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 218 毫秒
1.
Incubation of resting cells of Sphingobium indicum B90A, Sphingobium japonicum UT26, and Sphingobium francense Sp+ showed that they were able to transform β- and δ-hexachlorocyclohexane (β- and δ-HCH, respectively), the most recalcitrant hexachlorocyclohexane isomers, to pentachlorocyclohexanols, but only resting cells of strain B90A could further transform the pentachlorocyclohexanol intermediates to the corresponding tetrachlorocyclohexanediols. Moreover, experiments with resting cells of Escherichia coli expressing the LinB proteins of strains B90A, UT26, and Sp+ indicated that LinB was responsible for these transformations. Purified LinB proteins from all three strains also effected the formation of the respective pentachlorocyclohexanols. Although the three LinB enzymes differ only marginally with respect to amino acid sequence, they showed interesting differences with respect to substrate specificity. When LinB from strain B90A was incubated with β- and δ-HCH, the pentachlorocyclohexanol products were further transformed and eventually disappeared from the incubation mixtures. In contrast, the LinB proteins from strains UT26 and Sp+ could not catalyze transformation of the pentachlorocyclohexanols, and these products accumulated in the incubation mixture. A mutant of strain Sp+ lacking linA and linB did not degrade any of the HCH isomers, including β-HCH, and complementation of this mutant by linB from strain B90A restored the ability to degrade β- and δ-HCH.  相似文献   

2.
β-Hexachlorocyclohexane (β-HCH) is the most recalcitrant among the α-, β-, γ-, and δ-isomers of HCH and causes serious environmental pollution problems. We demonstrate here that the haloalkane dehalogenase LinB, reported earlier to mediate the second step in the degradation of γ-HCH in Sphingomonas paucimobilis UT26, metabolizes β-HCH to produce 2,3,4,5,6-pentachlorocyclohexanol.  相似文献   

3.
The alpha-proteobacterial strain Sphingobium japonicum UT26 utilizes a highly chlorinated pesticide, gamma-hexachlorocyclohexane (gamma-HCH), as a sole source of carbon and energy, and haloalkane dehalogenase LinB catalyzes the second step of gamma-HCH degradation in UT26. Functional complementation of a linB mutant of UT26, UT26DB, was performed by the exogenous plasmid isolation technique using HCH-contaminated soil, leading to our successful identification of a plasmid, pLB1, carrying the linB gene. Complete sequencing analysis of pLB1, with a size of 65,998 bp, revealed that it carries (i) 50 totally annotated coding sequences, (ii) an IS6100 composite transposon containing two copies of linB, and (iii) potential genes for replication, maintenance, and conjugative transfer with low levels of similarity to other homologues. A minireplicon assay demonstrated that a 2-kb region containing the predicted repA gene and its upstream region of pLB1 functions as an autonomously replicating unit in UT26. Furthermore, pLB1 was conjugally transferred from UT26DB to other alpha-proteobacterial strains but not to any of the beta- or gamma-proteobacterial strains examined to date. These results suggest that this exogenously isolated novel plasmid contributes to the dissemination of at least some genes for gamma-HCH degradation in the natural environment. To the best of our knowledge, this is the first detailed report of a plasmid involved in gamma-HCH degradation.  相似文献   

4.
Sphingobium japonicum strain UT26 utilizes γ-hexachlorocyclohexane (γ-HCH), a man-made chlorinated pesticide that causes serious environmental problems due to its toxicity and long persistence, as a sole source of carbon and energy. Here, we report the complete genome sequence of UT26, which consists of two chromosomes and three plasmids. The 15 lin genes involved in γ-HCH degradation are dispersed on the two chromosomes and one of the three plasmids.γ-Hexachlorocylohexane (γ-HCH) is a man-made insecticide that has caused serious environmental problems worldwide (9). Although only several decades have passed since the initial release of γ-HCH into the environment, many γ-HCH-degrading bacterial strains have been isolated (9), suggesting that γ-HCH-degrading ability was acquired by such strains within a short period (14). Sphingobium japonicum UT26 was isolated from upland γ-HCH-contaminated soil in Japan and utilizes γ-HCH as its sole source of carbon and energy (8, 16). In UT26, 15 lin genes are involved in γ-HCH utilization (2, 11, 12). Our clarification of the complete genome sequence of this strain is expected to provide insights into the mechanisms by which bacteria adapt to xenobiotics.One S. japonicum UT26 colony was designated UT26S (NBRC 101211), and its complete genome sequence was determined by a whole-genome shotgun sequencing strategy using the Sanger method (10, 15, 17). The sequences of ca. 92,400 reads were assembled by the Phrap and CONSED assembly tools (4, 5, 7), and the gaps between contigs were closed by sequencing PCR products which were amplified from genomic DNA using the appropriate primers. The prediction and annotation of open reading frames (ORFs) were performed using Glimmer3 (1), BLASTP, the In Silico Molecular Cloning software suite (In Silico Biology, Inc.), and the GenomeMatcher software (13). Nontranslating genes were predicted using the Rfam, tRNAscan-SE, and ARAGORN programs.The genome of S. japonicum UT26S consists of two circular chromosomes (Chr), Chr 1 (3,514,822 bp, 64.8% G+C, 3,529 ORFs) and Chr 2 (681,892 bp, 65.9% G+C, 589 ORFs), and three circular plasmids, pCHQ1 (190,974 bp, 63.0% G+C, 224 ORFs), pUT1 (31,776 bp, 63.7% G+C, 44 ORFs), and pUT2 (5,398 bp, 61.0% G+C, 8 ORFs). Chr 1 and Chr 2 have one and two copies, respectively, of rRNA operons. Fifty-one and 4 tRNA genes were located on Chr 1 and Chr 2, respectively. One hundred ninety-six out of 206 bacterial essential genes proposed by Gil et al. (6) were all located on Chr 1, indicating that this is clearly a “main” chromosome. The 15 lin genes for γ-HCH degradation are dispersed on Chr 1, Chr 2, and pCHQ1. Comparison of the UT26S genome with those of five other sphingomonad (TM1) strains revealed that the lin genes (linA, linB, linC, linRED, and linF) specific for the conversion of γ-HCH to β-ketoadipate are located in the DNA regions unique to the UT26S genome. On the other hand, linGHIJ for the β-ketoadipate pathway (12) and linKLMN for the ABC transporter system (3) are located in conserved genomic regions of these sphingomonads. Based on these results, we propose a model in which UT26S was established by recruiting the specific lin genes into an ancestral strain having core functions of sphingomonads.  相似文献   

5.
Sphingomonas (formerly Pseudomonas) paucimobilis UT26 utilizes γ-hexachlorocyclohexane (γ-HCH), a halogenated organic insecticide, as a sole carbon and energy source. In a previous study, we showed that γ-HCH is degraded to 2,5-dichlorohydroquinone (2,5-DCHQ) (Y. Nagata, R. Ohtomo, K. Miyauchi, M. Fukuda, K. Yano, and M. Takagi, J. Bacteriol. 176:3117–3125, 1994). In the present study, we cloned and characterized a gene, designated linD, directly involved in the degradation of 2,5-DCHQ. The linD gene encodes a peptide of 343 amino acids and has a low level of similarity to proteins which belong to the glutathione S-transferase family. When LinD was overproduced in Escherichia coli, a 40-kDa protein was found after sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Northern blot analysis revealed that expression of the linD gene was induced by 2,5-DCHQ in S. paucimobilis UT26. Thin-layer chromatography and gas chromatography-mass spectrometry analyses with the LinD-overexpressing E. coli cells revealed that LinD converts 2,5-DCHQ rapidly to chlorohydroquinone (CHQ) and also converts CHQ slowly to hydroquinone. LinD activity in crude cell extracts was increased 3.7-fold by the addition of glutathione. All three of the Tn5-induced mutants of UT26, which lack 2,5-DCHQ dehalogenase activity, had rearrangements or a deletion in the linD region. These results indicate that LinD is a glutathione-dependent reductive dehalogenase involved in the degradation of γ-HCH by S. paucimobilis UT26.  相似文献   

6.
The technical formulation of hexachlorocyclohexane (HCH) mainly consists of the insecticidal γ-isomer and noninsecticidal α-, β-, and δ-isomers, among which β-HCH is the most recalcitrant and has caused serious environmental problems. A γ-HCH-utilizing bacterial strain, Sphingobium sp. MI1205, was isolated from soil which had been contaminated with HCH isomers. This strain degraded β-HCH more rapidly than the well-characterized γ-HCH-utilizing strain Sphingobium japonicum UT26. In MI1205, β-HCH was converted to 2,3,5,6-tetrachlorocyclohexane-1,4-diol (TCDL) via 2,3,4,5,6-pentachlorocyclohexanol (PCHL). A haloalkane dehalogenase LinB (LinBMI) that is 98% identical (seven amino-acid differences among 296 amino acids) to LinB from UT26 (LinBUT) was identified as an enzyme responsible for the two-step conversion of β-HCH to TCDL. This property of LinBMI contrasted with that of LinBUT, which catalyzed only the first step conversion of β-HCH to PCHL. Site-directed mutagenesis and computer modeling suggested that two of the seven different amino acid residues (V134 and H247) forming a catalytic pocket of LinB are important for the binding of PCHL in an orientation suitable for the reaction in LinBMI. However, mutagenesis also indicated the involvement of other residues for the activity unique to LinBMI. Sequence analysis revealed that MI1205 possesses the IS6100-flanked cluster that contains two copies of the linB MI gene. This cluster is identical to the one located on the exogenously isolated plasmid pLB1, suggesting that MI1205 had recruited the linB genes by a horizontal transfer event.  相似文献   

7.
Sphingomonas paucimobilis B90A contains two variants, LinA1 and LinA2, of a dehydrochlorinase that catalyzes the first and second steps in the metabolism of hexachlorocyclohexanes (R. Kumari, S. Subudhi, M. Suar, G. Dhingra, V. Raina, C. Dogra, S. Lal, J. R. van der Meer, C. Holliger, and R. Lal, Appl. Environ. Microbiol. 68:6021-6028, 2002). On the amino acid level, LinA1 and LinA2 were 88% identical to each other, and LinA2 was 100% identical to LinA of S. paucimobilis UT26. Incubation of chiral α-hexachlorocyclohexane (α-HCH) with Escherichia coli BL21 expressing functional LinA1 and LinA2 S-glutathione transferase fusion proteins showed that LinA1 preferentially converted the (+) enantiomer, whereas LinA2 preferred the (−) enantiomer. Concurrent formation and subsequent dissipation of β-pentachlorocyclohexene enantiomers was also observed in these experiments, indicating that there was enantioselective formation and/or dissipation of these enantiomers. LinA1 preferentially formed (3S,4S,5R,6R)-1,3,4,5,6-pentachlorocyclohexene, and LinA2 preferentially formed (3R,4R,5S,6S)-1,3,4,5,6-pentachlorocyclohexene. Because enantioselectivity was not observed in incubations with whole cells of S. paucimobilis B90A, we concluded that LinA1 and LinA2 are equally active in this organism. The enantioselective transformation of chiral α-HCH by LinA1 and LinA2 provides the first evidence of the molecular basis for the changed enantiomer composition of α-HCH in many natural environments. Enantioselective degradation may be one of the key processes determining enantiomer composition, especially when strains that contain only one of the linA genes, such as S. paucimobilis UT26, prevail.  相似文献   

8.
A gram-positive Microbacterium sp. strain, ITRC1, that was able to degrade the persistent and toxic hexachlorocyclohexane (HCH) isomers was isolated and characterized. The ITRC1 strain has the capacity to degrade all four major isomers of HCH present in both liquid cultures and aged contaminated soil. DNA fragments corresponding to the two initial genes involved in γ-HCH degradative pathway, encoding enzymes for γ-pentachlorocyclohexene hydrolytic dehalogenase (linB) and a 2,5-dichloro-2,5-cyclohexadiene-1,4-diol dehydrogenase (linC), were amplified by PCR and sequenced. Their presence in the ITRC1 genomic DNA was also confirmed by Southern hybridization. Sequencing of the amplified DNA fragment revealed that the two genes present in the ITRC1 strain were homologous to those present in Sphingomonas paucimobilis UT26. Both 16S rRNA sequencing and phylogenetic analysis resulted in the identification of the bacteria as a Microbacterium sp. We assume that these HCH-degrading bacteria evolved independently but possessed genes similar to S. paucimobilis UT26. The reported results indicate that catabolic genes for γ-HCH degradation are highly conserved in diverse genera of bacteria, including the gram-positive groups, occurring in various environmental conditions.  相似文献   

9.
The complete genome sequencing of a γ-hexachlorocyclohexane-degrading strain, Sphingobium japonicum UT26, revealed that the genome consists of two circular chromosomes [with sizes of 3.5 Mb (Chr1) and 682 kb (Chr2)], a 191-kb large plasmid (pCHQ1), and two small plasmids with sizes of 32 and 5 kb. The lin genes are dispersed on Chr1, Chr2, and pCHQ1. Comparison of the UT26 genome with those of other sphingomonad strains demonstrated that the “specific” lin genes for conversion of γ-HCH to β-ketoadipate (linA, linB, linC, linRED, and linF) are located on the DNA regions unique to the UT26 genome, suggesting the acquisition of these lin genes by horizontal transfer events. On the other hand, linGHIJ and linKLMN are located on the regions conserved in the genomes of sphingomonads, suggesting that the linGHIJ-encoded β-ketoadipate pathway and the LinKLMN-type ABC transporter system are involved in core functions of sphingomonads. Based on these results, we propose a hypothesis that UT26 was created by recruiting the specific lin genes into a strain having core functions of sphingomonads. Most of the specific lin genes in UT26 are associated with IS6100. Our analysis of spontaneous linA-, linC-, and linRED-deletion mutants of UT26 revealed the involvement of IS6100 in their deduced genome rearrangements. These facts strongly suggest that IS6100 plays important roles both in the dissemination of the specific lin genes and in the genome rearrangements.  相似文献   

10.
11.
This study investigated the feasibility of a slow-release inoculation approach as a bioaugmentation strategy for the degradation of lindane (γ-hexachlorocyclohexane [γ-HCH]). Slow-release inoculation of Sphingomonas sp. γ 1-7 was established in both liquid and soil slurry microcosms using open-ended silicone tubes in which the bacteria are encapsulated in a protective nutrient-rich matrix. The capacity of the encapsulated cells to degrade lindane under aerobic conditions was evaluated in comparison with inoculation of free-living cells. Encapsulation of cells in tubes caused the removal of lindane by adsorption to the silicone tubes but also ensured prolonged biodegradation activity. Lindane degradation persisted 2.2 and 1.4 times longer for liquid and soil slurry microcosms, respectively, than that for inoculation with free cells. While inoculation of free-living cells led to a loss in lindane-degrading activity in limited time intervals, encapsulation in tubes allowed for a more stable actively degrading community. The loss in degrading activity was linked to the loss of the linA gene, encoding γ-HCH dehydrochlorinase (LinA), which is involved in the initial steps of the lindane degradation pathway. This work shows that a slow-release inoculation approach using a catabolic strain encapsulated in open-ended tubes is a promising bioaugmentation tool for contaminated sites, as it can enhance pollutant removal and can prolong the degrading activity in comparison with traditional inoculation strategies.  相似文献   

12.
Urease-Encoding Genes in Ammonia-Oxidizing Bacteria   总被引:1,自引:1,他引:0       下载免费PDF全文
Many but not all ammonia-oxidizing bacteria (AOB) produce urease (urea amidohydrolase, EC 3.5.1.5) and are capable of using urea for chemolithotrophic growth. We sequenced the urease operons from two AOB, the β-proteobacterium Nitrosospira sp. strain NpAV and the γ-proteobacterium Nitrosococcus oceani. In both organisms, all seven urease genes were contiguous: the three structural urease genes ureABC were preceded and succeeded by the accessory genes ureD and ureEFG, respectively. Green fluorescent protein reporter gene fusions revealed that the ure genes were under control of a single operon promoter upstream of the ureD gene in Nitrosococcus oceani. Southern analyses revealed two copies of ureC in the Nitrosospira sp. strain NpAV genome, while a single copy of the ure operon was detected in the genome of Nitrosococcus oceani. The ureC gene encodes the alpha subunit protein containing the active site and conserved nickel binding ligands; these conserved regions were suitable primer targets for obtaining further ureC sequences from additional AOB. In order to develop molecular tools for detecting the ureolytic ecotype of AOB, ureC genes were sequenced from several β-proteobacterial AOB. Pairwise identity values ranged from 80 to 90% for the UreC peptides of AOB within a subdivision. UreC sequences deduced from AOB urease genes and available UreC sequences in the public databases were used to construct alignments and make phylogenetic inferences. The UreC proteins from β-proteobacterial AOB formed a distinct monophyletic group. Unexpectedly, the peptides from AOB did not group most closely with the UreC proteins from other β-proteobacteria. Instead, it appears that urease in β-proteobacterial autotrophic ammonia oxidizers is the product of divergent evolution in the common ancestor of γ- and β-proteobacteria that was initiated before their divergence during speciation. Sequence motifs conserved for the proteobacteria and variable regions possibly discriminatory for ureC from β-proteobacterial AOB were identified for future use in environmental analysis of ureolytic AOB. These gene sequences are the first publicly available for ure genes from autotrophic AOB.  相似文献   

13.
The α-proteobacterium Wolbachia pipientis infects more than 65% of insect species worldwide and manipulates the host reproductive machinery to enable its own survival. It can live in mutualistic relationships with hosts that cause human disease, including mosquitoes that carry the Dengue virus. Like many other bacteria, Wolbachia contains disulfide bond forming (Dsb) proteins that introduce disulfide bonds into secreted effector proteins. The genome of the Wolbachia strain wMel encodes two DsbA-like proteins sharing just 21% sequence identity to each other, α-DsbA1 and α-DsbA2, and an integral membrane protein, α-DsbB. α-DsbA1 and α-DsbA2 both have a Cys-X-X-Cys active site that, by analogy with Escherichia coli DsbA, would need to be oxidized to the disulfide form to serve as a disulfide bond donor toward substrate proteins. Here we show that the integral membrane protein α-DsbB oxidizes α-DsbA1, but not α-DsbA2. The interaction between α-DsbA1 and α-DsbB is very specific, involving four essential cysteines located in the two periplasmic loops of α-DsbB. In the electron flow cascade, oxidation of α-DsbA1 by α-DsbB is initiated by an oxidizing quinone cofactor that interacts with the cysteine pair in the first periplasmic loop. Oxidizing power is transferred to the second cysteine pair, which directly interacts with α-DsbA1. This reaction is inhibited by a non-catalytic disulfide present in α-DsbA1, conserved in other α-proteobacterial DsbAs but not in γ-proteobacterial DsbAs. This is the first characterization of the integral membrane protein α-DsbB from Wolbachia and reveals that the non-catalytic cysteines of α-DsbA1 regulate the redox relay system in cooperation with α-DsbB.  相似文献   

14.
Objectives: To assess the Gene-Environmental interaction between maternal organochlorine pesticides (OCPs) level and CYP17 gene polymorphism with the risk of preterm delivery (PTD). Materials and methods: Maternal blood samples of hundred cases (n = 100) of PTD and of equal number of healthy controls were collected at the time of delivery. OCPs levels were estimated by Gas chromatography system equipped with electron capture detector and PCR-RFLP was used for polymorphic analysis of CYP17 gene. Results: Significantly (p < 0.05) higher levels of α-HCH, β-HCH, and γ-HCH were found in maternal blood samples of PTD cases as compared to controls. We did not found any significant difference in the frequency genotype distribution CYP17 gene in PTD cases as compared to controls. When gene environmental interaction between the CYP17 gene polymorphism and OCPs level was considered, a significant interaction was observed between ≥ 50th percentile of γ-HCH and CYP17 A1A1 (wild type) genotype. Conclusions: Higher levels of OCPs along with wild type state of CYP17 gene (A1A1) in women may be considered as an important etiological factor in ‘idiopathic’ PTD. The present study provides evidence that genetic variation and its interaction with the environmental exposure may increase the risk of PTD.  相似文献   

15.
LinA is the first enzyme of the microbial degradation pathway of a chlorinated insecticide, hexachlorocyclohexane (HCH), and mediates the dehydrochlorination of α-, γ-, and δ-HCH. Its two variants, LinA type 1 and LinA type 2, which differ at 10 out of 156 amino acid residues, have been described. Their activities for the metabolism of different HCH isomers differ considerably but overall are high for γ-HCH, moderate for α-HCH, low for δ-HCH, and lacking for β-HCH. Here, we describe the characterization of a new variant of this enzyme, LinA type 3, whose gene was identified from the metagenome of an HCH-contaminated soil sample. Its deduced primary structure in the region spanning amino acid residues 1 to 147 of the protein exhibits 17 and 12 differences from LinA type 1 and LinA type 2, respectively. In addition, the residues GIHFAPS, present at the region spanning residues 148 to 154 in both LinA type 1 and LinA type 2, are deleted in LinA type 3.The activity of LinA type 3 for the metabolism of δ-HCH is several orders of magnitude higher than that of LinA type 1 or LinA type 2 and can be useful for improvement of the metabolism of δ-HCH.  相似文献   

16.
The enzymes LinBUT and LinBMI (LinB from Sphingobium japonicum UT26 and Sphingobium sp. MI1205, respectively) catalyze the hydrolytic dechlorination of β-hexachlorocyclohexane (β-HCH) and yield different products, 2,3,4,5,6-pentachlorocyclohexanol (PCHL) and 2,3,5,6-tetrachlorocyclohexane-1,4-diol (TCDL), respectively, despite their 98% identity in amino acid sequence. To reveal the structural basis of their different enzymatic properties, we performed site-directed mutagenesis and X-ray crystallographic studies of LinBMI and its seven point mutants. The mutation analysis revealed that the seven amino acid residues uniquely found in LinBMI were categorized into three groups based on the efficiency of the first-step (from β-HCH to PCHL) and second-step (from PCHL to TCDL) conversions. Crystal structure analyses of wild-type LinBMI and its seven point mutants indicated how each mutated residue contributed to the first- and second-step conversions by LinBMI. The dynamics simulation analyses of wild-type LinBMI and LinBUT revealed that the entrance of the substrate access tunnel of LinBUT was more flexible than that of LinBMI, which could lead to the different efficiencies of dehalogenation activity between these dehalogenases.  相似文献   

17.
Bioremediation of anthropogenic organic pollutants in cold climates is often limited by lower microbial or enzyme activity induced by low temperature. The present study addressed this issue through the degradation of ??-hexachlorocyclohexane (??-HCH) by three Sphingobium strains (S. indicum B90A, S. japonicum UT26 and S. francense Sp+) under low temperature (4 °C). After 5 days incubation at 4 °C, 79.7% and 43.8% of 5 and 25 mg L−1 of ??-HCH added were degraded, respectively by the inoculation of 1.75 × 107 cells mL−1 of S. indicum B90A. An increase in inoculum concentration to 1.72 × 108 cells mL−1 significantly increased the degradation to 98.1 ± 1.7% of 5 mg L−1 within 24 h. Further, S. indicum B90A and S. japonicum UT26 can rapidly degrade ??-HCH at 4 °C, while the degradation capability of S. francense Sp+ is relatively low. At 4 °C, ??-HCH is transformed to extremely low amounts of 1,2,4-trichlorobenzene (1,2,4-TCB) and 2,5-dichlorophenol (2,5-DCP) by S. indicum B90A, but most of ??-HCH were transformed to 2,5-Dichloro-2,5-cyclohexadiene-1,4-diol (2,5-DDOL) by S. japonicum UT26. These results revealed that haloalkane dehalogenases in some Sphingobium species are very active at temperature as low as 4 °C and S. indicum B90A might be a good candidate for developing novel bioremediation techniques for cold regions to decontaminate ??-HCH from soils/waters.  相似文献   

18.
The period when the snowpack melts in late spring is a dynamic time for alpine ecosystems. The large winter microbial community begins to turn over rapidly, releasing nutrients to plants. Past studies have shown that the soil microbial community in alpine dry meadows of the Colorado Rocky Mountains changes in biomass, function, broad-level structure, and fungal diversity between winter and early summer. However, little specific information exists on the diversity of the alpine bacterial community or how it changes during this ecologically important period. We constructed clone libraries of 16S ribosomal DNA from alpine soil collected in winter, spring, and summer. We also cultivated bacteria from the alpine soil and measured the seasonal abundance of selected cultured isolates in hybridization experiments. The uncultured bacterial communities changed between seasons in diversity and abundance within taxa. The Acidobacterium division was most abundant in the spring. The winter community had the highest proportion of Actinobacteria and members of the Cytophaga/Flexibacter/Bacteroides (CFB) division. The summer community had the highest proportion of the Verrucomicrobium division and of β-Proteobacteria. As a whole, α-Proteobacteria were equally abundant in all seasons, although seasonal changes may have occurred within this group. A number of sequences from currently uncultivated divisions were found, including two novel candidate divisions. The cultured isolates belonged to the α-, β-, and γ-Proteobacteria, the Actinobacteria, and the CFB groups. The only uncultured sequences that were closely related to the isolates were from winter and spring libraries. Hybridization experiments showed that actinobacterial and β-proteobacterial isolates were most abundant during winter, while the α- and γ-proteobacterial isolates tested did not vary significantly. While the cultures and clone libraries produced generally distinct groups of organisms, the two approaches gave consistent accounts of seasonal changes in microbial diversity.  相似文献   

19.
Mitochondrial DNA (mtDNA) is unquestionably the remnant of an α-proteobacterial genome, yet only ∼10%–20% of mitochondrial proteins are demonstrably α-proteobacterial in origin (the “α-proteobacterial component,” or APC). The evolutionary ancestry of the non-α-proteobacterial component (NPC) is obscure and not adequately accounted for in current models of mitochondrial origin. I propose that in the host cell that accommodated an α-proteobacterial endosymbiont, much of the NPC was already present, in the form of a membrane-bound metabolic organelle (the premitochondrion) that compartmentalized many of the non-energy-generating functions of the contemporary mitochondrion. I suggest that this organelle also possessed a protein import system and various ion and small-molecule transporters. In such a scenario, an α-proteobacterial endosymbiont could have been converted relatively directly and rapidly into an energy-generating organelle that incorporated the extant metabolic functions of the premitochondrion. This model (the “pre-endosymbiont hypothesis”) effectively represents a synthesis of previous, contending mitochondrial origin hypotheses, with the bulk of the mitochondrial proteome (much of the NPC) having an endogenous origin and the minority component (the APC) having a xenogenous origin.Considering the central role played in all eukaryotic cells by mitochondria or mitochondrion-related organelles (MROs, such as hydrogenosomes and mitosomes) (Hjort et al. 2010; Shiflett and Johnson 2010; Müller et al. 2012), the question of the origin and subsequent evolution of the mitochondrion has long captivated and challenged biologists. In a recent article in this series (Gray 2012), I discussed in detail several aspects of mitochondrial evolution, focusing particularly on how well the accumulating molecular data can be accommodated in current models of mitochondrial origin. In this context, the origin and evolution of the mitochondrial proteome, as opposed to the origin and evolution of the mitochondrial genome, were examined from the perspective of comparative mitochondrial proteomics. Somewhat disconcertingly, as more data have become available, we find ourselves considerably less certain about key aspects of how mitochondria originated than we were (or thought we were) several decades ago.Here, I summarize key points discussed in more detail in the previous article before presenting a novel perspective on how the mitochondrion might have originated. The new model proposed here, which represents a synthesis of both endogenous (“origin from within”) and xenogenous (“origin from outside”) modes, is advanced in an attempt to account for the inability of a purely endosymbiotic model, whose strongest support has come from studies of the mitochondrial genome, to adequately accommodate data on the mitochondrial proteome.  相似文献   

20.
Partially purified cell wall proteinases of eight strains of Streptococcus cremoris were compared in their action on bovine αs1-, β-, and κ-casein, as visualized by starch gel electrophoresis, sodium dodecyl sulfate-polyacrylamide gel electrophoresis, and thin-layer chromatography. Characteristic degradation profiles could be distinguished, from which the occurrence of two proteinases, represented by strain HP and strain AM1, was concluded. The action of the HP-type proteinase P1 (also detectable in strains Wg2, C13, and TR) was established by electrophoretic methods to be directed preferentially towards β-casein. The AM1-type proteinase PIII (also detectable in strain SK11) was also able to degrade β-casein, but at the same time split αs1- and κ-casein more extensively than did PI. Strain FD27 exhibited mainly PI activity but also detectable PIII degradation characteristics. The cell wall proteinase preparation of strain E8 showed low PI as well as low PIII activity. All proteinase preparations produced from κ-casein positively charged degradation products with electrophoretic mobilities similar to those of degradation products released by the action of the milk-clotting enzyme chymosin. The differences between PI and PIII in mode of action, as detected by gel electrophoresis and thin-layer chromatography, were reflected by the courses of the initial degradation of methyl-14C-labeled β-casein and by the effect of αs1- plus κ-casein on these degradations. The results are discussed in the light of previous comparative studies of cell wall proteinases in strains of S. cremoris and with respect to the growth of this organism in milk.  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