006) in showing that the bifunctional CODH/acetyl-CoA synthase (ACS) system, two enzymes involved in CO2-reduction, and a peculiar protein homologous to both corrinoid proteins and methyltransferases are synthesized at elevated levels in response to CO; however, the finding that the latter protein is also abundant when trimethylamine serves as growth substrate questions its proposed involvement in the reduction of methyl-groups to methane. Potential catabolic mechanisms and metabolic adaptations employed by M. acetivorans to effectively utilize CO are discussed.  相似文献   

12.
13.
Growing cultures of Fibrobacter succinogenes assimilated more ammonia than could be accounted for by cellular protein, RNA, or DNA and released large amounts of nonammonia nitrogen. The difference between net and true growth was most dramatic at low dilution rates, but mathematical derivations indicated that the lysis rate was a growth rate-independent function. The lysis rate was sevenfold greater than the true maintenance rate (0.07 h-1 versus 0.01 h-1). Because slowly growing cells had as much proton motive force and ATP as fast-growing cells, lysis was not a starvation response per se. Stationary-phase cells had a lysis rate that was 10-fold less than that of growing cells. Rapidly growing cells were not susceptible to phenylmethylsulfonyl fluoride, but phenylmethylsulfonyl fluoride increased the lysis rate of the cultures when they reached the stationary phase. This latter result indicated that autolysins of stationary-phase cells were being inactivated by a serine proteinase. When growing cells were treated with the glycolytic inhibitor iodoacetate, the proteinase-dependent transition to the stationary phase was circumvented, and the rate of lysis could be increased by as much as 50-fold.  相似文献   

14.
Because of the highly conserved pattern of expression of the eucaryotic heat shock genes hsp70 and hsp84 or their cognates during sporulation in Saccharomyces cerevisiae and development in higher organisms, the role of the Escherichia coli homologs dnaK and htpG was examined during the response to starvation. The htpG deletion mutant was found to be similar to its wild-type parent in its ability to survive starvation for essential nutrients and to induce proteins specific to starvation conditions. The dnaK103 mutant, however, was highly susceptible to killing by starvation for carbon and, to a lesser extent, for nitrogen and phosphate. Analysis of proteins induced under starvation conditions on two-dimensional gels showed that the dnaK103 mutant was defective for the synthesis of some proteins induced in wild-type cells by carbon starvation and of some proteins induced under all starvation conditions, including the stationary phase in wild-type cells. In addition, unique proteins were synthesized in the dnaK103 mutant in response to starvation. Although the synthesis of some proteins under glucose starvation control was drastically affected by the dnaK103 mutation, the synthesis of proteins specifically induced by nitrogen starvation was essentially unaffected. Similarly, the dnaK103 mutant was able to grow, utilizing glutamine or arginine as a source of nitrogen, at a rate approximate to that of the wild-type parent, but it inefficiently utilized glycerol or maltose as carbon sources. Several differences between the protein synthetic pattern of the dnaK103 mutant and the wild type were observed after phosphate starvation, but these did not result in a decreased ability to survive phosphate starvation, compared with nitrogen starvation.  相似文献   

15.
16.
Acetylcoenzyme A synthase/carbon monoxide dehydrogenase (ACS/CODH) contains two Ni–Fe–S active-site clusters (called A and C) connected by a tunnel through which CO and CO2 migrate. Site-directed mutants A578C, L215F, and A219F were designed to block the tunnel at different points along the region between the two C-clusters. Two other mutant proteins F70W and N101Q were designed to block the region that connects the tunnel at the ββ interface with a water channel also located at that interface. Purified mutant proteins were assayed for Ni/Fe content and examined by electron paramagnetic resonance spectroscopy. Analyses indicate that same metal clusters found in wild-type (WT) ACS/CODH (i.e., the A-, B-, C-, and probably D-clusters) are properly assembled in the mutant enzymes. Stopped-flow kinetics revealed that these centers in the mutants are rapidly reducible by dithionite but are only slowly reducible by CO, suggesting an impaired ability of CO to migrate through the tunnel to the C-cluster. Relative to the WT enzyme, mutant proteins exhibited little CODH or ACS activity (using CO2 as a substrate). Some ACS activity was observed when CO was a substrate, but not the cooperative CO inhibition effect characteristic of WT ACS/CODH. These results suggest that CO and CO2 enter and exit the enzyme at the water channel along the ββ subunit interface. They also suggest two pathways for CO during synthesis of acetylcoenzyme A, including one in which CO enters the enzyme and migrates through the tunnel before binding at the A-cluster, and another in which CO binds the A-cluster directly from the solvent.  相似文献   

17.
18.
J Heo  C R Staples  P W Ludden 《Biochemistry》2001,40(25):7604-7611
Carbon monoxide dehydrogenase (CODH) from Rhodospirillum rubrum catalyzes both the oxidation of CO and the reduction of CO(2). Studies of the redox dependence of CO(2) reduction by R. rubrum CODH show that (1) CODH is unable to catalyze CO(2) reduction at potentials greater than -300 mV; (2) the maximum activity is observed at potentials less than -480 mV; and (3) the midpoint potential (E(m)) of the transition from minimum to maximum CO(2) reduction activity occurs at approximately -339 mV. These results indicate that the C(red1) state of R. rubrum CODH (E(m) = -110 mV; g(zyx)() = 2.03, 1.88, 1.71) is not competent to reduce CO(2). Nernst analyses suggest that the reduction of CODH from the C(red1) state to the CO(2)-reducing form (C(unc), g(zyx)() = 2.04, 1.93, 1.89; E < approximately -300 mV) of the enzyme is a one-electron process. For the entire redox range, viologens stimulate CO(2) reduction by CODH more than 50-fold, and it is proposed that viologens accelerate the redox equilibration of redox buffers and [Fe(4)S(4)](B) during catalysis.  相似文献   

19.
Regulation of phenylalanine ammonia lyase in Rhodotorula glutinis.   总被引:5,自引:0,他引:5       下载免费PDF全文
In the red yeast Rhodotorula glutinis, phenylalanine ammonia lyase (PAL) was induced 10-fold during carbon starvation even in the absence of exogenous phenylalanine, although maximal induction occurred when phenylalanine was the nitrogen (40-fold) or carbon (100-fold) source. Apparent regulatory mutations that affected the expression of PAL were isolated by selecting mutants resistant to the analog p-fluoro-D,L-phenylalanine (PFP). One such mutant, designated FP1, could use phenylalanine as a nitrogen source but not as a carbon source. Similarly, FP1 failed to utilize intermediates of the phenylalanine degradative pathway, namely, benzoate, p-hydroxybenzoate, or 3,4-dihydroxybenzoate, as carbon sources. Although the PFP-resistant mutant contained a low level of PAL, no increase was found when it was grown with phenylalanine as the nitrogen source. A derivative of FP1, FP1a, was isolated that simultaneously regained an inducible PAL and the ability to use phenylalanine, benzoate, p-hydroxybenzoate, and 3,4-dihydroxybenzoate as carbon sources. In addition, when p-hydroxybenzoate was the carbon source, PAL was induced in the mutant FP1a but not in the PFP-sensitive parental strain. We propose that the mutation to PFP resistance occurred in a regulatory gene that controls the entire phenylalanine degradative pathway. Secondary mutations at this locus, as found in strain FP1a, not only restored expression of this pathway, but also altered the induction of PAL by metabolites of this pathway.  相似文献   

20.
The effects of nitrogen starvation on biomass composition and photosynthetic function were examined in the marine cryptophyte Rhodomonas sp. Batch-cultured cells in N-sufficient medium showed a 2.5-fold increase in total carbohydrate content, and a 33% increase in cell volume when the cultures reached the stationary growth phase. These cultures also increased the ratio of phycoerythrin (PE)/hydrosoluble proteins from 6 to 22% by the 4th and 10th day of culture, respectively. In contrast, light-saturated photosynthetic activity (Pm) progressively decreased, and the value obtained at the beginning of the stationary phase was about 45% of that obtained for cells in the late exponential growth phase. Transfer to N-lacking medium caused a 3.2-fold increase in cell volume. N starvation also triggered a rapid decline in N-containing compounds such as hydrosoluble proteins and photosynthetic pigments, causing an almost complete loss of PE. The ratio of PE/hydrosoluble proteins decreased from 6 to 1% after 6 d of N deprivation. Furthermore, the PSII fluorescence capacity declined under N-starved conditions, which caused a pronounced decrease in both the Pm (circa 90%) and the apparent photosynthetic efficiency (circa 55%). Under these conditions, photosynthetically fixed carbon was used to synthesize large amounts of carbohydrates. We suggest that, in addition to the role of phycoerythrin as a light-harvesting pigment, Rhodomonas sp. responds to N-depleted conditions by mobilizing combined nitrogen from biliproteins.  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