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 46 毫秒
1.
Lichenysins are surface-active lipopeptides with antibiotic properties produced nonribosomally by several strains of Bacillus licheniformis. Here, we report the cloning and sequencing of an entire 26.6-kb lichenysin biosynthesis operon from B. licheniformis ATCC 10716. Three large open reading frames coding for peptide synthetases, designated licA, licB (three modules each), and licC (one module), could be detected, followed by a gene, licTE, coding for a thioesterase-like protein. The domain structure of the seven identified modules, which resembles that of the surfactin synthetases SrfA-A to -C, showed two epimerization domains attached to the third and sixth modules. The substrate specificity of the first, fifth, and seventh recombinant adenylation domains of LicA to -C (cloned and expressed in Escherichia coli) was determined to be Gln, Asp, and Ile (with minor Val and Leu substitutions), respectively. Therefore, we suppose that the identified biosynthesis operon is responsible for the production of a lichenysin variant with the primary amino acid sequence l-Gln–l-Leu–d-Leu–l-Val–l-Asp–d-Leu–l-Ile, with minor Leu and Val substitutions at the seventh position.Many strains of Bacillus are known to produce lipopeptides with remarkable surface-active properties (11). The most prominent of these powerful lipopeptides is surfactin from Bacillus subtilis (1). Surfactin is an acylated cyclic heptapeptide that reduces the surface tension of water from 72 to 27 mN m−1 even in a concentration below 0.05% and shows some antibacterial and antifungal activities (1). Some B. subtilis strains are also known to produce other, structurally related lipoheptapeptides (Table (Table1),1), like iturin (32, 34) and bacillomycin (3, 27, 30), or the lipodecapeptides fengycin (50) and plipastatin (29).

TABLE 1

Lipoheptapeptide antibiotics of Bacillus spp.
LipopeptideOrganismStructureReference
Lichenysin AB. licheniformisFAa-L-Glu-L-Leu-D-Leu-L-Val-L-Asn-D-Leu-L-Ile51, 52
Lichenysin BFAa-L-Glu-L-Leu-D-Leu-L-Val-L-Asp-D-Leu-L-Leu23, 26
Lichenysin CFAa-L-Glu-L-Leu-D-Leu-L-Val-L-Asp-D-Leu-L-Ile17
Lichenysin DFAa-L-Gln-L-Leu-D-Leu-L-Val-L-Asp-D-Leu-L-IleThis work
Surfactant 86B. licheniformisFAa-L-Glxd-L-Leu-D-Leu-L-Val-L-Asxd-D-Leu-L-Ilee14, 15
L-Val
SurfactinB. subtilisFAa-L-Glu-L-Leu-D-Leu-L-Val-L-Asp-D-Leu-L-Leu1, 7, 49
EsperinB. subtilisFAb-L-Glu-L-Leu-D-Leu-L-Val-L-Asp-D-Leu-L-Leue45
L-Val 
Iturin AB. subtilisFAc-L-Asn-D-Tyr-D-Asn-L-Gln-L-Pro-D-Asn-L-Ser32
Iturin CFAc-L-Asn-D-Tyr-D-Asn-L-Gln-L-Pro-D-Asne-L-Asne34
D-Ser-L-Thr 
Bacillomycin LB. subtilisFAc-L-Asp-D-Tyr-D-Asn-L-Ser-L-Gln-D-Proe-L-Thr3
D-Ser- 
Bacillomycin DFAc-L-Asp-D-Tyr-D-Asn-L-Pro-L-Glu-D-Ser-L-Thr30, 31
Bacillomycin FFAc-L-Asn-D-Tyr-D-Asn-L-Gln-L-Pro-D-Asn-L-Thr27
Open in a separate windowaFA, β-hydroxy fatty acid. The β-hydroxy group forms an ester bond with the carboxy group of the C-terminal amino acid. bFA, β-hydroxy fatty acid. The β-hydroxy group forms an ester bond with the carboxy group of Asp5. cFA, β-amino fatty acid. The β-amino group forms a peptide bond with the carboxy group of the C-terminal amino acid. dOnly the following combinations of amino acid 1 and 5 are allowed: Gln-Asp or Glu-Asn. eWhere an alternative amino acid may be present in a structure, the alternative is also presented. In addition to B. subtilis, several strains of Bacillus licheniformis have been described as producing the lipopeptide lichenysin (14, 17, 23, 26, 51). Lichenysins can be grouped under the general sequence l-Glx–l-Leu–d-Leu–l-Val–l-Asx–d-Leu–l-Ile/Leu/Val (Table (Table1).1). The first amino acid is connected to a β-hydroxyl fatty acid, and the carboxy-terminal amino acid forms a lactone ring to the β-OH group of the lipophilic part of the molecule. In contrast to the lipopeptide surfactin, lichenysins seem to be synthesized during growth under aerobic and anaerobic conditions (16, 51). The isolation of lichenysins from cells growing on liquid mineral salt medium on glucose or sucrose basic has been studied intensively. Antimicrobial properties and the ability to reduce the surface tension of water have also been described (14, 17, 26, 51). The structural elucidation of the compounds revealed slight differences, depending on the producer strain. Various distributions of branched and linear fatty acid moieties of diverse lengths and amino acid variations in three defined positions have been identified (Table (Table11).In contrast to the well-defined methods for isolation and structural characterization of lichenysins, little is known about the biosynthetic mechanisms of lichenysin production. The structural similarity of lichenysins and surfactin suggests that the peptide moiety is produced nonribosomally by multifunctional peptide synthetases (7, 13, 25, 49, 53). Peptide synthetases from bacterial and fungal sources describe an alternative route in peptide bond formation in addition to the ubiquitous ribosomal pathway. Here, large multienzyme complexes affect the ordered recognition, activation, and linking of amino acids by utilizing the thiotemplate mechanism (19, 24, 25). According to this model, peptide synthetases activate their substrate amino acids as aminoacyl adenylates by ATP hydrolysis. These unstable intermediates are subsequently transferred to a covalently enzyme-bound 4′-phosphopantetheinyl cofactor as thioesters. The thioesterified amino acids are then integrated into the peptide product through a stepwise elongation by a series of transpeptidations directed from the amino terminals to the carboxy terminals. Peptide synthetases have not only awakened interest because of their mechanistic features; many of the nonribosomally processed peptide products also possess important biological and medical properties.In this report we describe the identification and characterization of a putative lichenysin biosynthesis operon from B. licheniformis ATCC 10716. Cloning and sequencing of the entire lic operon (26.6 kb) revealed three genes, licA, licB, and licC, with structural patterns common to peptide synthetases and a gene designated licTE, which codes for a putative thioesterase. The modular organization of the sequenced genes resembles the requirements for the biosynthesis of the heptapeptide lichenysin. Based on the arrangement of the seven identified modules and the tested substrate specificities, we propose that the identified genes are involved in the nonribosomal synthesis of the portion of the lichenysin peptide with the primary sequence l-Gln–l-Leu–d-Leu–l-Val–l-Asp–d-Leu–l-Ile (with minor Val and Leu substitutions).  相似文献   

2.
3.
4.
Pseudomonas strain ADP metabolizes the herbicide atrazine via three enzymatic steps, encoded by the genes atzABC, to yield cyanuric acid, a nitrogen source for many bacteria. Here, we show that five geographically distinct atrazine-degrading bacteria contain genes homologous to atzA, -B, and -C. The sequence identities of the atz genes from different atrazine-degrading bacteria were greater than 99% in all pairwise comparisons. This differs from bacterial genes involved in the catabolism of other chlorinated compounds, for which the average sequence identity in pairwise comparisons of the known members of a class ranged from 25 to 56%. Our results indicate that globally distributed atrazine-catabolic genes are highly conserved in diverse genera of bacteria.Atrazine [2-chloro-4-(ethylamino)-6-(isopropylamino)- 1,3,5-triazine] is a herbicide used for controlling broad-leaf and grassy weeds and is relatively persistent in soils (51). Atrazine and other s-triazine compounds have been detected in ground and surface waters at levels exceeding the Environmental Protection Agency’s maximum contaminant level of 3 ppb (30).Microbial populations exposed to synthetic chlorinated compounds, such as atrazine, often respond by producing enzymes that degrade these molecules. Most of our current understanding of the genes and enzymes involved in atrazine degradation derives from studies using Pseudomonas strain ADP, in which the first three enzymatic steps in atrazine degradation have been defined (6, 14, 15, 48). The genes atz A, -B, and -C, which encode these enzymes, have been cloned and sequenced. Atrazine chlorohydrolase (AtzA), hydroxyatrazine ethylaminohydrolase (AtzB), and N-isopropylammelide isopropylaminohydrolase (AtzC) sequentially convert atrazine to cyanuric acid (6, 14, 15, 48) (Fig. (Fig.1).1). Cyanuric acid and related compounds are catabolized by many soil bacteria (10, 11, 17, 24, 26, 61), and by Pseudomonas sp. ADP, to carbon dioxide and ammonia (35). This provides the evolutionary pressure for the atzA, -B, and -C genes to permit bacterial growth on the more than one billion pounds of atrazine that have been applied to soils globally (20). Here we used a knowledge of the atzA, -B, and -C gene sequences to investigate the presence of homologous genes in other atrazine-degrading bacteria. In this study, we report that five atrazine-degrading microorganisms, which were recently isolated from geographically separated sites exposed to atrazine, contained nearly identical atzA, -B, and -C genes. Open in a separate windowFIG. 1Pathway for atrazine catabolism to cyanuric acid in Pseudomonas sp. strain ADP.

Atrazine-catabolizing bacteria used in this study.

Until recently, attempts at isolating bacteria (18) or fungi (27) that completely degrade atrazine to carbon dioxide, ammonia, and chloride were unsuccessful. While several microorganisms were shown to dealkylate atrazine, they were unable to displace the chlorine atom (41, 54). Since 1994, several research groups have independently isolated atrazine-degrading bacteria that displaced the chlorine atom and mineralized atrazine (3, 7, 13, 35, 39, 46). Six of these bacterial cultures, listed in Table Table1,1, were studied here, and the Clavibacter strain had been investigated previously (13).

TABLE 1

Recently isolated atrazine-catabolizing bacteria
GenusStrainLocation where isolatedYr reported (reference)
PseudomonasaADPAgricultural-chemical dealership site, Little Falls, Minn.1995 (35)
RalstoniaaM91-3Agricultural soil, Ohio1995 (46, 55)
Mixed cultureBasel, Switzerland1995 (57)
ClavibacterAgricultural soil, Riverside, Calif.1996 (13)
AgrobacteriumJ14aAgricultural soil, Nebraska1996 (39)
NDb38/38Atrazine-contaminated soil, Indiana1996 (3)
AlcaligenesaSG1Industrial settling pond, San Gabriel, La.1997 (7)
Open in a separate windowaIsolate identity based on 16S rRNA sequence analysis. bND, not determined. 

Detection of atzA, -B, and -C homologs in atrazine-degrading microorganisms by PCR analysis.

Recently isolated atrazine-degrading bacteria were screened for the presence of DNA homologous to the Pseudomonas strain ADP atzABC genes, which encode enzymes transforming atrazine to cyanuric acid (Fig. (Fig.1).1). Total genomic DNA was isolated from each of these bacteria as described elsewhere (49), and the PCR technique was used to amplify sequences internal to the atzA, -B, and -C genes as described elsewhere (13). Custom primers were designed specifically for atzA (5′CCATGTGAACCAGATCCT3′ and 5′TGAAGCGTCCACATTACC3′), atzB (5′TCACCGGGGATGTCGCGGGC3′ and 5′CTCTCCCGCATGGCATCGGG3′), and atzC (5′GCTCACATGCAGGTACTCCA3′ and 5′GTACCATATCACCGTTTGCCA3′) by using the Primer Designer package, version 2.01 (Scientific and Educational Software, State Line, Pa.), and were synthesized by Gibco BRL (Gaithersburg, Md.). PCR fragments were amplified by using Taq DNA polymerase (Gibco BRL) (22) and were separated from primers on a 1.0% agarose gel. The results of these studies (Fig. (Fig.2)2) indicated that PCR amplification consistently produced DNA fragments of 0.5 kb for all organisms when the atzA or -B primers were used and fragments of 0.6 kb when the atzC primers were used. Open in a separate windowFIG. 2PCR analysis with primers designed to amplify internal regions of atzA (lanes 1 to 5), atzB (lanes 6 to 10), and atzC (lanes 11 to 15). The atrazine-degrading bacteria analyzed were Pseudomonas strain ADP (35) (lanes 1, 6, and 11), Alcaligenes strain SGI (7) (lanes 2, 7, and 12), Ralstonia strain M91-3 (46) (lanes 3, 8, and 13), Agrobacterium strain J14a (39) (lanes 4, 9, and 14), and isolate 38/38 (3) (lanes 5, 10, and 15). Values to the right of the gel are sizes (in kilobase pairs).Southern hybridization analyses were performed on the PCR-amplified DNA as described elsewhere (49) to confirm the presence of homologous DNA. We used a 0.6-kb ApaI/PstI fragment from pMD4 (15), a 1.5-kb BglII fragment from pATZB-2 (6), and a 2.0-kb EcoRI/AvaI fragment from pTD2.5 (48) as probes for atzA, -B, and -C genes, respectively. DNA probes were labeled with [α-32P]dCTP by using the Rediprime Random Primer Labeling Kit (Amersham Life Science, Arlington Heights, Ill.) according to the manufacturer’s instructions. Southern hybridization analyses, performed under stringent conditions, confirmed that each strain contained DNA homologous to atzA, -B, and -C (data not shown). With strain M91-3 and isolate 38/38, however, in addition to the expected 0.5-kb atzB PCR product (Fig. (Fig.2,2, lanes 8 and 10), a 1.2-kb fragment was also obtained. However, no hybridization to this fragment was seen with the atzB probe. Similar investigations showed that a mixed culture obtained from Switzerland (Table (Table1),1), capable of degrading atrazine, also contained DNA homologous to all three atz genes (12).As a negative control, bacteria known not to degrade atrazine were analyzed. PCR analyses were carried out with genomic DNA from the following randomly chosen laboratory strains: Rhodococcus chlorophenolicus (1), Flavobacterium sp. (47), Streptomyces coelicolor M145 (21), Amycolatopsis mediterranei (19), Agrobacterium strain A136 and strain A348 (A136/pTiA6NC) (60), Arthrobacter globiformis MN1 (45), Bradyrhizobium japonicum (33), Rhizobium sp. strain NGR 234 (44), Pseudomonas NRRLB12228, and Klebsiella pneumoniae 99 (16). None of these strains contained DNA that was amplified by PCR using the primers designed to identify the atzA, -B, or -C gene (data not shown).

DNA sequences of atzA, -B, and -C homologs in atrazine-degrading microorganisms.

DNAs amplified from the five strains in Table Table11 with the atzA, -B, and -C primers were purified from gel slices by using the GeneClean II System (Bio 101, Inc., Vista, Calif.) and sequenced with a PRISM Ready Reaction DyeDeoxy Terminator Cycle Sequencing kit (Perkin-Elmer Corp., Norwalk, Conn.) and an ABI model 373A DNA sequencer (Applied Biosystems, Foster City, Calif.). The GCG sequence analysis software package (Genetics Computer Group, Inc., Madison, Wis.) was used for all DNA and protein sequence comparisons and alignments. Table Table22 summarizes these data. The PCR-amplified genes were ≥99% identical to the Pseudomonas strain ADP atzA, -B, and -C genes in all pairwise comparisons of DNA sequences. This remarkable sequence identity suggested that each atz gene in the different genera was derived from a common ancestor and that they have diverged evolutionarily only to a limited extent.

TABLE 2

Sequence identities of atzABC homologs from different atrazine-degrading bacteria
Strain% DNA sequence identitya
atzAatzBatzC
Pseudomonas ADP100100100
Alcaligenes SG199.2100100
Ralstonia M91-399.0100100
Agrobacterium J14a99.1100100
Isolate 38/3899.310099.8
Open in a separate windowaDNA sequences obtained from each strain by using the ataA, -B, and -C primers were compared with the atzABC gene sequences from Pseudomonas strain ADP. A review of the literature on other bacterial catabolic pathways indicated a much greater degree of divergence when genes encoding enzymes for the catabolism of other commercially relevant chlorinated compounds were compared (Table (Table3).3). As with atrazine, multiple bacterial strains that catabolize 1,2-dichloroethane, chloroacetic acid, 2,4-dichlorophenoxyacetate, dichloromethane, and 4-chlorobenzoate have been isolated. A comparison of the gene sequences encoding the initiating reactions in the catabolism of each of those compounds revealed that sequence divergence was comparatively high. In pairwise comparisons within each gene class, the average sequence identities ranged from 25 to 56% (divergence was 46 to 75%). With the atzABC genes, by contrast, there is at most a 1% sequence difference within the sequenced gene region (Table (Table2).2). Moreover, the atzB sequences were completely identical, and the atzC genes diverged by only 1 bp in one of the five strains tested. This suggests that the atz genes recently arose from a single origin and have become distributed globally. Similarly, identical parathion hydrolase genes were isolated from two bacteria representing different genera and global locations (40, 52, 53).

TABLE 3

Sequence comparisons of isofunctional bacterial enzymes that catabolize chlorinated compounds
GeneEnzymeAverage % protein sequence identitya (no. of pairwise comparisons)References
dhlA, dhaAHaloalkane dehalogenase25.0 (1)23, 31
dehC, hadL, dehH, dehH1, dehH2, dhlB, dehCI, dehCII2-Haloacid dehalogenase36.6 ± 3.9 (36)5, 25, 28, 29, 42, 43, 50, 59
tfdA2,4-Dichlorophenoxyacetate monooxygenase43.2 ± 4.6 (21)b34, 37, 38, 56, 58
dcmADichloromethane dehalogenase56.0 (1)4, 32
atzAAtrazine chlorohydrolase98.6 ± 0.12 (15)cThis study
atzBHydroxyatrazine ethylaminohydrolase100 (10)cThis study
atzCN-Isopropylammelide isopropylaminohydrolase99.0 ± 0.43 (10)cThis study
Open in a separate windowaAll possible pairwise alignments of translated gene sequences were made. The average percent identity is the mean of the percent identity values for all pairwise alignments ± standard error of the mean. bIncludes full protein sequences as well as partial protein sequences of ≥100 amino acids. cSequence identity within a 0.5-kb PCR product for atzA and -B and within a 0.6-kb PCR product for atzC. Six sequences were analyzed for atzA, and five were analyzed for atzB and -C. The data presented here provide further support for previous studies suggesting that hydroxyatrazine in the environment derives from biological processes (36), and not solely from abiotic reactions (2, 9). The present data, and a recent report by Bouquard et al. (8), indicate that the gene encoding atrazine chlorohydrolase is widespread in the United States and Europe.Our observations argue for a single, recent evolutionary origin of the atz genes and their subsequent global distribution. We have recently localized the atzA, -B, and -C genes to a large, self-transmissible plasmid in Pseudomonas strain ADP (12), and possible mechanisms of transfer of the atzABC genes are currently under investigation.  相似文献   

5.
6.
7.
8.
In the last two years, because of advances in protein separation and mass spectrometry, top-down mass spectrometry moved from analyzing single proteins to analyzing complex samples and identifying hundreds and even thousands of proteins. However, computational tools for database search of top-down spectra against protein databases are still in their infancy. We describe MS-Align+, a fast algorithm for top-down protein identification based on spectral alignment that enables searches for unexpected post-translational modifications. We also propose a method for evaluating statistical significance of top-down protein identifications and further benchmark various software tools on two top-down data sets from Saccharomyces cerevisiae and Salmonella typhimurium. We demonstrate that MS-Align+ significantly increases the number of identified spectra as compared with MASCOT and OMSSA on both data sets. Although MS-Align+ and ProSightPC have similar performance on the Salmonella typhimurium data set, MS-Align+ outperforms ProSightPC on the (more complex) Saccharomyces cerevisiae data set.In the past two decades, proteomics was dominated by bottom-up mass spectrometry that analyzes digested peptides rather than intact proteins. Bottom-up approaches, although powerful, do have limitations in analyzing protein species, e.g. various proteolytic forms of the same protein or various protein isoforms resulting from alternative splicing. Top-down mass spectrometry focuses on analyzing intact proteins and large peptides (110) and has advantages in localizing multiple post-translational modifications (PTMs)1 in a coordinated fashion (e.g. combinatorial PTM code) and identifying multiple protein species (e.g. proteolytically processed protein species) (11). Until recently, most top-down studies were limited to single purified proteins (1215). Top-down studies of protein mixtures were restricted by difficulties in separating and fragmenting intact proteins and a shortage of robust computational tools.In the last two years, because of advances in protein separation and top-down instrumentation, top-down mass spectrometry moved from analyzing single proteins to analyzing complex samples containing hundreds and even thousands of proteins (1621). Because algorithms for interpreting top-down spectra are still in their infancy, many recent developments include computational innovations in protein identification.Because top-down spectra are complex, the first step in top-down spectral interpretation is usually spectral deconvolution, which converts a complex top-down spectrum to a list of monoisotopic masses (a deconvolved spectrum). Every protein (possibly with modifications) can be scored against a top-down deconvoluted spectrum, resulting in a Protein-Spectrum-Match (PrSM). The top-down protein identification problem is finding a protein in a database with the highest scoring PrSM for a top-down spectrum and further output the PrSM if it is statistically significant. There are several software tools for top-down protein identification (
SoftwareIdentification of unexpected modificationsProteogenomics search against 6-frame translationSpeedEstimation of statistical significance
ProSightPC+/−a+Fast/Slowb+
PIITA+/−Fast
UStag++Fast
MS-TopDown+Slow
MS-Align+++Fast+
Open in a separate windowa ProSightPC has various search modes that contribute to bridging the gap between blind and restrictive modes of MS/MS database search. It can identify truncated proteins by using biomarker search and identify unexpected modifications by using Δm mode and setting the error tolerance of precursor mass to a large value (e.g., 1999 Da). However, it is not designed for identifying truncated proteins with unexpected PTMs which are not represented in the “shotgun annotated” database.b In its most advances mode, ProSightPC can search the annotated top-down database that contains various protein species. However, ProSightPC searches in this mode become an order of magnitude slower.We describe MS-Align+, a fast software tool for top-down protein identification. MS-Align+ shares the spectral alignment approach with MS-TopDown, but greatly improves on speed, statistical analysis (providing E-values of PrSMs), and the number of identified PrSMs (e.g. by finding spectral alignments between spectra and truncated proteins). We benchmarked various tools for top-down protein identification on two data sets from Saccharomyces cerevisiae (SC) and Salmonella typhimurium (ST). We demonstrate that MS-Align+ significantly increase the number of identified spectra as compared with MASCOT and OMSSA on both data sets. Although MS-Align+ and ProSightPC have similar performance on the ST data set, MS-Align+ outperforms ProSightPC on the more complex SC data set.  相似文献   

9.
Evidence for Interspecies Gene Transfer in the Evolution of 2,4-Dichlorophenoxyacetic Acid Degraders     
Catherine McGowan  Roberta Fulthorpe  Alice Wright  J. M. Tiedje 《Applied and environmental microbiology》1998,64(10):4089-4092
Small-subunit ribosomal DNA (SSU rDNA) from 20 phenotypically distinct strains of 2,4-dichlorophenoxyacetic acid (2,4-D)-degrading bacteria was partially sequenced, yielding 18 unique strains belonging to members of the alpha, beta, and gamma subgroups of the class Proteobacteria. To understand the origin of 2,4-D degradation in this diverse collection, the first gene in the 2,4-D pathway, tfdA, was sequenced. The sequences fell into three unique classes found in various members of the beta and gamma subgroups of Proteobacteria. None of the α-Proteobacteria yielded tfdA PCR products. A comparison of the dendrogram of the tfdA genes with that of the SSU rDNA genes demonstrated incongruency in phylogenies, and hence 2,4-D degradation must have originated from gene transfer between species. Only those strains with tfdA sequences highly similar to the tfdA sequence of strain JMP134 (tfdA class I) transferred all the 2,4-D genes and conferred the 2,4-D degradation phenotype to a Burkholderia cepacia recipient.Bacteria capable of mineralizing 2,4-dichlorophenoxyacetic acid (2,4-D), a commonly used herbicide, are found in many different phylogenetic groups (2, 3, 7, 11, 22, 23). Evidence suggests that numerous variants of 2,4-D catabolic genes exist and that catabolic operons consist of a near-random mixing of these variants (7). Interspecies gene transfer is a well-documented phenomenon (13), and horizontal gene transfer of the 2,4-D-degrading plasmid pJP4 has been shown (3, 5). However, not all 2,4-D catabolic operons are found on plasmids (10, 11, 16, 20). The extent to which other 2,4-D genes have been exchanged in nature is unknown. The aim of this research was to assess the role of horizontal gene transfer in the evolution of 2,4-D-degrading strains. This article summarizes the results of two aspects of this work—the study of the transfer of the entire 2,4-D pathway by using standard mating experiments and a phylogenetic study of the tfdA gene. The tfdA gene codes for an α-ketoglutarate-dependent 2,4-D dioxygenase which converts 2,4-D into 2,4-dichlorophenol and glyoxylate (6). This 861-bp gene was first sequenced from Ralstonia eutropha JMP134 (19). Two more tfdA genes were cloned from chromosomal locations in Burkholderia strain RASC and Burkholderia strain TFD6 (16, 20). These proved to be identical to each other and 78.5% similar to the original. An alignment of the two variants allowed conserved areas to be identified and primers to be designed for the amplification of tfdA-like genes from other sources (24). Sequence analysis of putative tfdA fragments and the small-subunit ribosomal DNA (SSU rDNA) of the strains carrying them allowed us to construct phylogenies of the genes and their hosts and to look for congruency between them.

Mating experiments.

A collection of 2,4-D degraders containing 15 unique strains as determined by genomic fingerprinting (7) was used as a source of donors in a series of mating experiments (Table (Table1).1). Burkholderia cepacia D5, lacking the ability to grow on 2,4-D and not hybridizing to any tfd genes, was used as a recipient in mating experiments. Strain D5 contains neomycin phosphotransferase genes (nptII) carried on transposon Tn5 and is resistant to 50 μg each of kanamycin, carbenicillin, and bacitracin per ml. All of the 2,4-D strains used were sensitive to these antibiotics. Filter matings were performed with a donor-to-recipient ratio of 1:10. Colonies which grew on selective medium (500 ppm of 2,4-D in mineral salts agar [MMO] [23] including 50 μg of kanamycin, carbenicillin, and bacitracin per ml) were subjected to further tests. Their ability to catabolize 2,4-D was tested in liquid medium (same composition as that described above).

TABLE 1

2,4-D-degrading strains, geographic origins, and GenBank accession numbers
StrainGenBank accession no. (SSU rDNA)OriginMost similar to genus and/or speciesaTransferbtfdA typecGenBank accession no. (tfdA gene)Reference or source
JMP134AF049542AustraliaRalstonia eutropha+IM167303
EML1549AF049546OregonBurkholderia sp.+I2
TFD39AF049539SaskatchewanBurkholderia sp.+IU4319723
K712AF049543MichiganBurkholderia sp.+IU4327611
TFD9AF049537SaskatchewanAlcaligenes xylosoxidans+IU4327623
TFD41AF049541MichiganRalstonia eutropha+I23
TFD38AF049540MichiganRalstonia eutropha+NDc23
TFD23AF049536MichiganRhodoferax fermentans+IU4327623
RASCAF049544OregonBurkholderia sp.(+)IIU257172
TFD6AF049546MichiganBurkholderia sp.II23
TFD2AF049545MichiganBurkholderia sp.II23
TFD31AF049536SaskatchewanRhodoferax fermentansIII23
B6-9AF049538OntarioRhodoferax fermentansNDIIIU431969
I-18U22836OregonHalomonas sp.NDIIIU2249915
K1443AF049531MichiganSphingomonas sp.d11
2,4-D1AF049535MontanaSphingomonas sp.R. Sanford
B6-5AF049533OntarioSphingomonas sp.ND9
B6-10AF049534OntarioSphingomonas sp.ND9
EML146AF049532OregonSphingomonas sp.2
M1AF049530French PolynesiaRhodospeudomonas sp.NDR. Fulthorpe
Open in a separate windowaThe generus and/or species most similar to the strain is given based on similarities of SSU rDNA sequences. bSymbols: +, able to transfer 2,4-D degradation to B. cepacia D5; (+), able to transfer at very low frequency; −, no transfer detected. cND, not determined. d—, no amplificate was obtained. The disappearance of 2,4-D from the culture medium was monitored by high-performance liquid chromatography. Cells were removed by centrifugation, and the supernatant was filtered through 0.2-μm-pore-size filters. These samples were then analyzed on a Lichrosorb Rp-18 column (Anspec Co., Ann Arbor, Mich.) with 60% methanol–40% 0.1% H3PO4 as the eluant. 2,4-D was detected by measuring light absorption at 230 nm. The presence of tfd genes was detected by hybridizing colony blots with a DNA probe derived from the entire pJP4 plasmid. The identity of the colonies was confirmed by probing with the nptII gene of Tn5 (found in B. cepacia D5). Probes were labeled with random hexanucleotides incorporating [32P]dCTP (3,000 Ci/mmol; New England Nuclear, Boston, Mass.). Hybridizations were done under high-stringency conditions by using 50% formamide and Denhardt’s solution (18) at 42°C. Of the 15 unique strains tested, 9 transferred 2,4-D degradation abilities to D5. This transfer was confirmed by hybridization with pJP4 for eight of these strains. B. cepacia RASC could transfer degradative abilities, but neither it nor the transconjugant hybridized to the pJP4 probe. Work subsequent to this study has confirmed that the genes carried by RASC do not hybridize to those found on pJP4 under high-stringency conditions (7).

Phylogenetic analyses.

Total genomic DNA was isolated from 20 unique 2,4-D-degrading strains (including all 15 used for mating experiments) grown on 500 ppm of 2,4-D mineral salts medium amended with 50 ppm of yeast extract. SSU rDNA was amplified by using fD1 and rD1 as primers (25). Putative tfdA fragments were amplified by using primers TVU and TVL as previously described (24). PCR products were purified with a Gene Clean kit (Bio 101, La Jolla, Calif.). Sequencing was done with an Applied Biosystems model 373A automatic sequencer (Perkin-Elmer Cetus) by using fluorescently labeled dye termination at the Michigan State University Sequencing Facility. The sequencing primer used for SSU rDNA fragments was 519R (5′ GTA TTA CCG CGG CTG CTG G-3′). For tfdA fragments, the sequencing primers were the same as the amplification primers. GenBank accession numbers for these sequences are given in Table Table11.The SSU rDNA sequences were compared to sequences in GenBank by using the Basic Local Alignment Search Tool (BLAST) (1), and those strains with the highest maximal segment pair scores were retrieved from GenBank and included in the phylogenetic analysis. Sequences were aligned manually with the software SeqEd (Applied Biosystems) and with MacClade (14). Sites where nucleotides were not resolved for all sequences were deleted from the alignment, as were those nucleotides corresponding to the small loop in this region that is absent in the alpha subgroup of the class Proteobacteria. These deletions left 283 unambiguous sites for the construction of the SSU rDNA phylogenies. Phylogenetic trees were constructed by using the neighbor-joining analysis of pairwise Jukes-Cantor distances (4), and the topology was confirmed by using the maximum parsimony method PAUP (21). Desulfomonile tiedjei of the δ-Proteobacteria was used as an outgroup. Bootstrap analysis based on 100 replicates was used to place confidence estimates on the tree. Only bootstrap values of greater than 50 were used.

2,4-D degrader diversity.

The 2,4-D degraders in this study were distributed throughout the alpha, beta, and gamma subgroups of the Proteobacteria (Fig. (Fig.1).1). The lack of representation of gram-positive bacteria is likely a reflection of isolation methods, not of the lack of gram-positive 2,4-D degraders. The majority of these strains were members of the beta subgroup of Proteobacteria, five of which were most closely related to the genus Burkholderia, having at least 92% sequence similarity with each other. Three were closely related to Rhodoferax fermentans (close to the class Comamonadaceae), three were related to Ralstonia eutropha, and one was related to Alcaligenes xylosoxidans. TFD39 falls outside any clear cluster. One member of the γ-Proteobacteria, strain I-18, a haloalkaliphile, was found to be closely related to the salt-loving genus Halomonas (15). The remaining six strains all clustered in the alpha branch of Proteobacteria (Fig. (Fig.1).1). Of this subgroup, five were most closely related to the genus Sphingomonas. One member of the α-Proteobacteria, strain M1, which is the most oligotrophic and slow growing of all the strains used in this study, is 97% similar to Rhodopseudomonas palustris. The character of strain M1 correlates well with its phylogenetic placement near the slow-growing genus Bradyrhizobium. Open in a separate windowFIG. 1Neighbor-joining dendrogram (Jukes-Cantor distances) of SSU rDNA from 2,4-D-degrading bacteria (indicated in boldface type) and reference strains (indicated in italic type). Class I (•), class II (▴), and class III (■) types of tfdA genes are indicated. Bootstrap confidence limits (percentages) are indicated above each branch. Scale bar represents a Jukes-Cantor distance of 0.01.

tfdA gene fragments.

tfdA gene fragments were successfully amplified and sequenced from 10 strains of β-Proteobacteria and 1 strain of γ-Protobacteria. None of the strains from the α-Proteobacteria gave any amplificates with these primers. These 313 contiguous nucleotides were aligned with additional tfdA sequences from JMP134 and from strain RASC (Fig. (Fig.2).2). Three distinct classes of tfdA gene sequences with slight variations in each class were found. Class I included fragments from JMP134, TFD39, TFD23, K712, and TFD9 that differed from each other by 2 bp at the most. Class I tfdA genes are probably plasmid encoded. All strains with a class I tfdA gene examined so far contained broad-host-range, self-transmissible plasmids containing 2,4-D genes (2, 3, 11, 17). All of the strains with a class I tfdA gene were able to transfer the 2,4-D phenotype in the mating studies reported above. The class II tfdA sequences included identical fragments amplified from RASC, TFD6, and TFD2 which were 76% similar to those in class I. Class III included identical fragments from strains TFD31, B6-9, and I-18 which were 77% similar to class I genes and 80% similar to class II genes. Both class II and III tfdA genes differed from each other and from class I genes in the same nine sites corresponding to the third base pair of the codons. The tfdA phylogenetic tree is a simple one, with three distinct branches that are incongruent with the SSU rDNA-derived phylogeny (Fig. (Fig.3).3). Class I tfdA sequences were found in Burkholderia-like strains, in strains related to the Comamonas-Rhodoferax group, and in the Ralstonia-Acaligenes group, all in the β-Proteobacteria. Class II sequences are less widely distributed, found only in Burkholderia-like branches. However, even in this subgroup, this tfdA variant is found in strains that differ by 7% at the SSU rDNA level (RASC and TFD2). However, the class III sequences were most interesting, being found both in the Comamonas-Rhodoferax group and in a strain of the γ-Proteobacteria, I-18, strains that differ by 24% at the SSU rDNA level. Class III genes have since been found in a collection of randomly isolated non-2,4-D degraders, including gram-positive bacilli, as well as in various gram-negative bacteria, even though the gene is not expressed (10). Open in a separate windowFIG. 2Alignment of 313 nucleotides of internal fragments of tfdA genes from representative strains. Nucleotides identical to tfdA from pJP4 are represented by periods.Open in a separate windowFIG. 3Phylogenetic incongruency of tfdA genes and SSU rDNA from diverse 2,4-D-degrading bacteria. Dendrograms for tfdA and SSU rDNA are indicated. Shading indicates the type of tfdA sequence, either class I, II, or III. Note that branch lengths are not drawn to scale.An interesting result was the detection of two different tfdA gene variants in sibling strains. TFD23 and TFD31 are identical at the ribosomal gene level, but one harbors a class I gene and the other harbors a class III gene. Similarly, TFD6 and EML159 are rRNA siblings that carry a class II and class I gene, respectively.None of the α-Proteobacteria yielded a PCR product when amplified with the conserved tfdA primers. This finding complements our observation that none of these bacteria hybridized to the tfdA gene, even under conditions of low stringency, indicating that any tfdA-like genes in the α-Proteobacteria are likely to be more divergent from the ones sequenced here (7, 11). In addition, none of the Sphingomonas strains in the study hybridized with a whole pJP4 probe, and similarly, no Sphingomonas strains scored positive for transfer of 2,4-D-degrading ability to recipient B. cepacia D5. Together these results suggest a reduced gene flow between members of the α- and β- or γ-Proteobacteria or poor gene expression of β- or γ-derived genes by α-Proteobacteria. Although plasmid pJP4 is a broad-host-range plasmid and has been known to transfer to α-Proteobacteria such as Rhizobium and Agrobacterium species and to γ-Proteobacteria such as Pseudomonas putida, Pseudomonas fluorescens, and Pseudomonas aeruginosa, the 2,4-D pathway is not expressed in these strains of the α- or γ-Proteobacteria (3). Phylogenetically limited expression of plasmid-borne 3-chlorobenzoate-degradative genes has also been noted for the pseudomonads (8). Subsequent studies have found divergent but related sequences for the tfdB and tfdC genes in 2,4-D-degrading Sphingomonas strains (7, 12, 24).With the exceptions of the minor differences within the class I pJP4-like tfdA sequences, there were no intermediate tfdA sequences. The most likely explanation of this is that the rate of horizontal transfer of the tfd genes is high relative to the rate at which mutations can accumulate. Examination of sequences of tfdA genes from a greater variety of organisms may turn up more intermediate variation.  相似文献   

10.
Point Mutations in the Integron Integrase IntI1 That Affect Recombination and/or Substrate Recognition     
Annie Gravel  Nancy Messier  Paul H. Roy 《Journal of bacteriology》1998,180(20):5437-5442
The site-specific recombinase IntI1 found in class 1 integrons catalyzes the excision and integration of mobile gene cassettes, especially antibiotic resistance gene cassettes, with a site-specific recombination system. The integron integrase belongs to the tyrosine recombinase (phage integrase) family. The members of this family, exemplified by the lambda integrase, do not share extensive amino acid identities, but three invariant residues are found within two regions, designated box I and box II. Two conserved residues are arginines, one located in box I and one in box II, while the other conserved residue is a tyrosine located at the C terminus of box II. We have analyzed the properties of IntI1 variants carrying point mutations at the three conserved residues of the family in in vivo recombination and in vitro substrate binding. We have made four proteins with mutations of the conserved box I arginine (R146) and three mutants with changes of the box II arginine (R280); of these, MBP-IntI1(R146K) and MBP-IntI1(R280K) bind to the attI1 site in vitro, but only MBP-IntI1(R280K) is able to excise cassettes in vivo. However, the efficiency of recombination and DNA binding for MBP-IntI1(R280K) is lower than that obtained with the wild-type MBP-IntI1. We have also made two proteins with mutations of the tyrosine residue (Y312), and both mutant proteins are similar to the wild-type fusion protein in their DNA-binding capacity but are unable to catalyze in vivo recombination.Integrons are DNA elements that capture genes, especially antibiotic resistance genes, by a site-specific recombination system (32). The recombination system consists of a DNA integrase (Int) and two types of recombination sites, attI and attC (59-base element). The integrase gene (int) is located in the 5′ conserved segment of the integron structure (Fig. (Fig.1)1) and is a member of the tyrosine recombinase family (1, 4, 13, 23, 24). Three types of integrases, sharing around 50% identity among themselves, have been identified; they define integron classes 1, 2, and 3 (30). The 5′ conserved segment found in class 1 integrons also contains a promoter region responsible for the expression of inserted cassettes (11, 21) and the recombination site attI1 (31). The 3′ conserved segment of the class 1 integrons includes an ethidium bromide resistance determinant (qacEΔ1), a sulfonamide resistance gene (sulI), an open reading frame (ORF5) of unknown function, and further sequences that differ from one integron to another (5, 6, 28). The 3′ conserved segment of class 2 integrons includes transposition genes (20) while that of class 3 integrons has not yet been studied (2). The variable region, located between the two conserved segments, usually contains antibiotic resistance genes; In0 contains no inserted genes while In21 possesses eight cassettes with ten genes (or ORFs) in this region (5, 16). These genes are part of mobile cassettes which include a recombination site, attC, that differs from one gene to another (18, 33). Incoming genes must be associated with an attC to be recognized by the integron integrase and are preferentially inserted at the recombination site attI1 (11). Cassettes are excised as circular intermediates and integrated at core sites by the action of the integrase (810). The core site, defined as GTTRRRY, makes up the 3′ end of attI1 and attC, with the crossover taking place between the G and the first T (19). Antibiotic selection pressure can reveal cassette rearrangements in which a given resistance is nearest the promoter and thus most strongly expressed (10). Open in a separate windowFIG. 1General structure of class 1 integrons. Cassettes are inserted in the integron variable region by a site-specific recombination mechanism. The attI1 site is shown by a black circle, core sites are represented by ovals, the attC site is indicated by a black rectangle, and promoters are denoted by P. intIl, integrase gene; qacEΔ1, antiseptic resistance gene; sulI, sulfonamide resistance gene; orf5, gene of unknown function.Site-specific recombination, unlike homologous recombination, is characterized by relatively short, specific DNA sequences and requires only limited homology of the recombining partners (12). Site-specific recombination is an entirely conservative process since all DNA strands that are broken (two per exchange site) are rejoined in a process that involves neither ATP nor DNA synthesis. Homology alignments of site-specific recombinases assign them to two families: the resolvase family, named after the TnpR proteins encoded by the transposons γδ and Tn3, and the integrase family. The integrase family includes over 140 members to date, but they are highly diversified proteins (13, 23). Members of this family, which include the well-studied λ integrase, recombine DNA duplexes by executing two consecutive strand breakage and rejoining steps and a topoisomerization of the reactants. The first pair of exchanges form a four-way Holliday junction and the second pair resolve the junction to complete the recombination. The integrase nucleophile is a conserved tyrosine that becomes associated with a phosphate group on DNA. The cleavage sites on each DNA duplex are separated by 6 to 8 bp with a 5′ stagger, and the tyrosine joins to the 3′ phosphate (17).The initial definition of the integrase family was based on comparisons of seven sequences, and three invariant residues were identified: an HXXR cluster and a Y residue (4). Alignment of 28 sequences identified a fourth invariant position, occupied by an arginine residue (1). These four conserved residues are found in two boxes located in the second half of the protein. A recent analysis has shown that the conserved histidine is present in 136 of the 147 members (93%); this residue is then not conserved in all members of the family (13). Another recent analysis has identified three patches of residues located around box I, which seem to be important in the secondary structure of these proteins (23). In this study, we analyzed the properties of several mutants of the conserved residues R146, R280, and Y312 of the integron integrase IntI1 in in vivo recombination and in vitro substrate binding.

Construction of plasmids overexpressing mutant MBP-IntI1 fusion proteins.

The plasmids encoding various mutants of MBP-IntI1 were constructed by PCR using pLQ369 (50 ng) as a template (15). Two primer pairs, designed with the OLIGO software package (version 4.1; National Biosciences, Plymouth, Minn.), were used to construct each set of mutants. The R146 mutants were constructed with an XcmI-BamHI primer pair [IntI1(R146)-XcmI, 5′-TTCACCAGCTTCTGTATGGAACGGGCATG(A/G)(A/T)AATCAG-3′; IntI1(R146)-BamHI, 5′-CCGGATCCCTACCTCTCACT-3′], the R280 mutants were constructed with an NruI-XmnI primer pair [IntI1(R280)-NruI, 5′-AGCCGTCGCGAACGAGTGC(C/T)(C/T)GAGGG-3′; IntI1(R280)-XmnI, 5′-ACCCCTAATGAAGTGGTTCGTATCC-3′], and the Y312 mutants were constructed with a AatII-ScaI primer pair [IntI1(Y312)-AatII, 5′-ATTCCGACGTCTCTACTACGATGATTT(C/T)CACGC-3′; pLQ369-ScaI, 5′-ATGCTTTTCTGTGACTGGTG-3′] (restriction sites within primer sequences are underlined). PCR conditions were 10 min at 94°C, three cycles consisting of 45 s at 94°C, 45 s at 47°C, and 90 s at 72°C, 30 cycles consisting of 45 s at 94°C, 45 s at 60°C (50°C for Y312 mutants), and 90 s at 72°C, and a final elongation step of 10 min at 72°C. The XcmI, NruI, and AatII primers were degenerate in one or two positions, so that a single primer could give all mutants. Mutant PCR fragments were digested and cloned directly into pLQ369 digested with the same enzymes, except for the R146 mutant fragments that were subcloned into pLQ364 at first. New mutant PCR fragments were then amplified on these subclones, using IntI1(R146)-BamHI and IntI1(R280)-XmnI primers. These mutant PCR fragments were cleaved with BamHI and XmnI, and the resulting fragments were cloned into pLQ369. This avoids the necessity of partial digestion of pLQ369 with XcmI. Mutant clones were digested with restriction endonucleases and sequenced to determine the mutation.

In vivo recombination.

Mutant MBP-IntI1 clones were introduced into Escherichia coli TB1 {F′ araΔ(lac-proAB) rpsL (Strr) [φ80dlacΔ(lacZ)M15] hsdR(rKmK)} containing pLQ428 by transformation (Fig. (Fig.22 and Table Table1).1). E. coli TB1 cells containing pLQ428 and one of the MBP-IntI1 mutants were grown at 37°C for 3 h in Luria-Bertani medium. Excision of the aacA1-ORFG and/or ORFH cassettes was induced by the overexpression of the malE-intI1 gene by using 0.3 mM isopropyl-β-d-thiogalactopyranoside (IPTG; Sigma Chemical Co.) and by incubation at 37°C for another 3 h. Cell culture was done in the presence of 50 μg of ampicillin per ml, 15 μg of amikacin per ml, and 50 μg of chloramphenicol per ml. Plasmid DNA was then prepared from 5-ml cultures with the Perfect Prep DNA extraction kit (Mandel Corporation). In order to determine the capacity of mutant MBP-IntI1 proteins to excise aacA1-ORFG and/or ORFH cassettes of In21, we used PCR primers pACYC184-5′ (5′-TGTAGCACCTGAAGTCAGCC-3′) and pACYC184-3′ (5′-ATACCCACGCCGAAACAAG-3′) (Fig. (Fig.2,2, primers 1 and 2) to detect the reduction of pLQ428 length. PCR conditions were 10 min at 94°C, 30 cycles consisting of 1 min at 94°C, 1 min at 60°C, and 5 min at 72°C, and a final elongation step of 10 min at 72°C. A major PCR fragment can be seen in each lane containing a DNA preparation from a mutant clone (Fig. (Fig.3,3, lanes 2 to 9). This band is 2,499 bp long and, as determined by restriction enzyme digestions, represents the pLQ428 clone without any cassette excision (data not shown). This band is also observed in the negative control, which is the pMAL-c2 vector without any gene fused to malE (Fig. (Fig.3,3, lane 12). Open in a separate windowFIG. 2Representation of plasmids used in this study. The positions of the three invariant residues of the integrase family are indicated, along with restriction sites used to construct mutant proteins. Core sites are represented by black circles, and attCs are shown by white boxes. The numbered arrows represent the PCR primers used to detect excision events, pACYC184-5′ (1) and pACYC184-3′ (2). bla, gene encoding β-lactamase; cat, gene encoding chloramphenicol acetyltransferase; intIl, gene encoding the integron integrase (IntI1); malE, gene encoding the maltose binding protein (MBP); ori, origin of replication; Ptac, tac promoter; Ptet, tetracycline promoter. Only relevant restriction sites are indicated.

TABLE 1

Plasmids used in this study
PlasmidCharacteristic(s)aReference or source
pLQ3632,190-bp EcoRI-HincII fragment of pLQ161 cloned in pLQ402 (Apr)16
pLQ3641,027-bp NcoI-BamHI PCR fragment amplified on pLQ860 and cloned in pET-3d (Apr)This study
pLQ3691,019-bp NdeI-BamHI PCR fragment modified to create a blunt-end 5′-ATG and cloned in pMAL-c2 cut with XmnI-BamHI (Apr)15
pLQ376pLQ369 MBP-IntI1(R146K) (Apr)This study
pLQ377pLQ369 MBP-IntI1(R146E) (Apr)This study
pLQ378pLQ369 MBP-IntI1(R146I) (Apr)This study
pLQ379pLQ369 MBP-IntI1(R146V) (Apr)This study
pLQ388pLQ369 MBP-IntI1(R280G) (Apr)This study
pLQ390pLQ369 MBP-IntI1(R280E) (Apr)This study
pLQ391pLQ369 MBP-IntI1(R280K) (Apr)This study
pLQ393pLQ369 MBP-IntI1(Y312S) (Apr)This study
pLQ394pLQ369 MBP-IntI1(Y312F) (Apr)This study
pLQ4282,133-bp EcoRI (filled in)-BglII fragment of pLQ363 cloned in pACYC184 cut with EcoRV-BamHI (Akr Cmr)This study
pLQ8602,900-bp BamHI fragment of pVS1 cloned in pTZ19R (Apr Sulr)5
Open in a separate windowaAkr, Apr, and Cmr, resistance to amikacin, ampicillin, and chloramphenicol. Open in a separate windowFIG. 3Electrophoresis of PCR products obtained with the pACYC184 primer pair and 100 ng of DNA preparations from overexpressed cultures on a 1% agarose gel. Lane 1, 1-kb DNA ladder (Gibco BRL); lane 2, DNA preparation of pLQ428-pLQ377 (R146E); lane 3, pLQ428-pLQ378 (R146I); lane 4, pLQ428-pLQ376 (R146K); lane 5, pLQ428-pLQ379 (R146V); lane 6, pLQ428-pLQ390 (R280E); lane 7, pLQ428-pLQ388 (R280G); lane 8, pLQ428-pLQ391 (R280K); lane 9, pLQ428-pLQ394 (Y312F); lane 10, pLQ428-pLQ393 (Y312S); lane 11, pLQ428-pLQ369 (wild type); lane 12, pLQ428-pMAL-c2 (MBP).The 2,499-bp PCR product was not obtained in the reaction containing the wild-type MBP-IntI1-expressing clone pLQ369 (Fig. (Fig.3,3, lane 11), indicating that there were no remaining full-length pLQ428 molecules. This shows that the wild-type fusion protein is very efficient in site-specific recombination and that all pLQ428 clones have undergone an excision of one or both cassettes. In this PCR, we observed two major bands of 1,341 and 889 bp. The 1,341-bp PCR product was digested with restriction enzymes to show that it represents a pLQ428 clone which has lost the aacA1-ORFG cassette (data not shown). The 889-bp band was also digested with restriction enzymes to show that it represents a pLQ428 clone which has lost both aacA1-ORFG and ORFH cassettes (data not shown). These two PCR products are also observed in the reaction containing the mutant clone pLQ391, which expresses the MBP-IntI1(R280K) fusion protein. This mutant protein is, however, less efficient than the wild-type protein, as seen by the intensity of the PCR products (Fig. (Fig.3,3, lane 8). We were not able to detect a PCR product of 2,047 bp, corresponding to the excision of the ORFH cassette alone; this is not surprising since this event has been shown in another study to be rare (16). It is possible to observe another band in pLQ428-pLQ391 (R280K) and pLQ428-pLQ369 (wild type) PCRs (Fig. (Fig.3,3, lanes 8 and 11); this PCR product is 1,100 bp long and probably represents a recombination event at a secondary site. Restriction enzyme digestions were done on this product, but we were unable to identify its origin. This product results from an event mediated by the integron integrase since it is seen only in reactions containing active proteins. An 1,800-bp PCR band is also present in the negative control and in all PCRs containing a mutant clone. This product appears to be nonspecific, and the fact that it is not seen in the PCR containing the pLQ428-pLQ369 (wild-type) clones probably results from the PCR being more favorable to smaller PCR products.

In vitro substrate binding.

The experiments described above demonstrate that only one of our mutants of IntI1 protein is able to catalyze in vivo recombination. Can all mutant proteins recognize and bind to the IntI1 recombination site in a manner similar to the wild-type protein? To investigate this question, we used purified fusion proteins and a gel retardation assay with the complete attI1 site (5′ site) of the integron. MBP-IntI1 fusion proteins were purified as suggested by New England Biolabs. The concentration of the purified fusion protein was determined by using the Bradford protein assay (Bio-Rad). The protein solution was then made 20% in glycerol and stored at −80°C. The purity of MBP-IntI1 was evaluated as >90% by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (data not shown). Binding reactions were done with labeled 5′-site DNA fragments (20,000 cpm, 0.25 pmol), incubated with different concentrations of MBP-IntI1 in a 10-μl volume containing 10 mM HEPES (K+, pH 8.0), 60 mM KCl, 4 mM MgCl2, 100 μM EDTA (pH 8.0), 100 μg of bovine serum albumin per ml, 250 μM dithiothreitol, 100 ng of poly(dI-dC), and 10% glycerol. Reaction mixtures were incubated at room temperature for 15 min prior to electrophoresis through 4 or 5% prerun, nondenaturing polyacrylamide gels buffered with 0.5× Tris-borate-EDTA. Dried gels were subjected to autoradiography. The wild-type fusion protein and native IntI1 were shown to lead to the same four distinct complexes (I, II, III, and IV) with this DNA substrate (Fig. (Fig.4)4) (15). These complexes represent the binding of four IntI1 molecules to four different sites in the attI1 site (15). Figure Figure44 shows results obtained with nine mutants of the MBP-IntI1 fusion protein. We observed that MBP-IntI1(R146E), MBP-IntI1(R146I), and MBP-IntI1(R146V) lost their ability to bind to the attI1 site, as no complexes are seen in the gel retardation experiment (Fig. (Fig.4A).4A). However, MBP-IntI1(R146K) formed four IntI1-DNA complexes with the 5′ site DNA fragment. The band pattern and the intensity observed with this mutant protein are similar to those observed with the wild-type protein, suggesting that MBP-IntI1(R146K) and MBP-IntI1 bind DNA with similar affinities. Open in a separate windowFIG. 4Binding of mutant MBP-IntI1 fusion proteins purified from E. coli TB1 to the 5′-site DNA fragment containing the complete attI1 site of the In2 integron (from nucleotide −96 to nucleotide +71, relative to the G residue of the core site as position 0). (A) MBP-IntI1(R146) mutants; (B) MBP-IntI1(R280) mutants; (C) MBP-IntI1(Y312) mutants. A purified labeled fragment was incubated with different concentrations of mutant fusion proteins. Free DNA (F) and protein-DNA complexes (I, II, III, and IV) were separated on 4 or 5% polyacrylamide gels and are indicated by arrows. Lanes 1, free DNA; lanes 2 through 7, purified fusion protein at 250, 375, 500, 12.5, 37.5, and 62.5 nM, respectively. The wild-type (WT) lanes in panel C were from a separate gel.Competition with a specific fragment with a 30-fold excess of unlabeled DNA competed away all four complexes, while a 100-fold excess of a nonspecific unlabeled DNA fragment did not compete away any complexes, indicating their specificity (data not shown) (15). We observed that MBP-IntI1(R280G) and MBP-IntI1(R280E) lost their ability to bind the 5′-site DNA fragment, while the MBP-IntI1(R280K) could still bind the attI1 site (Fig. (Fig.4B).4B). However, the band pattern obtained with this mutant protein is weaker than that obtained with the wild-type integrase. For example, at a protein concentration of 250 nM MBP-IntI1(R280K) (lane 2), we observed the formation of complexes I, II, and III, with a stronger intensity for the fastest-migrating complexes, while the intensity of the fourth complex was very weak. At the same concentration of the wild-type protein, we observed the formation of all four complexes, with a stronger intensity for the slowest-migrating complexes and no unbound DNA. These results show that MBP-IntI1(R280K) binds the attI1 site with a lower affinity than the wild-type fusion protein. As shown in Fig. Fig.4C,4C, both MBP-IntI1(Y312F) and MBP-IntI1(Y312S) lead to the formation of four complexes that migrate similarity to those obtained with wild-type MBP-IntI1, as judged by the gel migration of these complexes. The band pattern observed shows that the binding affinity of these mutant proteins is the same as or even better than that of the wild-type protein.

Relationships with other members of the family.

We found that MBP-IntI1 recombinase in which Arg-146 has been changed to lysine [MBP-IntI1(R146K)] by PCR mutagenesis cannot excise cassettes but can bind to the attI1 site with the same efficiency as the wild-type fusion protein. However, MBP-IntI1(R146I), MBP-IntI1(R146E), and MBP-IntI1(R146V) mutant proteins have completely lost both phenotypes. These findings are different from those for other members of the family. The only mutant protein of the lambda integrase at this residue [λ(R212Q)] binds the core site partially and is not able to catalyze in vivo or in vitro recombination (22). Mutants of the Cre recombinase with a change at this residue [Cre(R173K)] bind DNA as well as the wild-type protein but cannot catalyze in vivo or in vitro recombination (1). Mutants of Flp [Flp(R191K) and Flp(R191E)] bind FRT recombination sites as well as the wild-type protein but cannot carry out in vivo or in vitro recombination, except for the Flp(R191K) protein, which has shown slight activity in in vivo recombination (Table (Table2)2) (7, 14, 25). Therefore, the Cre(R173K) and Flp(R146K) mutants have the same phenotype as the MBP-IntI1(R146K) protein. However, the Flp(R191E) mutant protein shows efficient DNA binding while MBP-IntI1(R146E) does not bind to the attI1 site. We interpret these results according to the charge of the Arg-146 residue. The positively charged side chain of this residue makes contact with the DNA, which is negatively charged. This contact is probably important for the good conformation of the protein molecule in positioning the tyrosine residue to perform recombination. When this residue is exchanged for a lysine, DNA contacts are still able to take place because of the charge of the residue, but the side chain is smaller and the lysine is probably not able to position the tyrosine to catalyze recombination. We think that the charge of this residue is very important in the formation of DNA-protein complexes in the integron system, since all other MBP-IntI1 mutants tested are unable to bind DNA. This observation differs from those for Flp, because even when the wild-type residue was replaced by a negatively charged one, it could still bind DNA as well as the wild-type protein (Table (Table2).2).

TABLE 2

Mutational analysis of IntI1 and corresponding residues of other recombinases from the Int family
RecombinaseMutationDNA bindingRecombinationReference(s)
λIntR212QYesaNo22
λIntY342FYesNo22, 26
FlpR191EYesNo7
FlpR191KYesYes7, 14
FlpR308GYesNo27
FlpR308KYesYesa27
FlpY343FYesNo29
FlpY343SYesNo29
CreR173KYesNo1
P2R272KNDbNo23
XerCY275FYesNo3
XerDY279FYesNo3
IntI1R146ENoNoThis study
IntI1R146INoNoThis study
IntI1R146KYesNoThis study
IntI1R146VNoNoThis study
IntI1R280ENoNoThis study
IntI1R280GNoNoThis study
IntI1R280KYesYesaThis study
IntI1Y312FYesNoThis study
IntI1Y312SYesNoThis study
Open in a separate windowaLess efficient than the wild-type protein. bND, not determined. We have also made proteins with mutations at position 280; these were MBP-IntI1(R280E), MBP-IntI1(R280G), and MBP-IntI1(R280K). We found that the MBP-IntI1(R280K) mutant protein binds the attI1 site and excises integron cassettes with a lower efficiency than the wild-type MBP-IntI1, while MBP-IntI1(R280E) and MBP-IntI1(R280G) have completely lost both phenotypes. The Flp(R308K) mutant protein has been shown to bind DNA as well as the wild-type protein, but it recombines DNA with a lower efficiency than wild-type Flp (27). Another mutant protein of Flp [Flp(R308G)] has also been shown to bind DNA as well as the wild-type protein, but it was unable to catalyze in vivo or in vitro recombination (27). These results show that Flp(R308K) and MBP-IntI1(R280K) act similarly but that the other Flp mutant [Flp(R308G)] can bind DNA while the MBP-IntI1 mutant [MBP-IntI1(R280G)] cannot (Table (Table2).2). We also think that the positive charge of this residue is important for the binding of the recombinase to DNA, but Arg-280 does not seem to be implicated in the positioning of the tyrosine residue, since the MBP-IntI1(R280K) mutant protein can perform recombination.We found that MBP-IntI1(Y312S) and MBP-IntI1(Y312F) mutant proteins bind the attI1 site with the same efficiency as the wild-type protein but are not able to catalyze in vivo recombination. As expected, these results are the same as those obtained with the lambda integrase [λ(Y342F)], the XerC and XerD recombinases [XerC(Y275F) and XerD(Y279F)], and the Flp recombinases [Flp(Y343S) and Flp(Y343F)] (Table (Table2)2) (3, 22, 26, 29). The loss of the catalytic activity of the MBP-IntI1(Y312F) mutant protein is not surprising, since the hydroxyl group of the tyrosine, which is responsible for the nucleophilic attack of the DNA at the recombination site, is not present on the phenylalanine residue. The phenotype of MBP-IntI1(Y312S) indicates that the conformation of the tyrosine residue is important for the good activity of the recombinase, because even if the serine residue has a hydroxyl group, it is not able to catalyze recombination. These results indicate that the integron integrase IntI1 uses the hydroxyl group of the conserved tyrosine (Y312) to catalyze site-specific recombination, like other members of the family. However, in vitro recombination using this mutant protein needs to be done to confirm this.These results of point mutations show that mutations of the conserved arginines by nonpositively charged residues abolish substrate recognition, unlike the corresponding mutants of other members of the family. However, further mutational analysis, such as of residues around and in patch III, would be interesting, since only integron integrases contain more residues in this region than other members of the family (23). In vitro recombination assays with purified mutant proteins also need to be done in order to study thoroughly the mechanism of site-specific recombination in integrons.  相似文献   

11.
Multiple roles for cytokinin receptors and cross-talk of signaling pathways     
Teodoro Coba de la Pe?a  Claudia B Cárcamo  M Mercedes Lucas  José J Pueyo 《Plant signaling & behavior》2008,3(10):791-794
  相似文献   

12.
The Chlamydomonas reinhardtii ODA3 Gene Encodes a Protein of the Outer Dynein Arm Docking Complex     
Anthony Koutoulis  Gregory J. Pazour  Curtis G. Wilkerson  Kazuo Inaba  Hong Sheng  Saeko Takada  George B. Witman 《The Journal of cell biology》1997,137(5):1069-1080
  相似文献   

13.
Current Status of Hemostatic Agents and Sealants in Urologic Surgical Practice     
Sashi S Kommu  Robert McArthur  Amr M Emara  Utsav D Reddy  Christopher J Anderson  Neil J Barber  Raj A Persad  Christopher G Eden 《Reviews in urology》2015,17(3):150-159
  相似文献   

14.
Old knowledge and new technologies allow rapid development of model organisms     
Charles E. Cook  Janet Chenevert  Tomas A. Larsson  Detlev Arendt  Evelyn Houliston  Péter Lénárt 《Molecular biology of the cell》2016,27(6):882-887
  相似文献   

15.
Phanerochaete chrysosporium Cellobiohydrolase and Cellobiose Dehydrogenase Transcripts in Wood     
Marcelo A. Vallim  Bernard J. H. Janse  Jill Gaskell  Aline A. Pizzirani-Kleiner  Daniel Cullen 《Applied and environmental microbiology》1998,64(5):1924
  相似文献   

16.
Disease Mutations in the Human Mitochondrial DNA Polymerase Thumb Subdomain Impart Severe Defects in Mitochondrial DNA Replication     
Rajesh Kasiviswanathan  Matthew J. Longley  Sherine S. L. Chan    William C. Copeland 《The Journal of biological chemistry》2009,284(29):19501-19510
Forty-five different point mutations in POLG, the gene encoding the catalytic subunit of the human mitochondrial DNA polymerase (pol γ), cause the early onset mitochondrial DNA depletion disorder, Alpers syndrome. Sequence analysis of the C-terminal polymerase region of pol γ revealed a cluster of four Alpers mutations at highly conserved residues in the thumb subdomain (G848S, c.2542g→a; T851A, c.2551a→g; R852C, c.2554c→t; R853Q, c.2558g→a) and two Alpers mutations at less conserved positions in the adjacent palm subdomain (Q879H, c.2637g→t and T885S, c.2653a→t). Biochemical characterization of purified, recombinant forms of pol γ revealed that Alpers mutations in the thumb subdomain reduced polymerase activity more than 99% relative to the wild-type enzyme, whereas the palm subdomain mutations retained 50–70% wild-type polymerase activity. All six mutant enzymes retained physical and functional interaction with the pol γ accessory subunit (p55), and none of the six mutants exhibited defects in misinsertion fidelity in vitro. However, differential DNA binding by these mutants suggests a possible orientation of the DNA with respect to the polymerase during catalysis. To our knowledge this study represents the first structure-function analysis of the thumb subdomain in pol γ and examines the consequences of mitochondrial disease mutations in this region.As the only DNA polymerase found in animal cell mitochondria, DNA polymerase γ (pol γ)3 bears sole responsibility for DNA synthesis in all replication and repair transactions involving mitochondrial DNA (1, 2). Mammalian cell pol γ is a heterotrimeric complex composed of one catalytic subunit of 140 kDa (p140) and two 55-kDa accessory subunits (p55) that form a dimer (3). The catalytic subunit contains an N-terminal exonuclease domain connected by a linker region to a C-terminal polymerase domain. Whereas the exonuclease domain contains essential motifs I, II, and III for its activity, the polymerase domain comprising the thumb, palm, and finger subdomains contains motifs A, B, and C that are crucial for polymerase activity. The catalytic subunit is a family A DNA polymerase that includes bacterial pol I and T7 DNA polymerase and possesses DNA polymerase, 3′ → 5′ exonuclease, and 5′-deoxyribose phosphate lyase activities (for review, see Refs. 1 and 2). The 55-kDa accessory subunit (p55) confers processive DNA synthesis and tight binding of the pol γ complex to DNA (4, 5).Depletion of mtDNA as well as the accumulation of deletions and point mutations in mtDNA have been observed in several mitochondrial disorders (for review, see Ref. 6). mtDNA depletion syndromes are caused by defects in nuclear genes responsible for replication and maintenance of the mitochondrial genome (7). Mutation of POLG, the gene encoding the catalytic subunit of pol γ, is frequently involved in disorders linked to mutagenesis of mtDNA (8, 9). Presently, more than 150 point mutations in POLG are linked with a wide variety of mitochondrial diseases, including the autosomal dominant (ad) and recessive forms of progressive external ophthalmoplegia (PEO), Alpers syndrome, parkinsonism, ataxia-neuropathy syndromes, and male infertility (tools.niehs.nih.gov/polg) (9).Alpers syndrome, a hepatocerebral mtDNA depletion disorder, and myocerebrohepatopathy are rare heritable autosomal recessive diseases primarily affecting young children (1012). These diseases generally manifest during the first few weeks to years of life, and symptoms gradually develop in a stepwise manner eventually leading to death. Alpers syndrome is characterized by refractory seizures, psychomotor regression, and hepatic failure (11, 12). Mutation of POLG was first linked to Alpers syndrome in 2004 (13), and to date 45 different point mutations in POLG (18 localized to the polymerase domain) are associated with Alpers syndrome (9, 14, 15). However, only two Alpers mutations (A467T and W748S, both in the linker region) have been biochemically characterized (16, 17).During the initial cloning and sequencing of the human, Drosophila, and chicken pol γ genes, we noted a highly conserved region N-terminal to motif A in the polymerase domain that was specific to pol γ (18). This region corresponds to part of the thumb subdomain that tracks DNA into the active site of both Escherichia coli pol I and T7 DNA polymerase (1921). A high concentration of disease mutations, many associated with Alpers syndrome, is found in the thumb subdomain.Here we investigated six mitochondrial disease mutations clustered in the N-terminal portion of the polymerase domain of the enzyme (Fig. 1A). Four mutations (G848S, c.2542g→a; T851A, c.2551a→g; R852C, c.2554c→t; R853Q, c.2558g→a) reside in the thumb subdomain and two (Q879H, c.2637g→t and T885S, c.2653a→t) are located in the palm subdomain. These mutations are associated with Alpers, PEO, mitochondrial myopathy, encephalopathy, lactic acidosis, and stroke-like episodes (MELAS), ataxia-neuropathy syndrome, Leigh syndrome, and myocerebrohepatopathy (POLG mutationDiseaseGeneticsReferenceG848SAlpers syndromeIn trans with A467T, Q497H, T251I-P587L, or W748S-E1143G in Alpers syndrome15, 35, 4350Leigh syndromeIn trans with R232H in Leigh syndrome49MELASIn trans with R627Q in MELAS38PEO with ataxia-neuropathyIn trans with G746S and E1143G in PEO with ataxia50PEOIn trans with T251I and P587L in PEO51, 52T851AAlpers syndromeIn trans with R1047W48, 53In trans with H277CR852CAlpers syndromeIn trans with A467T14, 48, 50In cis with G11D and in trans with W748S-E1143G or A467TAtaxia-neuropathyIn trans with G11D-R627Q15R853QMyocerebrohepatopathyIn trans with T251I-P587L15Q879HAlpers syndrome with valproate-induced hepatic failureIn cis with E1143G and in trans with A467T-T885S35, 54T885SAlpers syndrome with valproate-induced hepatic failureIn cis with A467T and in trans with Q879H-E1143G35, 54Open in a separate windowOpen in a separate windowFIGURE 1.POLG mutations characterized in this study. A, the location of the six mutations characterized is shown in red in the primary sequence of pol γ. Four mutations, the G848S, T851A, R852C, and R853Q, are located in the thumb domain, whereas two mutations, the Q879H and T885S, are in the palm domain of the polymerase region. B, sequence alignment of pol γ from yeast to humans. The amino acids characterized in this study are shown in red. Yellow-highlighted amino acids are highly conserved, and blue-highlighted amino acids are moderately conserved.  相似文献   

17.
Mutational Analysis of the Rous Sarcoma Virus DR Posttranscriptional Control Element     
Robert A. Ogert  Karen L. Beemon 《Journal of virology》1998,72(4):3407
  相似文献   

18.
RNA Polymerase I Transcription Silences Noncoding RNAs at the Ribosomal DNA Locus in Saccharomyces cerevisiae     
Elisa Cesarini  Francesca Romana Mariotti  Francesco Cioci  Giorgio Camilloni 《Eukaryotic cell》2010,9(2):325-335
  相似文献   

19.
Autoregulation of the pTF-FC2 Proteic Poison-Antidote Plasmid Addiction System (pas) Is Essential for Plasmid Stabilization     
Anthony S. G. Smith  Douglas E. Rawlings 《Journal of bacteriology》1998,180(20):5463-5465
  相似文献   

20.
Stress-induced flowering     
Kaede C Wada  Kiyotoshi Takeno 《Plant signaling & behavior》2010,5(8):944-947
Many plant species can be induced to flower by responding to stress factors. The short-day plants Pharbitis nil and Perilla frutescens var. crispa flower under long days in response to the stress of poor nutrition or low-intensity light. Grafting experiments using two varieties of P. nil revealed that a transmissible flowering stimulus is involved in stress-induced flowering. The P. nil and P. frutescens plants that were induced to flower by stress reached anthesis, fruited and produced seeds. These seeds germinated, and the progeny of the stressed plants developed normally. Phenylalanine ammonialyase inhibitors inhibited this stress-induced flowering, and the inhibition was overcome by salicylic acid (SA), suggesting that there is an involvement of SA in stress-induced flowering. PnFT2, a P. nil ortholog of the flowering gene FLOWERING LOCUS T (FT) of Arabidopsis thaliana, was expressed when the P. nil plants were induced to flower under poor-nutrition stress conditions, but expression of PnFT1, another ortholog of FT, was not induced, suggesting that PnFT2 is involved in stress-induced flowering.Key words: flowering, stress, phenylalanine ammonia-lyase, salicylic acid, FLOWERING LOCUS T, Pharbitis nil, Perilla frutescensFlowering in many plant species is regulated by environmental factors, such as night-length in photoperiodic flowering and temperature in vernalization. On the other hand, a short-day (SD) plant such as Pharbitis nil (synonym Ipomoea nil) can be induced to flower under long days (LD) when grown under poor-nutrition, low-temperature or high-intensity light conditions.19 The flowering induced by these conditions is accompanied by an increase in phenylalanine ammonia-lyase (PAL) activity.10 Taken together, these facts suggest that the flowering induced by these conditions might be regulated by a common mechanism. Poor nutrition, low temperature and high-intensity light can be regarded as stress factors, and PAL activity increases under these stress conditions.11 Accordingly, we assumed that such LD flowering in P. nil might be induced by stress. Non-photoperiodic flowering has also been sporadically reported in several plant species other than P. nil, and a review of these studies suggested that most of the factors responsible for flowering could be regarded as stress. Some examples of these factors are summarized in 1214

Table 1

Some cases of stress-induced flowering
Stress factorSpeciesFlowering responseReference
high-intensity lightPharbitis nilinduction5
low-intensity lightLemna paucicostatainduction29
Perilla frutescens var. crispainduction14
ultraviolet CArabidopsis thalianainduction23
droughtDouglas-firinduction30
tropical pasture Legumesinduction31
lemoninduction3235
Ipomoea batataspromotion36
poor nutritionPharbitis nilinduction3, 4, 13
Macroptilium atropurpureumpromotion37
Cyclamen persicumpromotion38
Ipomoea batataspromotion36
Arabidopsis thalianainduction39
poor nitrogenLemna paucicostatainduction40
poor oxygenPharbitis nilinduction41
low temperaturePharbitis nilinduction9, 12
high conc. GA4/7Douglas-firpromotion42
girdlingDouglas-firinduction43
root pruningCitrus sp.induction44
Pharbitis nilinduction45
mechanical stimulationAnanas comosusinduction46
suppression of root elongationPharbitis nilinduction7
Open in a separate window  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