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.
Phenoxyalkanoic acid (PAA) herbicides are widely used in agriculture. Biotic degradation of such herbicides occurs in soils and is initiated by α-ketoglutarate- and Fe2+-dependent dioxygenases encoded by tfdA-like genes (i.e., tfdA and tfdAα). Novel primers and quantitative kinetic PCR (qPCR) assays were developed to analyze the diversity and abundance of tfdA-like genes in soil. Five primer sets targeting tfdA-like genes were designed and evaluated. Primer sets 3 to 5 specifically amplified tfdA-like genes from soil, and a total of 437 sequences were retrieved. Coverages of gene libraries were 62 to 100%, up to 122 genotypes were detected, and up to 389 genotypes were predicted to occur in the gene libraries as indicated by the richness estimator Chao1. Phylogenetic analysis of in silico-translated tfdA-like genes indicated that soil tfdA-like genes were related to those of group 2 and 3 Bradyrhizobium spp., Sphingomonas spp., and uncultured soil bacteria. Soil-derived tfdA-like genes were assigned to 11 clusters, 4 of which were composed of novel sequences from this study, indicating that soil harbors novel and diverse tfdA-like genes. Correlation analysis of 16S rRNA and tfdA-like gene similarity indicated that any two bacteria with D > 20% of group 2 tfdA-like gene-derived protein sequences belong to different species. Thus, data indicate that the soil analyzed harbors at least 48 novel bacterial species containing group 2 tfdA-like genes. Novel qPCR assays were established to quantify such new tfdA-like genes. Copy numbers of tfdA-like genes were 1.0 × 106 to 65 × 106 per gram (dry weight) soil in four different soils, indicating that hitherto-unknown, diverse tfdA-like genes are abundant in soils.Phenoxyalkanoic acid (PAA) herbicides such as MCPA (4-chloro-2-methyl-phenoxyacetic acid) and 2,4-D (2,4-dichlorophenoxyacetic acid) are widely used to control broad-leaf weeds in agricultural as well as nonagricultural areas (19, 77). Degradation occurs primarily under oxic conditions in soil, and microorganisms play a key role in the degradation of such herbicides in soil (62, 64). Although relatively rapidly degraded in soil (32, 45), both MCPA and 2,4-D are potential groundwater contaminants (10, 56, 70), accentuating the importance of bacterial PAA herbicide-degrading bacteria in soils (e.g., references 3, 5, 6, 20, 41, 59, and 78).Degradation can occur cometabolically or be associated with energy conservation (15, 54). The first step in the degradation of 2,4-D and MCPA is initiated by the product of cadAB or tfdA-like genes (29, 30, 35, 67), which constitutes an α-ketoglutarate (α-KG)- and Fe2+-dependent dioxygenase. TfdA removes the acetate side chain of 2,4-D and MCPA to produce 2,4-dichlorophenol and 4-chloro-2-methylphenol, respectively, and glyoxylate while oxidizing α-ketoglutarate to CO2 and succinate (16, 17).Organisms capable of PAA herbicide degradation are phylogenetically diverse and belong to the Alpha-, Beta-, and Gammproteobacteria and the Bacteroidetes/Chlorobi group (e.g., references 2, 14, 29-34, 39, 60, 68, and 71). These bacteria harbor tfdA-like genes (i.e., tfdA or tfdAα) and are categorized into three groups on an evolutionary and physiological basis (34). The first group consists of beta- and gammaproteobacteria and can be further divided into three distinct classes based on their tfdA genes (30, 46). Class I tfdA genes are closely related to those of Cupriavidus necator JMP134 (formerly Ralstonia eutropha). Class II tfdA genes consist of those of Burkholderia sp. strain RASC and a few strains that are 76% identical to class I tfdA genes. Class III tfdA genes are 77% identical to class I and 80% identical to class II tfdA genes and linked to MCPA degradation in soil (3). The second group consists of alphaproteobacteria, which are closely related to Bradyrhizobium spp. with tfdAα genes having 60% identity to tfdA of group 1 (18, 29, 34). The third group also harbors the tfdAα genes and consists of Sphingomonas spp. within the alphaproteobacteria (30).Diverse PAA herbicide degraders of all three groups were identified in soil by cultivation-dependent studies (32, 34, 41, 78). Besides CadAB, TfdA and certain TfdAα proteins catalyze the conversion of PAA herbicides (29, 30, 35). All groups of tfdA-like genes are potentially linked to the degradation of PAA herbicides, although alternative primary functions of group 2 and 3 TfdAs have been proposed (30, 35). However, recent cultivation-independent studies focused on 16S rRNA genes or solely on group 1 tfdA sequences in soil (e.g., references 3-5, 13, and 41). Whether group 2 and 3 tfdA-like genes are also quantitatively linked to the degradation of PAA herbicides in soils is unknown. Thus, tools to target a broad range of tfdA-like genes are needed to resolve such an issue. Primers used to assess the diversity of tfdA-like sequences used in previous studies were based on the alignment of approximately 50% or less of available sequences to date (3, 20, 29, 32, 39, 47, 58, 73). Primers specifically targeting all major groups of tfdA-like genes to assess and quantify a broad diversity of potential PAA degraders in soil are unavailable. Thus, the objectives of this study were (i) to develop primers specific for all three groups of tfdA-like genes, (ii) to establish quantitative kinetic PCR (qPCR) assays based on such primers for different soil samples, and (iii) to assess the diversity and abundance of tfdA-like genes in soil.  相似文献   

2.
2-Methyl-4-chlorophenoxyacetic acid (MCPA) is a widely used phenoxyalkanoic acid herbicide and subject to aerobic microbial degradation. Earthworms stimulate both growth and activity of MCPA-degrading bacteria in soil. Thus, active MCPA degraders in soil and drilosphere (i.e. burrow walls, gut content and cast) were assessed by 16S rRNA stable isotope probing in soil columns under experimental conditions designed to minimize laboratory incubation biases. Agriculturally relevant concentrations of [(13) C]MCPA (20 μg g(dw) (-1)) were degraded in soil within 23 and 27 days in the presence and absence of earthworms respectively. Total 16S rRNA analysis revealed 73 operational taxonomic units indicative of active Acidobacteria, Actinobacteria, Bacteroidetes, Chloroflexi, Cyanobacteria, Firmicutes, Gemmatimonadetes, Planctomycetes, Proteobacteria and Verrucomicrobia in soil and drilosphere derived material. Seven operational taxonomic units indicative of Alpha-, Beta-, Gammaproteobacteria and Firmicutes consumed MCPA-[(13) C]. Dominant consumers of MCPA-[(13) C] were Alphaproteobacteria (Sphingomonadaceae and Bradyrhizobiaceae) in soil and drilosphere. Beta- (Comamonadaceae) and Gammaproteobacteria (Xanthomonadaceae) were also important MCPA-[(13) C] consumers in burrow walls only, indicating that earthworms favour betaproteobacterial MCPA degraders. In oxic microcosms with bulk soil, burrow walls and cast, 20 and 300-400 μg g(dw) (-1) [(13) C]MCPA were consumed within 24 h and 20 days respectively. Gut contents did not facilitate the degradation of [(13) C]MCPA. Sphingomonadaceae dominated MCPA-[(13) C] consumers in bulk soil and burrow wall microcosms, while Beta- and Gammaproteobacteria (Burkholderiacea, Comamonadaceae, Oxalobacteraceae and Xanthomonadaceae) dominated MCPA-[(13) C] consumers in microcosms of cast, indicating that the latter taxa are prone to respond to MCPA in cast. The collective data indicated that Alphaproteobacteria are major MCPA degraders in soil and drilosphere.  相似文献   

3.
In soil, the way biotic parameters impact the relationship between bacterial diversity and function is still unknown. To understand these interactions better, we used RNA-based stable-isotope probing to study the diversity of active atrazine-degrading bacteria in relation to atrazine degradation and to explore the impact of earthworm-soil engineering with respect to this relationship. Bulk soil, burrow linings and earthworm casts were incubated with 13C-atrazine. The pollutant degradation was quantified by liquid chromatography–mass spectrometry for 8 days, whereas active atrazine degraders were identified at 2 and 8 days by sequencing the 16S ribosomal RNA in the 13C-RNA fractions from the three soil microsites. An original diversity of atrazine degraders was found. Earthworm soil engineering greatly modified the taxonomic composition of atrazine degraders with dominance of α-, β- and γ-proteobacteria in burrow linings and of Actinobacteria in casts. Earthworm soil bioturbation increased the γ-diversity of atrazine degraders over the soil microsites generated. Atrazine degradation was enhanced in burrow linings in which primary atrazine degraders, closely related to Pelomonas aquatica, were detected only 2 days after atrazine addition. Atrazine degradation efficiency was not linearly related to the species richness of degraders but likely relied on keystone species. By enhancing soil heterogeneity, earthworms sustained high phylogenetic bacterial diversity and exerted a biotic control on the bacterial diversity–function relationships. Our findings call for future investigations to assess the ecological significance of biotic controls on the relationships between diversity and function on ecosystem properties and services (for example, soil detoxification) at larger scales.  相似文献   

4.
To follow the anaerobic degradation of organic matter in tidal-flat sediments, a stimulation experiment with 13C-labeled Spirulina biomass (130 mg per 21 g sediment slurry) was conducted over a period of 24 days. A combination of microcalorimetry to record process kinetics, chemical analyses of fermentation products and RNA-based stable-isotope probing (SIP) to follow community changes was applied. Different degradation phases could be identified by microcalorimetry: Within 2 days, heat output reached its maximum (55 μW), while primary fermentation products were formed (in μmol) as follows: acetate 440, ethanol 195, butyrate 128, propionate 112, H2 127 and smaller amounts of valerate, propanol and butanol. Sulfate was depleted within 7 days. Thereafter, methanogenesis was observed and secondary fermentation proceeded. H2 and alcohols disappeared completely, whereas fatty acids decreased in concentration. Three main degraders were identified by RNA-based SIP and denaturant gradient gel electrophoresis. After 12 h, two phylotypes clearly enriched in 13C: (i) Psychrilyobacter atlanticus, a fermenter known to produce hydrogen and acetate and (ii) bacteria distantly related to Propionigenium. A Cytophaga-related bacterium was highly abundant after day 3. Sulfate reduction appeared to be performed by incompletely oxidizing species, as only sulfate-reducing bacteria related to Desulfovibrio were labeled as long as sulfate was available.  相似文献   

5.

Background

Phylogenetic heterogeneity across Pseudomonas genus is complemented by its diverse genome architecture enriched by accessory genetic elements (plasmids, transposons, and integrons) conferring resistance across this genus. Here, we sequenced a stress tolerant genotype i.e. Pseudomonas sp. strain RL isolated from a hexachlorocyclohexane (HCH) contaminated pond (45 mg of total HCH g−1 sediment) and further compared its gene repertoire with 17 reference ecotypes belonging to P. stutzeri, P. mendocina, P. aeruginosa, P. psychrotolerans and P. denitrificans, representing metabolically diverse ecosystems (i.e. marine, clinical, and soil/sludge). Metagenomic data from HCH contaminated pond sediment and similar HCH contaminated sites were further used to analyze the pan-genome dynamics of Pseudomonas genotypes enriched across increasing HCH gradient.

Results

Although strain RL demonstrated clear species demarcation (ANI ≤ 80.03%) from the rest of its phylogenetic relatives, it was found to be closest to P. stutzeri clade which was further complemented functionally. Comparative functional analysis elucidated strain specific enrichment of metabolic pathways like α-linoleic acid degradation and carbazole degradation in Pseudomonas sp. strain RL and P. stutzeri XLDN-R, respectively. Composition based methods (%codon bias and %G + C difference) further highlighted the significance of horizontal gene transfer (HGT) in evolution of nitrogen metabolism, two-component system (TCS) and methionine metabolism across the Pseudomonas genomes used in this study. An intact mobile class-I integron (3,552 bp) with a captured gene cassette encoding for dihydrofolate reductase (dhfra1) was detected in strain RL, distinctly demarcated from other integron harboring species (i.e. P. aeruginosa, P. stutzeri, and P. putida). Mobility of this integron was confirmed by its association with Tnp21-like transposon (95% identity) suggesting stress specific mobilization across HCH contaminated sites. Metagenomics data from pond sediment and recently surveyed HCH adulterated soils revealed the in situ enrichment of integron associated transposase gene (TnpA6100) across increasing HCH contamination (0.7 to 450 mg HCH g−1 of soil).

Conclusions

Unlocking the potential of comparative genomics supplemented with metagenomics, we have attempted to resolve the environment and strain specific demarcations across 18 Pseudomonas gene complements. Pan-genome analyses of these strains indicate at astoundingly diverse metabolic strategies and provide genetic basis for the cosmopolitan existence of this taxon.

Electronic supplementary material

The online version of this article (doi:10.1186/s12864-015-1488-2) contains supplementary material, which is available to authorized users.  相似文献   

6.
7.
Enhanced biodegradation in the rhizosphere has been reported for many organic xenobiotic compounds, although the mechanisms are not fully understood. The purpose of this study was to discover whether rhizosphere-enhanced biodegradation is due to selective enrichment of degraders through growth on compounds produced by rhizodeposition. We monitored the mineralization of [U-14C]2,4-dichlorophenoxyacetic acid (2,4-D) in rhizosphere soil with no history of herbicide application collected over a period of 0 to 116 days after sowing of Lolium perenne and Trifolium pratense. The relationships between the mineralization kinetics, the number of 2,4-D degraders, and the diversity of genes encoding 2,4-D/α-ketoglutarate dioxygenase (tfdA) were investigated. The rhizosphere effect on [14C]2,4-D mineralization (50 μg g−1) was shown to be plant species and plant age specific. In comparison with nonplanted soil, there were significant (P < 0.05) reductions in the lag phase and enhancements of the maximum mineralization rate for 25- and 60-day T. pratense soil but not for 116-day T. pratense rhizosphere soil or for L. perenne rhizosphere soil of any age. Numbers of 2,4-D degraders in planted and nonplanted soil were low (most probable number, <100 g−1) and were not related to plant species or age. Single-strand conformational polymorphism analysis showed that plant species had no impact on the diversity of α-Proteobacteria tfdA-like genes, although an impact of 2,4-D application was recorded. Our results indicate that enhanced mineralization in T. pratense rhizosphere soil is not due to enrichment of 2,4-D-degrading microorganisms by rhizodeposits. We suggest an alternative mechanism in which one or more components of the rhizodeposits induce the 2,4-D pathway.  相似文献   

8.
The impact of temperature and cycle length on microbial competition between polyhydroxybutyrate (PHB)-producing populations enriched in feast-famine sequencing batch reactors (SBRs) was investigated at temperatures of 20 °C and 30 °C, and in a cycle length range of 1–18 h. In this study, the microbial community structure of the PHB-producing enrichments was found to be strongly dependent on temperature, but not on cycle length. Zoogloea and Plasticicumulans acidivorans dominated the SBRs operated at 20 °C and 30 °C, respectively. Both enrichments accumulated PHB more than 75% of cell dry weight. Short-term temperature change experiments revealed that P. acidivorans was more temperature sensitive as compared with Zoogloea. This is particularly true for the PHB degradation, resulting in incomplete PHB degradation in P. acidivorans at 20 °C. Incomplete PHB degradation limited biomass growth and allowed Zoogloea to outcompete P. acidivorans. The PHB content at the end of the feast phase correlated well with the cycle length at a constant solid retention time (SRT). These results suggest that to establish enrichment with the capacity to store a high fraction of PHB, the number of cycles per SRT should be minimized independent of the temperature.  相似文献   

9.
Separate quantification of three classes of tfdA genes was performed using TaqMan quantitative real-time PCR for 13 different soils subsequent to mineralization of three phenoxy acids. Class III tfdA genes were found to be involved in mineralization more often than class I and II tfdA genes.The phenoxy acids (PA), including 2,4-dichlorophenoxyacetic acid (2,4-D), 4-chloro-2-methylphenoxyacetic acid (MCPA), and 2-(4-chloro-2-methylphenoxy)propanoic acid (MCPP), are herbicides that are intensively used for control of broadleaf weeds in cereal crops worldwide. Extensive research on these environmentally hazardous compounds has produced detailed information on pathways and gene sequences involved in their complete mineralization in several bacterial pure-culture isolates (3, 12, 15, 16).Decomposition of the ether bond resulting in a phenolic compound and acetic or propanoic acid is accepted to be the first step in the degradation pathway (5). This step is catalyzed by an α-ketoglutarate-dependent dioxygenase encoded by the tfdA or tfdA-like genes (6). The diversity of the tfdA genes has been investigated in detail, and three different classes have been proposed based on sequence information (10). Recently, the tfdA genes have been used as biomarkers in studies of the growth of degraders during mineralization of PA in natural soil samples in situ (1, 2, 7, 8, 19), but none of these studies successfully differentiated quantities of the individual tfdA gene classes.Previously, two different PCR assays were used to detect and quantify tfdA genes in environmental samples. Vallaeys et al. (17) developed a PCR assay suitable for targeting and proving the presence of the three classes of tfdA genes, while, in order to improve specificity and PCR efficiency, Bælum et al. (1) developed and used a novel PCR primer set more suitable for quantitative real-time PCR (qPCR) (1, 2, 11). Furthermore, based on endpoint analysis of PCR products, these two PCR assays have been used to study the functional diversity and dynamics of the different classes of tfdA genes during mineralization of phenoxyacetic acids in environmental samples. As this kind of analysis does not provide a quantitative measure of the relative composition of the different classes, the need for a tool to do this is evident. Therefore, the main objective of the present study was to develop a reliable quantitative real-time PCR-based method to separately quantify the different classes of tfdA genes directly in complex soil samples.Two novel qPCR assays (tfdA 81-bp and TaqMan assays) and one previously described qPCR assay (tfdA 215-bp assay) (1) targeting the known diversity of tfdA genes were tested to determine their PCR efficiency and specificity for the different classes of tfdA genes. Due to potential problems with production of PCR artifacts, the 215-bp PCR fragment is in the longer range of the fragment size recommended for qPCR (18), explaining why we designed a novel primer set amplifying a shorter fragment. The tfdA 81-bp and tfdA 215-bp assays were based on the SYBR green dye, while the TaqMan assay was based on TaqMan probes. Also, in the TaqMan assay our novel 81-bp primer set was used to amplify the target sequences. Primers for the novel tfdA 81-bp assay (Table (Table1;1; see Fig. S1 in the supplemental material), as well as probes for the TaqMan assay, were designed based on tfdA gene sequences obtained from the GenBank database (accession numbers M16730, U25717, and AF377325 for tfdA classes I to III, respectively). To confirm that no unwanted targets are amplified with our primer sets, the oligonucleotides were used as queries for a BLAST search in the GenBank database (for details see the supplemental Materials and Methods in the supplemental material). In the present study we had access to instrument facilities that detect only two fluorophores simultaneously. Therefore, we were not able to test multiplex PCR, where all three tfdA gene classes can be quantified in a single PCR vessel, but we believe that this should be possible by using probes with three different fluorophores.

TABLE 1.

Primers and probes used for qPCR
OligonucleotideTypeTarget genesSequence (5′-3′)aFragment size (bp)Annealing temp (°C)
tfdA-215bpForward primertfdA classes I, II, and IIIGAGCACTACGCRCTGAAYTCCCG21564
Reverse primerGTCGCGTGCTCGAGAAG
tfdA-81bpForward primertfdA classes I, II, and IIIGAGCACTACGCRCTGAAYTCCCG8162
Reverse primerSACCGGMGGCATSGCATT
tfdA-CIProbetfdA class IFAM-TTGCGCTTCCGAATAGTCGGTGTC-BBQ62
tfdA-CIIProbetfdA class IIFAM-CGTTGACTTTCAGAATACTCT GTGTCGCCA-BBQ62
tfdA-CIIIProbetfdA class IIIYAK-TTGACTTTCAGAATAGTCCGTATCGCCAAG-BBQ62
Open in a separate windowaR = A or G; Y = T or C; S = G or C; M = A or C. FAM, 6-carboxyfluorescein; BBQ, blackberry quencher.Using 10-fold standard dilution series of recombined plasmids with inserts of each of the class I, II, and III tfdA genes and 107 to 100 genes per reaction as a template, highly comparable PCR performances were achieved with the three different assays (see Fig. S2 and Table S2 in the supplemental material). Except for the tfdA 215-bp assay targeting the class II gene, consistent and reliable quantification was obtained down to a limit of 102 genes per reaction for all three tfdA classes (see Fig. S2 in the supplemental material). The comparable qPCR sensitivities for the SYBR green and TaqMan probe qPCR assays are in accordance the sensitivities reported elsewhere (4, 9, 13). Detection (i.e., the replicates were not consistent and thus unreliable quantification was obtained) could be performed with as few as 101 genes per reaction. Probably due to a 1-bp mismatch in the reverse primer region, reliable quantification of class II genes using the tfdA 215-bp assay could be obtained with only as few as 104 genes per reaction (see Fig. S2a in the supplemental material). Furthermore, a test with different combinations of the tfdA target sequences and with the three TaqMan probes together in one reaction vessel was performed in order to verify the specificity of each of the probes for its target sequence. Also, we tested the feasibility of using the probes to perform duplex real-time PCR with combinations of the class I- and III-specific probes and of the class II- and III-specific probes. These tests gave the same results for the detection level and specificity as the tests described above, indicating that duplex PCR is indeed possible with these probes (for detailed information, see the supplemental Materials and Methods in the supplemental material).A soil microcosm experiment including 13 different soils obtained from distinct locations around the world (for further details on soils, see Table S1 in the supplemental material) was performed to test the different qPCR approaches with DNA extracted directly from a wide variety of soils. In order to allow specific degraders harboring the tfdA genes to proliferate, we treated the microcosms with 90 μmol of 2,4-D, MCPA, or MCPP kg−1 of soil. Prior to and subsequent to ≥50% mineralization (measured using evolved 14CO2), a composite subsample consisting of 0.5 g of soil was removed, and DNA was extracted using a Power-Clean soil DNA kit (MoBio Laboratories, Carlsbad, CA) according to the manufacturer''s instructions. Total DNA was quantified and the extraction efficiency was normalized by running 4-μl aliquots of extracts on a standard 1.5% agarose gel stained with ethidium bromide, and qPCR was performed using the three different approaches (for further information on the experimental setup, nucleic acid preparation, and qPCR, see the supplemental Materials and Methods in the supplemental material).Despite the low detection limit of the qPCR approaches, we were able to detect low levels of tfdA genes (<105 genes g−1 soil) in only three soils, SjOreg, Suma-Paz, and KBSreg (for information on soils, see the supplemental Materials and Methods in the supplemental material) prior to PA application, while after mineralization of one of the PA (>50% mineralized) we were able to detect significant increases in the levels of tfdA genes in the soils. The potential to mineralize the three PA was investigated using the widely used and well-validated assay for trapping 14CO2 in an NaOH trap during mineralization in microcosms (14). The potential to mineralize >50% of the added PA was found for all 13 soils for 2,4-D, for 9 soils for MCPA, and for 4 soils for MCPP, indicating that the potential for 2,4-D mineralization is more widespread than the potential for MCPA and MCPP mineralization. Additionally, our data suggest that 2,4-D is mineralized more rapidly than MCPA and MCPP (data not shown). In all of the soils exhibiting >50% mineralization we detected increased levels of tfdA genes (Fig. (Fig.1),1), while in the soils with very slow and insignificant mineralization no such increase was detected.Open in a separate windowFIG. 1.Ratios of the three tfdA gene classes quantified using the TaqMan assay, expressed as the cumulative numbers of the three classes of tfdA genes in each of the 13 soils after mineralization of 2,4-D, MCPA, and MCPP. (A to C) Class I tfdA genes quantified in soils exposed to the three PA. (D to F) Class II tfdA genes quantified in soils exposed to the three PA. (G to I) Class III tfdA genes quantified in soils exposed to the three PA. The individual values for triplicate samples are indicated, and for clarity, the soils for which data are shown in each panel are indicated at the bottom. The exact numbers are shown in Table S3 in the supplemental material.Even though there was a slight tendency for the tfdA 81-bp SYBR green and TaqMan assays to reveal larger quantities (Fig. (Fig.2A2A and and2C),2C), the correlation between the tfdA quantities obtained using the three different qPCR approaches was very high (Fig. (Fig.2).2). Based on two-way analysis of variance statistics (P < 0.05), the only soil scenarios for which a difference between the quantification results could be detected were 2,4-D and MCPA in the Pradera soil and 2,4-D in the Suma-Paz soil and the KBSforest soil. For the Pradera soil treated with 2,4-D the difference was especially noticeable (the outliers are shown in Fig. Fig.2A2A and and2B).2B). For this soil scenario we were able to find 6 × 108 tfdA genes g−1 soil using the tfdA 81-bp SYBR green assay, while with the two other assays we were able to find only >5 × 104 tfdA genes g−1 soil. This suggests that a degrader harboring a novel class of tfdA genes was present, but due to difficulties in sequencing extremely short PCR amplicons we have not been able to obtain sequence information. We consider the generally high correlation between tfdA quantities obtained using the three qPCR assays strong evidence that these three assays are suitable for consistent and reliable quantification of the class I to III tfdA genes in environmental samples.Open in a separate windowFIG. 2.Correlations between the numbers of class I to III tfdA genes determined using the three different approaches. (A) tfdA 81-bp assay compared with the tfdA 215-bp assay. (B) TaqMan assay compared with the tfdA 81-bp assay. (C) TaqMan assay compared with the tfdA 215-bp assay. The lines represent linear regressions, and the R2 value is indicated in each panel. The error bars indicate standard errors of triplicate determinations.The usability of our novel TaqMan probe-based qPCR approach was demonstrated by quantifying the three different tfdA gene classes after mineralization of 2,4-D, MCPA, and MCPP in 13 different soils (Fig. (Fig.1;1; see Table S3 in the supplemental material). For most of the soils, we observed that if class III tfdA genes were present in a soil, they typically proliferated and became dominant among the tfdA genes after mineralization occurred. In the soils treated with 2,4-D this was reflected by generally higher ratios of class III tfdA genes to class I tfdA genes (Fig. (Fig.1A1A and and1G).1G). This trend was even more obvious for MCPA and MCPP mineralization, where class I tfdA genes proliferated only in the KBSreg and KBSforest soils treated with MCPA (Fig. (Fig.1B),1B), while in the remaining soils showing potential for MCPA and MCPP mineralization only the class III tfdA genes proliferated (Fig. (Fig.1B1B and and1I).1I). Further proof was the fact that the KBSreg and KBSforest soils were the only soils treated with 2,4-D in which class III tfdA genes did not proliferate, suggesting that there was no potential for proliferation of class III tfdA genes, which left the PA for the organisms harboring the class I tfdA genes.In previous studies we speculated that MCPA mineralization is linked to the class III tfdA gene (1) and 2,4-D mineralization is linked to both class I and class III tfdA genes (1, 2). In these studies only two different soils, originating from the same region, were studied. Furthermore, based on enrichment studies of microbial communities, Zakaria et al. (19) linked mineralization of MCPP to the class III tfdA gene. The present study adds significantly to our knowledge of the dynamics of tfdA genes during mineralization of the PA 2,4-D, MCPA, and MCPP. Here we describe a method to quantitatively measure the relative levels of the different tfdA gene classes in a wide variety of soils, which provides significantly better data.In conclusion, we successfully developed a TaqMan-based qPCR method to quantify three different classes of tfdA genes in environmental samples. Applying the method, we were able to quantify tfdA genes in 13 different soils subsequent to mineralization of 2,4-D, MCPA, and MCPP. In situ in natural soils, mineralization of 2,4-D can induce growth of organisms harboring one of the three tfdA genes, preferentially the class I and III tfdA genes, and the class III tfdA gene is most often the dominant gene. Mineralization of MCPA induces proliferation of class III tfdA genes, and in cases where the potential for class III gene proliferation is absent, class I genes may proliferate as well. Mineralization of MCPP induces proliferation of only class III tfdA genes.  相似文献   

10.
蚯蚓在植物修复芘污染土壤中的作用   总被引:1,自引:1,他引:0  
潘声旺  魏世强  袁馨  曹生宪 《生态学报》2011,31(5):1349-1355
采用盆栽试验法,研究了蚯蚓(Pheretima hupeiensis)在植物修复芘污染土壤中的作用。结果显示,试验浓度(20.24-321.42 mg/kg) 范围内,蚯蚓活动促进了芘污染土壤中修复植物黑麦草(Lolium multiforum)黑麦草的生长,其根冠比明显增大。添加蚯蚓72 d后,种植黑麦草的土壤中芘的去除率高达60.01%-86.26%,其平均去除率(74.66%)比无蚯蚓活动的土壤-植物系统(64.55%)提高10.11%,比无植物对照组(18.24%)提高56.42%。各种生物、非生物修复因子中,植物-微生物交互作用对芘去除的平均贡献率(51.75%)最为突出,比无蚯蚓活动时(44.94%)提高6.81%。说明蚯蚓活动可强化土壤-植物系统对土壤芘污染的修复作用。  相似文献   

11.
Understanding the microbial degradation of Microcystis biomass is crucial for determining the ecological consequences of Microcystis blooms in freshwater lakes. The purpose of this study was to identify bacteria involved in the anaerobic degradation of Microcystis blooms. Microcystis scum was anaerobically incubated for 90 days at three temperatures (15 °C, 25 °C and 35 °C). We used terminal restriction fragment length polymorphism (T-RFLP) analysis of bacterial 16S rRNA genes, followed by cloning and sequencing of selected samples, to reveal the community composition of bacteria and their dynamics during decomposition. Clostridium spp. were found to be the most dominant bacteria in the incubations, accounting for 72% of the sequenced clones. Eight new clusters or subclusters (designated CLOS.1–8) were identified in the Clostridium phylogenetic tree. The bacterial populations displayed distinct successions during Microcystis decomposition. Temperature had a strong effect on the dynamics of the bacterial populations. At 15 °C, the initial dominance of a 207-bp T-RF (Betaproteobacteria) was largely substituted by a 227-bp T-RF (Clostridium, new cluster CLOS.2) at 30 days. In contrast, at 25 °C and 35 °C, we observed an alternating succession of the 227-bp T-RF and a 231-bp T-RF (Clostridium, new cluster CLOS.1) that occurred more than four times; no one species dominated the flora for the entire experiment. Our study shows that novel Clostridium clusters and their diverse consortiums dominate the bacterial communities during anaerobic degradation of Microcystis, suggesting that these microbes'' function in the degradation process.  相似文献   

12.

Background

Essential oils extracted from aromatic and medicinal plants have many biological properties and are therefore an alternative to the use of synthetic products. The chemical composition of essential oils from two medicinal plants (Eucalyptus globulus and E. lehmannii) was determined and, their insecticidal effects on the third and fourth larval stages of Orgyia trigotephras were assessed.

Results

Larvae were collected from Jebel Abderrahmane (North-East of Tunisia), conserved in groups of 50/box (21 × 10 × 10 cm) at a temperature of 25°C. Larvae were tested for larvicidal activities of essential oils. Each oil was diluted in ethanol (96%) to prepare 3 test solutions (S1 = 0.05%, S2 = 0.10% and S3 = 0.50%). Essential oils were used for contact, ingestion and Olfactory actions and compared to reference products (Bacillus thuringiensis and Decis). Olfactory action of essential oils shows that larvae mortality is higher than contact action, lower than ingestion action. MTM and FTM of S3 of E. lehmannii were respectively 1 h 32 min and 1 h 39 min are higher than those of E. globulus (MTM = 51 min and FTM = 1 h 22 min 34 sec). Contact action of E. lehmannii oil shows low insecticidal activity compared to E. globulus. MTM are respectively (1 min 52 sec and 1 min 7 sec), FTM are (2 min 38 sec, 1 min 39 sec), are the shortest recorded for S3, on the third stage of larvae. The fourth stage of larvae, MTM are (2 min 20 sec and 2 min 9 sec), FTM are (3 min 25 sec, 3 min 19 sec). Ingestion action of essential oils is longer than the contact action, since the time of death exceeds 60 minutes for all species.

Conclusion

Results shows that essential oils have a toxic action on nerves leading to a disruption of vital system of insects. High toxic properties make these plant-derived compounds suitable for incorporation in integrated pest management programs.  相似文献   

13.
Influences of infaunal burrows constructed by the polychaete (Tylorrhynchus heterochaetus) on O2 concentrations and community structures and abundances of ammonia-oxidizing bacteria (AOB) and nitrite-oxidizing bacteria (NOB) in intertidal sediments were analyzed by the combined use of a 16S rRNA gene-based molecular approach and microelectrodes. The microelectrode measurements performed in an experimental system developed in an aquarium showed direct evidence of O2 transport down to a depth of 350 mm of the sediment through a burrow. The 16S rRNA gene-cloning analysis revealed that the betaproteobacterial AOB communities in the sediment surface and the burrow walls were dominated by Nitrosomonas sp. strain Nm143-like sequences, and most of the clones in Nitrospira-like NOB clone libraries of the sediment surface and the burrow walls were related to the Nitrospira marina lineage. Furthermore, we investigated vertical distributions of AOB and NOB in the infaunal burrow walls and the bulk sediments by real-time quantitative PCR (Q-PCR) assay. The AOB and Nitrospira-like NOB-specific 16S rRNA gene copy numbers in the burrow walls were comparable with those in the sediment surfaces. These numbers in the burrow wall at a depth of 50 to 55 mm from the surface were, however, higher than those in the bulk sediment at the same depth. The microelectrode measurements showed higher NH4+ consumption activity at the burrow wall than those at the surrounding sediment. This result was consistent with the results of microcosm experiments showing that the consumption rates of NH4+ and total inorganic nitrogen increased with increasing infaunal density in the sediment. These results clearly demonstrated that the infaunal burrows stimulated O2 transport into the sediment in which otherwise reducing conditions prevailed, resulting in development of high NH4+ consumption capacity. Consequently, the infaunal burrow became an important site for NH4+ consumption in the intertidal sediment.  相似文献   

14.
The competition for glucose between Escherichia coli ML30, a typical copiotrophic enterobacterium and Chelatobacter heintzii ATCC29600, an environmentally successful strain, was studied in a carbon-limited culture at low dilution rates. First, as a base for modelling, the kinetic parameters μmax and Ks were determined for growth with glucose. For both strains, μmax was determined in batch culture after different precultivation conditions. In the case of C. heintzii, μmax was virtually independent of precultivation conditions. When inoculated into a glucose-excess batch culture medium from a glucose-limited chemostat run at a dilution rate of 0.075 h−1 C. heintzii grew immediately with a μmax of 0.17±0.03 h−1. After five transfers in batch culture, μmax had increased only slightly to 0.18±0.03 h−1. A different pattern was observed in the case of E. coli. Inoculated from a glucose-limited chemostat at D=0.075 h−1 into glucose-excess batch medium E. coli grew only after an acceleration phase of ∼3.5 h with a μmax of 0.52 h−1. After 120 generations and several transfers into fresh medium, μmax had increased to 0.80±0.03 h−1. For long-term adapted chemostat-cultivated cells, a Ks for glucose of 15 μg l−1 for C. heintzii, and of 35 μg l−1 for E. coli, respectively, was determined in 14C-labelled glucose uptake experiments. In competition experiments, the population dynamics of the mixed culture was determined using specific surface antibodies against C. heintzii and a specific 16S rRNA probe for E. coli. C. heintzii outcompeted E. coli in glucose-limited continuous culture at the low dilution rates of 0.05 and 0.075 h−1. Using the determined pure culture parameter values for Ks and μmax, it was only possible to simulate the population dynamics during competition with an extended form of the Monod model, which includes a finite substrate concentration at zero growth rate (smin). The values estimated for smin were dependent on growth rate; at D=0.05 h−1, it was 12.6 and 0 μg l−1 for E. coli and C. heintzii, respectively. To fit the data at D=0.075 h−1, smin for E. coli had to be raised to 34.9 μg l−1 whereas smin for C. heintzii remained zero. The results of the mathematical simulation suggest that it is not so much the higher Ks value, which is responsible for the unsuccessful competition of E. coli at low residual glucose concentration, but rather the existence of a significant smin.  相似文献   

15.

Background

Several studies on the association of TNF-alpha (−308 G/A), IL-6 (−174 G/C) and IL-1beta (−511 C/T) polymorphisms with polycystic ovary syndrome (PCOS) risk have reported conflicting results. The aim of the present study was to assess these associations by meta-analysis.

Results

A total of 14 eligible articles (1665 cases/1687 controls) were included in this meta-analysis. The results suggested that there was no obvious association between the TNF-alpha (−308 G/A) polymorphism and PCOS in the overall population or subgroup analysis by ethnicity, Hardy–Weinberg equilibrium (HWE) in controls, genotyping method, PCOS diagnosis criteria, and study sample size. Also, no obvious association was found between the TNF-alpha (−308 G/A) polymorphism and obesity in patients with PCOS (body mass index [BMI] ≥ 25 kg/m2 vs. BMI < 25 kg/m2). Regarding the IL-6 (−174 G/C) polymorphism, also no association was found in the overall population in heterozygote comparison, dominant model, and recessive model. Even though an allelic model (odds ratio [OR] = 0.63, 95% confidence interval [CI] = 0.41–0.96) and a homozygote comparison (OR = 0.52, 95% CI = 0.30–0.93) showed that the IL-6 (−174 G/C) polymorphism was marginally associated with PCOS. Further subgroup analysis suggested that the effect size was not significant among HWE in controls (sample size ≤ 200) and genotyping method of pyrosequencing under all genetic models. Similarly, there was no association between the IL-1beta (−511 C/T) polymorphism and PCOS in the overall population or subgroup analysis under all genetic models. Furthermore, no significant association was found between the IL-1beta (−511 C/T) polymorphism and several clinical and biochemical parameters in patients with PCOS.

Conclusions

The results of this meta-analysis suggest that the TNF-alpha (−308 G/A), IL-6 (−174 G/C), and IL-1beta (−511 C/T) polymorphisms may not be associated with PCOS risk. However, further case–control studies with larger sample sizes are needed to confirm our results.

Electronic supplementary material

The online version of this article (doi:10.1186/s12863-015-0165-4) contains supplementary material, which is available to authorized users.  相似文献   

16.

Background

The current study aims at evaluating the analgesic, anti-pyretic and anti-inflammatory properties of methanolic extract of the stem, bark and leaves of Launaea sarmentosa and Aegialitis rotundifolia roxb.

Results

The AELS and AEAR extract presented a significant (***p < 0.001) dose dependent increase in reaction time in writhing method and showed inhibition of 63.1% and 57.1% respectively at the doses of 400 mg/kg body weight while standard drug showed (P < 0.001) inhibition of 69.23%. In tail immersion method, AELS and AEAR showed maximum time of tail retention at 30 min in hot water i.e. 6.93 sec and 6.54 sec respectively at highest doses of 400 mg/kg body weight than lower dose while standard pentazocine showed reaction time of 7.62 sec. The AELS and AEAR extract also exhibited promising anti-inflammatory effect as demonstrated by statistically significant inhibition of paw volume by 32.48% and 26.75% respectively at the dose of 400 mg/kg body weight while the value at the dose of 200 mg/kg body weight were linear to higher dose at the 3rd hour of study. On the other hand, Standard indomethacin inhibited 40.13% of inflammation (***P < 0.001). In Cotton-pellet granuloma method, AELS and AEAR extract at the dose of 400 mg/kg body weight exhibited inhibition of inflammation of 34.7% and 29.1% respectively while standard drug showed (P < 0.001) inhibition of 63.22%. Intraperitoneal administration of AELS and AEAR showed dose dependent decrease in body temperature in brewer’s yeast induced hyperthermia in rats at both doses. However, AELS significantly decreased body temperature (***p < 0.001) at 400 mg/kg compared to control.

Conclusions

Present work propose that the methanolic extract of Launaea sarmentosa and Aegialitis rotundifolia roxb possesses dose dependent pharmacological action which supports its therapeutic use in folk medicine possibly mediated through the inhibition or blocking of release of prostaglandin and/or actions of vasoactive substances such as histamine, serotonin and kinins.  相似文献   

17.
Zinc (Zn) is an important micronutrient for the physiology of plants. It is poorly available to the plants in soil solution. A pot experiment was conducted to evaluate effectiveness of various Zn application methods on key enzyme activities and protein content of two contrasting rice genotypes viz., PD16 (Zn efficient) and NDR359 (Zn inefficient). The treatments were, control (0 mg Zn kg−1 soil), soil application (5 mg Zn kg−1 soil), foliar application (0.5 % ZnSO4 + 0.25 % lime at 30, 60 and 90 days after transplanting), soil (5 mg Zn kg−1 soil) + foliar application of 0.5 % ZnSO4 + 0.25 % lime at 30, 60 and 90 days after transplanting. Among all the methods tested soil+foliar application of Zn fertilizers was found most effective in increasing superoxide dismutase (SOD) and carbonic anhydrase (CA) activities as well as chlorophyll and protein content in both the rice varieties. NDR359, showed higher enzyme activities and more chlorophyll content in leaves than PD16, when Zn was applied either through foliar spray alone or in soil along with foliar application. Regarding the protein content in grains, PD16 showed higher protein content than NDR359, thus showed better translocation of Zn from leaves to grains.  相似文献   

18.
19.
Mutations beneficial in one environment may cause costs in different environments, resulting in antagonistic pleiotropy. Here, we describe a novel form of antagonistic pleiotropy that operates even within the same environment, where benefits and deleterious effects exhibit themselves at different growth rates. The fitness of hfq mutations in Escherichia coli affecting the RNA chaperone involved in small-RNA regulation is remarkably sensitive to growth rate. E. coli populations evolving in chemostats under nutrient limitation acquired beneficial mutations in hfq during slow growth (0.1 h−1) but not in populations growing sixfold faster. Four identified hfq alleles from parallel populations were beneficial at 0.1 h−1 and deleterious at 0.6 h−1. The hfq mutations were beneficial, deleterious or neutral at an intermediate growth rate (0.5 h−1) and one changed from beneficial to deleterious within a 36 min difference in doubling time. The benefit of hfq mutations was due to the greater transport of limiting nutrient, which diminished at higher growth rates. The deleterious effects of hfq mutations at 0.6 h−1 were less clear, with decreased viability a contributing factor. The results demonstrate distinct pleiotropy characteristics in the alleles of the same gene, probably because the altered residues in Hfq affected the regulation of expression of different genes in distinct ways. In addition, these results point to a source of variation in experimental measurement of the selective advantage of a mutation; estimates of fitness need to consider variation in growth rate impacting on the magnitude of the benefit of mutations and on their fitness distributions.  相似文献   

20.

Background

Restricted space and close contact with conspecifics in captivity may be stressful for musk deer, as they are highly territorial and solitary in the wild. So we tested the effects of crowding on stress of forest musk deer (Moschus berezovskii) in heterosexual groups, using fecal cortisol analysis as a non-invasive method. 32 healthy adults during non-breeding seasons were chose as our experimental objects. Group 1 was defined as higher crowding condition, with 10-15 m2/deer (6 enclosures, 10♀ and 6♂); group 2 was defined as lower crowding condition, with 23-33 m2/deer (6 enclosures, 10♀ and 6♂). Every enclosure contained 1 male and 3 female. These patterns had been existed for years.

Results

The results showed that females in lower crowding condition (217.1 ± 9.5 ug/g) had significantly higher fecal cortisol levels than those in higher crowding condition (177.2 ± 12.1 ug/g). Interestingly, crowding seemed have no effect on male fecal cortisol levels (148.1 ± 9.1 ug/g and 140.5 ± 13.3 ug/g, respectively). At both groups, cortisol was significantly lower in males than in females.

Conclusions

These results showed that chronic crowding may affect stress status of captive forest musk deer. The captive environment should consider the space need for musk deer.  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