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.
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-(14)C]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/alpha-ketoglutarate dioxygenase (tfdA) were investigated. The rhizosphere effect on [(14)C]2,4-D mineralization (50 microg 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 alpha-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   

2.
Although metals are thought to inhibit the ability of microorganisms to degrade organic pollutants, several microbial mechanisms of resistance to metal are known to exist. This study examined the potential of cadmium-resistant microorganisms to reduce soluble cadmium levels to enhance degradation of 2,4-dichlorophenoxyacetic acid (2,4-D) under conditions of cocontamination. Four cadmium-resistant soil microorganisms were examined in this study. Resistant up to a cadmium concentration of 275 μg ml−1, these isolates represented the common soil genera Arthrobacter, Bacillus, and Pseudomonas. Isolates Pseudomonas sp. strain H1 and Bacillus sp. strain H9 had a plasmid-dependent intracellular mechanism of cadmium detoxification, reducing soluble cadmium levels by 36%. Isolates Arthrobacter strain D9 and Pseudomonas strain I1a both produced an extracellular polymer layer that bound and reduced soluble cadmium levels by 22 and 11%, respectively. Although none of the cadmium-resistant isolates could degrade 2,4-D, results of dual-bioaugmentation studies conducted with both pure culture and laboratory soil microcosms showed that each of four cadmium-resistant isolates supported the degradation of 500-μg ml−1 2,4-D by the cadmium-sensitive 2,4-D degrader Ralstonia eutropha JMP134. Degradation occurred in the presence of up to 24 μg of cadmium ml−1 in pure culture and up to 60 μg of cadmium g−1 in amended soil microcosms. In a pilot field study conducted with 5-gallon soil bioreactors, the dual-bioaugmentation strategy was again evaluated. Here, the cadmium-resistant isolate Pseudomonas strain H1 enhanced degradation of 2,4-D in reactors inoculated with R. eutropha JMP134 in the presence of 60 μg of cadmium g−1. Overall, dual bioaugmentation appears to be a viable approach in the remediation of cocontaminated soils.  相似文献   

3.
We examined the ability of a soil bacterium, Agrobacterium radiobacter J14a, to degrade the herbicide atrazine under a variety of cultural conditions, and we used this bacterium to increase the biodegradation of atrazine in soils from agricultural chemical distribution sites. J14a cells grown in nitrogen-free medium with citrate and sucrose as carbon sources mineralized 94% of 50 μg of [14C-U-ring]atrazine ml−1 in 72 h with a concurrent increase in the population size from 7.9 × 105 to 5.0 × 107 cells ml−1. Under these conditions cells mineralized the [ethyl-14C]atrazine and incorporated approximately 30% of the 14C into the J14a biomass. Cells grown in medium without additional carbon and nitrogen sources degraded atrazine, but the cell numbers did not increase. Metabolites produced by J14a during atrazine degradation include hydroxyatrazine, deethylatrazine, and deethyl-hydroxyatrazine. The addition of 105 J14a cells g−1 into soil with a low indigenous population of atrazine degraders treated with 50 and 200 μg of atrazine g−1 soil resulted in two to five times higher mineralization than in the noninoculated soil. Sucrose addition did not result in significantly faster mineralization rates or shorten degradation lag times. However, J14a introduction (105 cells g−1) into another soil with a larger indigenous atrazine-mineralizing population reduced the atrazine degradation lag times below those in noninoculated treatments but did not generally increase total atrazine mineralization.  相似文献   

4.
Prior to gene transfer experiments performed with nonsterile soil, plasmid pJP4 was introduced into a donor microorganism, Escherichia coli ATCC 15224, by plate mating with Ralstonia eutropha JMP134. Genes on this plasmid encode mercury resistance and partial 2,4-dichlorophenoxyacetic acid (2,4-D) degradation. The E. coli donor lacks the chromosomal genes necessary for mineralization of 2,4-D, and this fact allows presumptive transconjugants obtained in gene transfer studies to be selected by plating on media containing 2,4-D as the carbon source. Use of this donor counterselection approach enabled detection of plasmid pJP4 transfer to indigenous populations in soils and under conditions where it had previously not been detected. In Madera Canyon soil, the sizes of the populations of presumptive indigenous transconjugants were 107 and 108 transconjugants g of dry soil−1 for samples supplemented with 500 and 1,000 μg of 2,4-D g of dry soil−1, respectively. Enterobacterial repetitive intergenic consensus PCR analysis of transconjugants resulted in diverse molecular fingerprints. Biolog analysis showed that all of the transconjugants were members of the genus Burkholderia or the genus Pseudomonas. No mercury-resistant, 2,4-D-degrading microorganisms containing large plasmids or the tfdB gene were found in 2,4-D-amended uninoculated control microcosms. Thus, all of the 2,4-D-degrading isolates that contained a plasmid whose size was similar to the size of pJP4, contained the tfdB gene, and exhibited mercury resistance were considered transconjugants. In addition, slightly enhanced rates of 2,4-D degradation were observed at distinct times in soil that supported transconjugant populations compared to controls in which no gene transfer was detected.  相似文献   

5.
The distribution of tfdAα and cadA, genes encoding 2,4-dichlorophenoxyacetate (2,4-D)-degrading proteins which are characteristic of the 2,4-D-degrading Bradyrhizobium sp. isolated from pristine environments, was examined by PCR and Southern hybridization in several Bradyrhizobium strains including type strains of Bradyrhizobium japonicum USDA110 and Bradyrhizobium elkanii USDA94, in phylogenetically closely related Agromonas oligotrophica and Rhodopseudomonas palustris, and in 2,4-D-degrading Sphingomonas strains. All strains showed positive signals for tfdAα, and its phylogenetic tree was congruent with that of 16S rRNA genes in α-Proteobacteria, indicating evolution of tfdAα without horizontal gene transfer. The nucleotide sequence identities between tfdAα and canonical tfdA in β- and γ-Proteobacteria were 46 to 57%, and the deduced amino acid sequence of TfdAα revealed conserved residues characteristic of the active site of α-ketoglutarate-dependent dioxygenases. On the other hand, cadA showed limited distribution in 2,4-D-degrading Bradyrhizobium sp. and Sphingomonas sp. and some strains of non-2,4-D-degrading B. elkanii. The cadA genes were phylogenetically separated between 2,4-D-degrading and nondegrading strains, and the cadA genes of 2,4-D degrading strains were further separated between Bradyrhizobium sp. and Sphingomonas sp., indicating the incongruency of cadA with 16S rRNA genes. The nucleotide sequence identities between cadA and tftA of 2,4,5-trichlorophenoxyacetate-degrading Burkholderia cepacia AC1100 were 46 to 53%. Although all root nodule Bradyrhizobium strains were unable to degrade 2,4-D, three strains carrying cadA homologs degraded 4-chlorophenoxyacetate with the accumulation of 4-chlorophenol as an intermediate, suggesting the involvement of cadA homologs in the cleavage of the aryl ether linkage. Based on codon usage patterns and GC content, it was suggested that the cadA genes of 2,4-D-degrading and nondegrading Bradyrhizobium spp. have different origins and that the genes would be obtained in the former through horizontal gene transfer.  相似文献   

6.
7.
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   

8.
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   

9.
Rhizodegradation is a technique involving plants that offers interesting potential to enhance biodegradation of persistent organic pollutants such as polycyclic aromatic hydrocarbons (PAHs). Nevertheless, the behaviour of PAHs in plant rhizosphere, including micro-organisms and the physico-chemical soil properties, still needs to be clarified. The present work proposes to study the toxicity and the dissipation of phenanthrene in three artificially contaminated soils (1 g kg-1 DW). Experiments were carried out after 2 months of soil aging. They consisted in using different systems with two plant species (Ryegrass—Lolium perenne L. var. Prana and red clover—Trifolium pratense L. var. fourragère Caillard), three kinds of soils (a silty-clay-loam soil “La Bouzule”, a coarse sandy-loam soil “Chenevières” and a fine sandy-loam soil “Maconcourt”). Phenanthrene was quantified by HPLC in the beginning (T 0) and the end of the experiments (30 days). Plant biomass, microbial communities including mycorrhizal fungi, Rhizobium and PAH degraders were also recorded. Generally phenanthrene contamination did not affect plant biomass. Only the red clover biomass was enhanced in Chenevières and La Bouzule polluted soils. A stimulation of Rhizobium red clover colonisation was quantified in spiked soils whereas a drastic negative phenanthrene effect on the mycorrhization of ryegrass and red clover was recorded. The number of PAH degraders was stimulated by the presence of phenanthrene in all tested soils. Both in ryegrass and red clover planted soils, the highest phenanthrene dissipation due to the rhizosphere was measured in La Bouzule soils. On the contrary, in non-planted soils, La Bouzule soils had also the lowest pollutant dissipation. Thus, in rhizospheric and non-rhizospheric soils the phenanthrene dissipation was found to depend on soil clay content.  相似文献   

10.
Earthworm egg capsules (cocoons) may acquire bacteria from the environment in which they are produced. We found that Ralstonia eutropha (pJP4) can be recovered from Eisenia fetida cocoons formed in soil inoculated with this bacterium. Plasmid pJP4 contains the genes necessary for 2,4-dichlorophenoxyacetic acid (2,4-D) and 2,4-dichlorophenol (2,4-DCP) degradation. In this study we determined that the presence of R. eutropha (pJP4) within the developing earthworm cocoon can influence the degradation and toxicity of 2,4-D and 2,4-DCP, respectively. The addition of cocoons containing R. eutropha (pJP4) at either low or high densities (102 or 105 CFU per cocoon, respectively) initiated degradation of 2,4-D in nonsterile soil microcosms. Loss of 2,4-D was observed within the first week of incubation, and respiking the soil with 2,4-D showed depletion within 24 h. Microbial analysis of the soil revealed the presence of approximately 104 CFU R. eutropha (pJP4) g−1 of soil. The toxicity of 2,4-DCP to developing earthworms was tested by using cocoons with or without R. eutropha (pJP4). Results showed that cocoons containing R. eutropha (pJP4) were able to tolerate higher levels of 2,4-DCP. Our results indicate that the biodegradation of 2,4-DCP by R. eutropha (pJP4) within the cocoons may be the mechanism contributing to toxicity reduction. These results suggest that the microbiota may influence the survival of developing earthworms exposed to toxic chemicals. In addition, cocoons can be used as inoculants for the introduction into the environment of beneficial bacteria, such as strains with biodegradative capabilities.  相似文献   

11.
Enhanced biodegradation of organic xenobiotic compounds in the rhizosphere is frequently recorded although the specific mechanisms are poorly understood. We have shown that the mineralization of 2,4-dichlorophenoxyacetic acid (2,4-D) is enhanced in soil collected from the rhizosphere of Trifolium pratense[e.g. maximum mineralization rate=7.9 days-1 and time at maximum rate (t1)=16.7 days for 12-day-old T. pratense soil in comparison with 4.7 days-1 and 25.4 days, respectively, for non-planted controls). The purpose of this study was to gain a better understanding of the plant-microbe interactions involved in rhizosphere-enhanced biodegradation by narrowing down the identity of the T. pratense rhizodeposit responsible for stimulating the microbial mineralization of 2,4-D. Specifically, we investigated the distribution of the stimulatory component(s) among rhizodeposit fractions (exudates or root debris) and the influence of soil properties and plant species on its production. Production of the stimulatory rhizodeposit was dependent on soil pH (e.g. t1 for roots grown at pH 6.5 was significantly lower than for those grown at pH 4.4) but independent of soil inorganic N concentration. Most strikingly, the stimulatory rhizodeposit was only produced by T. pratense grown in non-sterile soil and was present in both exudates and root debris. Comparison of the effect of root debris from plant species (three each) from the classes monocotyledon, dicotyledon (non-legume) and dicotyledon (legume) revealed that legumes had by far the greatest positive impact on 2,4-D mineralization kinetics. We discuss the significance of these findings with respect to legume-rhizobia interactions in the rhizosphere.  相似文献   

12.
Uptake of 2,4-dichlorophenoxyacetate (2,4-D) by Ralstonia eutropha JMP134(pJP4) was studied and shown to be an energy-dependent process. The uptake system was inducible with 2,4-D and followed saturation kinetics in a concentration range of up to 60 μM, implying the involvement of a protein in the transport process. We identified an open reading frame on plasmid pJP4, which was designated tfdK, whose translation product TfdK was highly hydrophobic and showed resemblance to transport proteins of the major facilitator superfamily. An interruption of the tfdK gene on plasmid pJP4 decimated 2,4-D uptake rates, which implies a role for TfdK in uptake. A tfdA mutant, which was blocked in the first step of 2,4-D metabolism, still took up 2,4-D. A mathematical model describing TfdK as an active transporter at low micromolar concentrations fitted the observed uptake data best.  相似文献   

13.
A new and rapid protocol for optimum callus production and complete plant regeneration has been assessed in Malaysian upland rice (Oryza sativa) cv. Panderas. The effect of plant growth regulator (PGR) on the regeneration frequency of Malaysian upland rice (cv. Panderas) was investigated. Mature seeds were used as a starting material for callus induction experiment using various concentrations of 2,4-D and NAA. Optimal callus induction frequency at 90% was obtained on MS media containing 2,4-D (3 mg L−1) and NAA (2 mg L−1) after 6 weeks while no significant difference was seen on tryptophan and glutamine parameters. Embryogenic callus was recorded as compact, globular and light yellowish in color. The embryogenic callus morphology was further confirmed with scanning electron microscopy (SEM) analysis. For regeneration, induced calli were treated with various concentrations of Kin (0.5–1.5 mg L−1), BAP, NAA and 0.5 mg L−1 of TDZ. The result showed that the maximum regeneration frequency (100%) was achieved on MS medium containing BAP (0.5 mg L−1), Kin (1.5 mg L−1), NAA (0.5 mg L−1) and TDZ (0.5 mg L−1) within four weeks. Developed shoots were successfully rooted on half strength MS free hormone medium and later transferred into a pot containing soil for acclimatization. This cutting-edge finding is unique over the other existing publishable data due to the good regeneration response by producing a large number of shoots.Abbreviations: 2,4-D, 2,4-dichlorophenoxyacetic acid; NAA, naphthaleneacetic acid; Kin, kinetin; MS, Murashige and Skoog; BAP, benzylaminopurine; TDZ, thidiazuron  相似文献   

14.
The aim of this study was to evaluate how the in situ exposure of a Danish subsurface aquifer to phenoxy acid herbicides at low concentrations (<40 μg l−1) changes the microbial community composition. Sediment and groundwater samples were collected inside and outside the herbicide-exposed area and were analyzed for the presence of general microbial populations, Pseudomonas bacteria, and specific phenoxy acid degraders. Both culture-dependent and culture-independent methods were applied. The abundance of microbial phenoxy acid degraders (100 to 104 g−1 sediment) was determined by most probable number assays, and their presence was only detected in herbicide-exposed sediments. Similarly, PCR analysis showed that the 2,4-dichlorophenoxyacetic acid degradation pathway genes tfdA and tfdB (102 to 103 gene copies g−1 sediment) were only detected in sediments from contaminated areas of the aquifer. PCR-restriction fragment length polymorphism measurements demonstrated the presence of different populations of tfd genes, suggesting that the in situ herbicide degradation was caused by the activity of a heterogeneous population of phenoxy acid degraders. The number of Pseudomonas bacteria measured by either PCR or plating on selective agar media was higher in sediments subjected to high levels of phenoxy acid. Furthermore, high numbers of CFU compared to direct counting of 4′,6-diamidino-2-phenylindole-stained cells in the microscope suggested an increased culturability of the indigenous microbial communities from acclimated sediments. The findings of this study demonstrate that continuous exposure to low herbicide concentrations can markedly change the bacterial community composition of a subsurface aquifer.  相似文献   

15.
Our objective was to isolate and characterize indigenous bacteria able to use 2,4-D as a sole carbon (C) and energy source from an agricultural soil in the Sauce Grande River basin (Argentina). Culturable-dependant and molecular methods combined were used to identify and characterize putatively dominant indigenous degrading bacteria. Physiological traits, chloride release and biomass production showed the degradative capacity of the isolates obtained and high-performance liquid chromatography (HPLC) was used to corroborate the evidence. Degrading genes (tfdA and tfdB) were detected in all isolates, and their restriction fragment length polymorphisms (RFLP) were analyzed. Altogether, our results suggest that agricultural use of 2,4-D at recommended level leads to selection for a copiotrophic degrading population. The dominant genus able to metabolize 2,4-D in this soil was identified as Cupriavidus by 16S rRNA gene sequencing and the RFLP profiles of all isolates resembled that of Cupriavidus necator JMP134, the model organism for 2,4-D degradation. The strain EMA-G showed a remarkable performance in herbicide degradation (100 % removal in <1 day) in pure culture and is a favorite candidate for future biodegradation experiments.  相似文献   

16.
Colony hybridizations with a gene probe for enumeration of 2,4-dichlorophenoxy-acetic acid (2,4-D)-degrading bacteria were compared with classical enrichment and radiolabel most-probable-number (MPN) assay methods. Two natural water samples (rivers) and raw sewage were tested by each method. UV scans of enrichment cultures revealed 2,4-D degradation with raw sewage occurred in 4–11 days, 4–>22 days with Mary's River water, and 5–>22 days with Willamette River water. [14C]-2,4-D MPN analysis, measuring release of14CO2, yielded estimates of bacteria per milliliter able to degrade 2,4-D. Raw sewage estimates were 1.4 × 105 2,4-D degraders/ml, Mary's River >1.6 × 105/ml, and Willamette River water 1.6 × 104/ml. Activities noted by UV scan enrichment data supported these results.Autoradiograms of colony blots were also used to estimate numbers of 2,4-D-degrading bacteria. These estimates were also supported by the UV scan data from enrichment cultures. Raw sewage gave counts between 5 × 104 and 2.9 × 105 2,4-D-degrading bacteria/ml, which correlates well with the estimates obtained by14C-MPN analyses. River waters, both much lower in total bacterial counts and organic carbon than raw sewage, yielded fewer 2,4-D-degrading bacteria than estimated by14C-MPN. Media composition and cometabolism may account for discrepancies in estimates for 2,4-D-degrading bacteria observed when colony blot and14C-MPN analyses were compared.Replica plating made it possible to test for 2,4-D biodegradation from colonies reactive in autoradiograms. Five of 12 (42%) colonies reacting in the colony hybridization exhibited biodegradation activities. Nonreactive colonies failed to degrade 2,4-D.  相似文献   

17.
A genetically engineered microorganism, Pseudomonas putida PPO301(pRO103), and the plasmidless parent strain, PPO301, were added at approximately 107 CFU/g of soil amended with 500 ppm of 2,4-dichlorophenoxyacetate (2,4-D) (500 μg/g). The degradation of 2,4-D and the accumulation of a single metabolite, identified by gas chromatography-mass spectrophotometry as 2,4-dichlorophenol (2,4-DCP), occurred only in soil inoculated with PPO301(pRO103), wherein 2,4-DCP accumulated to >70 ppm for 5 weeks and the concentration of 2,4-D was reduced to <100 ppm. Coincident with the accumulation of 2,4-DCP was a >400-fold decline in the numbers of fungal propagules and a marked reduction in the rate of CO2 evolution, whereas 2,4-D did not depress either fungal propagules or respiration of the soil microbiota. 2,4-DCP did not appear to depress the numbers of total heterotrophic, sporeforming, or chitin-utilizing bacteria. In vitro and in situ assays conducted with 2,4-DCP and fungal isolates from the soil demonstrated that 2,4-DCP was toxic to fungal propagules at concentrations below those detected in the soil.  相似文献   

18.
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   

19.
31 different bacterial strains isolated using the herbicide 2,4-dichlorophenoxyacetic acid (2,4-D) as the sole source of carbon, were investigated for their ability to mineralize 2,4-D and the related herbicide 4-chloro-2-methylphenoxyacetic acid (MCPA). Most of the strains mineralize 2,4-D considerably faster than MCPA. Three novel primer sets were developed enabling amplification of full-length coding sequences (CDS) of the three known tfdA gene classes known to be involved in phenoxy acid degradation. 16S rRNA genes were also sequenced; and in order to investigate possible linkage between tfdA gene classes and bacterial species, tfdA and 16S rRNA gene phylogeny was compared. Three distinctly different classes of tfdA genes were observed, with class I tfdA sequences further partitioned into the two sub-classes I-a and I-b based on more subtle differences. Comparison of phylogenies derived from 16S rRNA gene sequences and tfdA gene sequences revealed that most class II tfdA genes were encoded by Burkholderia sp., while class I-a, I-b and III genes were found in a more diverse array of bacteria.  相似文献   

20.
The 2,4-dichlorophenoxyacetic acid (2,4-D) degrading bacterium, Burkholderia cepacia (formerly Pseudomonas cepacia) DBO1(pRO101) was coated on non-sterile barley (Hordeum vulgare) seeds, which were planted in two non-sterile soils amended with varying amounts of 2,4-D herbicide. In the presence of 10 or 100 mg 2,4-D per kg soil B. cepacia DBO1(pRO101) readily colonized the root at densities up to 107 CFU per cm root. In soil without 2,4-D the bacterium showed weak root colonization. The seeds coated with B. cepacia DBO1(pRO101) were able to germinate and grow in soils containing 10 or 100 mg kg–1 2,4-D, while non-coated seeds either did not germinate or quickly withered after germination. The results suggest that colonization of the plant roots by the herbicide-degrading B. cepacia DBO1(pRO101) can protect the plant by degradation of the herbicide in the rhizosphere soil. The study shows that the ability to degrade certain pesticides should be considered, when searching for potential plant growth-promoting rhizobacteria. The role of root colonization by xenobiotic degrading bacteria is further discussed in relation to bioremediation of contaminated soils.  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