Skip to main content
  • Research
  • Published:

Prediction and overview of the RpoN-regulon in closely related species of the Rhizobiales

Abstract

Background

In the rhizobia, a group of symbiotic Gram-negative soil bacteria, RpoN (σ54, σN, NtrA) is best known as the sigma factor enabling transcription of the nitrogen fixation genes. Recent reports, however, demonstrate the involvement of RpoN in other symbiotic functions, although no large-scale effort has yet been undertaken to unravel the RpoN-regulon in rhizobia. We screened two complete rhizobial genomes (Mesorhizobium loti, Sinorhizobium meliloti) and four symbiotic regions (Rhizobium etli, Rhizobium sp. NGR234, Bradyrhizobium japonicum, M. loti) for the presence of the highly conserved RpoN-binding sites. A comparison was also made with two closely related non-symbiotic members of the Rhizobiales (Agrobacterium tumefaciens, Brucella melitensis).

Results

A highly specific weight-matrix-based screening method was applied to predict members of the RpoN-regulon, which were stored in a highly annotated and manually curated dataset. Possible enhancer-binding proteins (EBPs) controlling the expression of RpoN-dependent genes were predicted with a profile hidden Markov model.

Conclusions

The methodology used to predict RpoN-binding sites proved highly effective as nearly all known RpoN-controlled genes were identified. In addition, many new RpoN-dependent functions were found. The dependency of several of these diverse functions on RpoN seems species-specific. Around 30% of the identified genes are hypothetical. Rhizobia appear to have recruited RpoN for symbiotic processes, whereas the role of RpoN in A. tumefaciens and B. melitensis remains largely to be elucidated. All species screened possess at least one uncharacterized EBP as well as the usual ones. Lastly, RpoN could significantly broaden its working range by direct interfering with the binding of regulatory proteins to the promoter DNA.

Background

The bacterial alternative sigma factor RpoN recognizes and binds a -24/-12-type promoter with the following consensus sequence: 5'-TGGCACG-N4-TTGCW-3' (the bold G and C are situated at position -24 and -12 relative to the transcription start site, respectively) [1]. Subsequently, the core DNA-dependent RNA polymerase (core-RNAP) binds to the RpoN-DNA secondary complex to form a stable closed promoter complex. The closed promoter complex is unable to initiate transcription by itself. For this, melting of the double-stranded (ds) DNA within the closed complex is required [2]. This is accomplished by the nucleotide hydrolysis activity of an activator or enhancer-binding protein (EBP). EBPs bind to enhancer sites situated 100 base-pairs (bp) or more upstream of the transcription initiation site. Each EBP is controlled by its own signal transduction pathway, thereby responding to different conditions [3,4,5]. As there is almost no leaky expression in the absence of EBPs, the expression of different RpoN-dependent genes is tightly regulated by the different EBPs.

RpoN is also known as σ54 (from the 54 kDa molecular weight of the Escherichia coli polypeptide), σN (this sigma factor was initially discovered for its requirement for the expression of nitrogen metabolism genes) and NtrA or GlnF (names now not in common use). Because the members of this protein family vary considerably in molecular weight, the designation σ54 cannot be correctly applied to all of them. Furthermore, we use the name RpoN instead of σN as the link with the encoding gene rpoN is more obvious. As some bacteria (for example, Bradyrhizobium japonicum, Rhizobium etli and Mesorhizobium loti) have two copies of the gene, the respective proteins can be easily distinguished as RpoN1 and RpoN2.

RpoN helps initiating the transcription of genes encoding proteins for very diverse functions in a broad range of bacteria [5,6,7]. The processes controlled by RpoN are not essential for cell survival and growth under favorable conditions, with the exception of Myxococcus xanthus [8]. The most widespread RpoN-regulated function in bacteria is the assimilation of ammonia [6,7]. The expression of genes coding for glutamine synthetase, an ammonium transporter and the PII proteins is controlled by the NtrC EBP and RpoN. In the enteric nitrogen regulation (Ntr) paradigm, NtrC is activated via phosphorylation by NtrB. The PII protein, a signal transducer, stimulates dephosphorylation of NtrC-P under conditions of nitrogen excess (reflected by a high glutamine/2-ketoglutarate ratio), whereas it does not affect the NtrC phosphorylation status under nitrogen-limiting conditions [9]. Consequently, under low nitrogen conditions, NtrC-P activates transcription of the NtrC-RpoN regulon.

Species of the genera Allorhizobium, Azorhizobium, Bradyrhizobium, Mesorhizobium, Rhizobium and Sinorhizobium are generally referred to as rhizobia. Rhizobial rpoN mutants are deficient in symbiotic nitrogen fixation [10,11,12,13,14]. RpoN ensures the transcription of most of the nitrogen fixation genes (nif/fix) whose gene products constitute the nitrogenase complex and accessory proteins [15,16]. However, several other symbiosis-related genes are reported to be RpoN-dependent [17,18,19,20,21,22,23,24,25,26,27]. So far, no large-scale effort has been undertaken to unravel the RpoN-regulon in rhizobia. The recent publication of several rhizobial genomes and symbiotic regions is a good opportunity to identify RpoN-regulated genes in rhizobia. To obtain a good view on what (symbiotic) functions are controlled by RpoN, we carried out an in silico analysis on the presence of -24/-12-type promoters in the symbiotic regions of R. etli CFN42, Rhizobium sp. NGR234, B. japonicum USDA110 and M. loti R7A [28,29,30] and in the genomes of M. loti MAFF303099 and Sinorhizobium meliloti 1021 [31,32]. Two closely related non-symbiotic species belonging to the Rhizobiales order of the α-proteobacteria, namely Agrobacterium tumefaciens C58 and Brucella melitensis 16M [33,34], were also included. These are non-symbiotic plant and animal pathogens, respectively. To date, there is only one report on RpoN-dependent functions in A. tumefaciens [35] whereas no information whatsoever is available on the RpoN-regulon of B. melitensis. The possible RpoN-dependent genes predicted by the screening were complemented with data from the literature and classified according to the function of the encoded proteins.

Results and discussion

In silicoidentification of potential RpoN-dependent promoters

The upstream intergenic sequences were extracted from the genomes of M. loti MAFF303099, S. meliloti 1021, A. tumefaciens C58 and B. melitensis M16 and the symbiotic regions of R. etli CFN42, Rhizobium sp. NGR234, B. japonicum USDA110 and M. loti R7A (see Materials and methods) [28,29,30,31,32,33,34]. The upper strand of these sequences was scored against a weight matrix using PATSER (see Materials and methods). Positive matches (possible RpoN-binding sites or -24/-12-type promoters) were classified according to the functions of the encoded gene products (see Additional data files, pages 8,9). To ensure a sound functional description, the predicted proteins were individually screened against the protein databases of the National Center for Biotechnology Information (NCBI) using BLASTP.

In our analysis, the number of false-positive matches is estimated to be very low. The use of PATSER in combination with a strong weight matrix is a preferred method to identify true binding sites [36]. The weight matrix used here was based on a set of 186 characterized RpoN-binding sites of 44 different bacterial species (Table 1) [1]. The lower threshold for the scores was chosen such that only matches strongly resembling the consensus were retained (see Materials and methods). This high stringency allows for a high specificity at the expense of the sensitivity. Indeed, some reports mention the presence of very poorly conserved -24/-12-type promoters that appear to be functional to some degree (see Additional data files, pages 8,9). Our stringent procedure most probably misses these sites. Therefore, the retained matches might represent an underestimation of the actual number of active -24/-12-type promoters. Second, only intergenic sequences were considered for the screening, as all known -24/-12-type promoters are situated in intergenic regions [7]. Moreover, the matches all have the correct orientation, as the upper strand of the intergenic sequences was used (for matches on the lower strand, see 'Additional control mechanisms involving RpoN'). The median of the distances from the -12 (C) position - the cytidylate residue on position -12 relative to the transcription initiation site - of a match to the start codon of the downstream coding sequence (CDS) varies from 74 to 156 bp (Table 2). This is somewhat different from the situation in E. coli, where the average distance amounts to 50 bp [7]. Although roughly 75% of the matches are within 200 bp upstream of the start codon, matches with distances over 1,000 bp were also retained (Table 2). A good example to justify this is the case of the mapped promoter of the B. japonicum gene fixB, which is situated 720 bp upstream of the coding region [37]. As is the case with other in silico prediction methods, the results of our analysis may have been biased by the approach used.

Table 1 Weight matrix based on 186 characterized -24/-12-type promoters [1]
Table 2 Comparison of different members of the Rhizobiales

A good test case to evaluate the reliability of our predictions is to compare the matches with experimental data. In our laboratory we confirmed the RpoN-dependent expression of several R. etli genes with predicted RpoN-binding sites (see Additional data files, pages 3,4,6,7,12,13). The products of these genes are involved in a wide variety of functions, such as nitrogen fixation (nifH and iscN-nifUS) [38,39], oxidative stress and gene regulation (spxA-rpoN2) [13,40], and transport (yp104-103 and yp100) [41] (G. Dirix, M. Van Guyse and J.M., unpublished results). For the genes on plasmid pNGR234a of Rhizobium sp. NGR234, an extensive expression analysis is available as well as an annotation of RpoN-dependent promoters [42]. Our screening confirmed the 15 previously annotated RpoN-dependent promoters on pNGR234a (controlling 50 CDSs). All these promoters are upregulated in bacteroids, resembling the NifA-RpoN-dependent symbiosis-specific expression pattern of the nitrogen fixation genes (see Additional data files). In addition to these promoters, eight other matches were found, controlling ten CDSs. Five out of these eight are highly expressed in bacteroids (see Additional data files), indicating that they are probably functional. In the symbiotic region of B. japonicum, 31 matches were found (controlling 87 CDSs) (Table 2), which were also previously annotated as RpoN-dependent promoters [29]. Thirteen of these promoters were shown to be functional and controlled by NifA and RpoN (see Additional data files). Furthermore, the fact that most - if not all - common RpoN-controlled genes were picked up for all species screened, including those already characterized in rhizobia, gives additional support to the methodology used. Common RpoN-dependent functions include assimilation of ammonium, transport of C4-dicarboxylates and nitrogen fixation (see Additional data files, pages 1-7, 10).

Functional classification of RpoN-dependent genes in the Rhizobiales

The matches (also referred to as possible RpoN-binding sites or possible -24/-12-type promoters) obtained from the in silico search were complemented with literature data on RpoN-regulated genes in rhizobia, and classified according to protein function (see Additional data files).

For the four genomes screened, matches were found upstream of the genes whose products are involved in the uptake and assimilation of ammonia: glnK-amtB, gInBA and glnII (see Additional data files, pages 1,2). A similar regulation and gene organization is found in other α-proteobacteria [9]. B. melitensis differs slightly in that amtB constitutes a monocistronic unit and the glnII ortholog was not identified.

Possible RpoN-binding sites were found upstream of many nitrogen fixation genes in the six symbiotic, nitrogen-fixing species (R. etli, Rhizobium sp. NGR234, B. japonicum, M. loti MAF303099 and R7A and S. meliloti). Most rhizobial nitrogen fixation genes are controlled by RpoN and the NifA EBP (see Additional data files, pages 2-7). The latter protein is activated under microaerobic conditions and is highly active in bacteroids [15,16].

A possible RpoN-binding site is also present in the promoter region of dctA, encoding a C4-dicarboxylate transporter, in all species screened, except for B. melitensis, where no such gene is present, and R. etli, where it is not present on the symbiotic plasmid (see Additional data files, page 10). In M. loti MAFF303099, -24/-12-type promoters were identified in front of the two C4-dicarboxylate transporter genes. Rhizobial dctA mutants are deficient in nitrogen fixation. The expression of dctA is controlled under free-living conditions by RpoN-DctD, whereas in bacteroids, the DctD EBP is (partially) replaced by another unknown EBP. The DctD activator protein is activated by the DctB sensor protein that senses the presence of C4-dicarboxylates together with DctA [16].

During symbiotic nitrogen fixation, hydrogen is produced as an obligate byproduct of the nitrogenase reaction. Some rhizobia possess a hydrogen-uptake (Hup) system that reoxidizes the hydrogen, a process that is thought to increase the efficiency of nitrogen fixation. This is, for instance, the case in B. japonicum, where the -24/-12-type promoters controlling these genes were confirmed in the screening (see Additional data files, pages 7,8). In B. japonicum, HoxA-RpoN is required for hupSL expression under free-living conditions and FixK2 under symbiotic conditions, whereas in Rhizobium leguminosarum, expression of hup genes is only induced in bacteroids and depends on NifA-RpoN [21,43,44].

As well as these previously well-characterized RpoN-dependent functions, several other groups of genes were found in the in silico screening. Potential target genes preceded by a possible -24/-12-type promoter belong to the following functional categories: nodulation; transport; detoxification; gene regulation; cell envelope; amino-acid biosynthesis; energy and DNA metabolism; translation; integration and recombination; and various others (see Additional data files, pages 8-28).

An interesting observation is the large number of hypothetical CDSs preceded by a possible RpoN-binding site (see Additional data files, pages 21-28). Of the 593 gene products identified in the screening, 22% have no function attributed, but do have homologs in other species. Proteins encoded by orphan genes, unique to one species, make up 9% of the total. Some of these hypothetical CDSs having a possible -24/-12-type promoter were picked up in several genomes: atu1754-smc01000-mll0186, smc00999-mll0185, mll9007-y4nG, mll9408-y4nH, orf196-y4vH, orf538-y4vI, orf354-y4vJ, id79-mlr5912-y4vQ-msi288-yp001, yp003-y4wP-yp021-yp099, mlr5819-y4wI, mll5852-msi351, mlr5886-msi324, mlr5914-msi286 and msr5928-msi276. The majority of these hypothetical genes in Rhizobium sp. NGR234 are highly expressed in bacteroids [42], suggesting a possible involvement of the putative gene products in the symbiosis.

Another interesting class of possible RpoN-regulated genes encodes proteins catalyzing detoxification reactions (see Additional data files, pages 12,13). During symbiosis, rhizobia are confronted with toxic compounds and oxidative stress in the form of reactive oxygen species [45]. Effective protection against these harmful agents would thus enhance the symbiotic performance of the rhizobia. Two good candidate proteins to fulfill such a function were identified in this screening: a peroxiredoxin and a cytochrome P450. Peroxiredoxins reduce peroxides to the alcohol and water. The gene encoding the symbiosis-specific peroxiredoxin SpxA is highly expressed in bacteroids of several rhizobial species in a NifA-RpoN-dependent way [13,40,42]. Cytochrome P450S catalyze the monoxygenation of a broad spectrum of substrates, including toxic compounds. Several genes encoding cytochrome P450S have a possible RpoN-binding site in their promoter region. One subclass of these P450S is conserved among several rhizobial species and is highly expressed in bacteroids [27,42]. Expression of the Rhizobium sp. BR816 homolog depends on NifA-RpoN and another, so far unidentified, factor [27]. Furthermore, two matches were found upstream of B. japonicum nrgA and nrgBC. The expression of these genes, possibly involved in detoxification reactions, was shown to be controlled by NifA and RpoN [46].

Transport systems have an important role at different stages of the rhizobium-legume symbiosis [41,47,48,49]. Many transporter-encoding gene loci are preceded by a possible -24/-12-type promoter (see Additional data files, pages 9-12). The majority of these transporters are of the ABC type, which is consistent with the fact that this type is the most abundant transporter in Rhizobiaceae family [33]. The RpoN-regulated transporters might be involved in symbiotic (as, for instance, DctA) or non-symbiotic (in A. tumefaciens) processes.

Many regulatory genes have a possible RpoN-binding site in their promoter region (see Additional data files, pages 13,14). The RpoN regulon is known to expand its working range by controlling the expression of other regulatory genes [6,7]. The induction of the y4qH gene (preceded by a possible RpoN-binding site and encoding a LuxR-type transcriptional regulator) in bacteroids of Rhizobium sp. NGR234 illustrates how RpoN might indirectly control the expression of genes involved in symbiosis [42]. Other examples include RpoN dependency of R. leguminosarum fnrN, or the autoregulation of R. etli rpoN2 [13,14]. The latter form of control also appears to be conserved in M. loti MAFF303099 and R7A, where a possible -24/-12-type promoter was identified upstream of the second copy of rpoN (see Additional data files, page 7).

The gene products of the rhizobial nodulation genes (nol, nod, noe) synthesize and export a class of signal molecules, the so-called Nod factors, involved in the early steps of the symbiotic interaction with leguminous plants. Although the transcription of most nodulation genes depends on the NodD regulator protein in the presence of a suitable plant-root-derived flavonoid, several reports mention an RpoN dependency (to some extent) of some of the nodulation genes ([22,23,24,26,50] and see Additional data files, pages 8,9). However, the screening did not reveal any matches upstream of rhizobial nodulation genes. Because of the high stringency of the screening, several promoters, such as those of Rhizobium sp. NGR234 nodD2 and S. meliloti nodD3, may have been missed. It thus seems that less conserved -24/-12-type promoters are functional to some degree. On the other hand, such poorly conserved promoters may be the exception rather than the rule. Nevertheless, the -24/-12-type promoter of the B. japonicum nfeC gene (involved in nodule formation) was confirmed in the screening.

The other functional categories (see Additional data files, page 14-21) include, among others, cell envelope, amino-acid biosynthesis, energy metabolism, DNA metabolism, translation, and integration and recombination. Unlike the functions discussed above, these categories comprise many different proteins from different bacterial species. It appears that the RpoN dependency of the encoding genes in these categories is not conserved in the Rhizobiales. This might reflect a species-specific recruitment of RpoN for gene transcription.

Signals controlling RpoN-dependent gene expression in the Rhizobiales

What signals control the activity of EBPs in species of the Rhizobiales? An in silico screening for EBPs in the proteomes of members of the Rhizobiales (see Materials and methods) revealed that the number of candidate EBPs is proportional to the genome size (Table 2). M. loti, S. meliloti, A. tumefaciens and B. melitensis all have roughly 1 EBP per million bases (Mb) (8, 7, 6 and 3, respectively), a number differing from a previous estimate (13, 12 and 9, respectively; the B. melitensis genome was not published at the time) [33]. However, our predicted EBPs scored significantly better (E-value 10-25 or less) with the Pfam Sigma54_activat domain HMM motif than the best of the rejected proteins (E-value 10-5 or greater), giving confidence to our prediction. Each of the candidate EBPs was subjected to a BLASTP search (default settings) against the NCBI protein databases and classified according to homology (Table 3). A similar analysis of Pseudomonas aeruginosa, Salmonella typhimurium and E. coli proteomes revealed that these bacteria have a higher number of possible EBPs (22, 15 and 13, respectively) than the species of the Rhizobiales (data not shown).

Table 3 Enhancer binding proteins (EBPs) of different members of the Rhizobiales

All four genomes screened have NtrC and NtrX orthologs (Table 3), which are both nitrogen-responsive EBPs [9]. Rhizobial NtrC-RpoN-regulated genes are downregulated early in bacteroid differentiation [51,52]. This downregulation seems to be necessary for an effective symbiosis as ectopic expression of R. etli amtB alters the bacteroid differentiation process and the ability to enter the host cells [53].

M. loti, S. meliloti and A. tumefaciens all have two DctD EBP paralogs (Table 3). DctD is activated by the DctB sensor protein, which senses C4-dicarboxylates in collaboration with the DctA C4-dicarboxylate transporter. The second dctD copy might encode the so-called alternative symbiotic activator, ensuring dctA expression in S. meliloti and R. leguminosarum dctD mutants during symbiosis [16].

All nitrogen-fixing species possess a NifA homolog - M. loti MAFF303099 and R7A even have two (Table 3). Rhizobial NifA proteins are activated by the low oxygen tension in the nodules, which is the key signal controlling expression of the nitrogen fixation genes and a prerequisite for nitrogen fixation [54]. NifA activates RpoN-dependent transcription of many rhizobial genes, including nitrogen fixation genes (see Additional data files). For the NifA-RpoN-dependent expression of some genes, microaerobiosis alone is not sufficient to equal the expression levels observed in bacteroids (see Additional data files, pages 2,4,6,7,13,24) [27,39]. An additional signal appears to control the optimal functioning of NifA.

The presence of additional uncharacterized EBPs suggests that other signals control RpoN-dependent gene expression in the Rhizobiales (Table 3). These signals could control expression of genes involved in symbiosis (M. loti and S. meliloti) or other, so far uncharacterized, functions (A. tumefaciens and B. melitensis). One common signal might control the activity of EBP-1, present in the four genomes (Table 3).

Comparison between symbiotic and non-symbiotic Rhizobiales

Globally, a higher number of possible RpoN-binding sites per megabases of genome is observed in symbiotic members of the Rhizobiales than in non-symbiotic ones (Table 2). This ratio is even higher in the symbiotic regions and that might reflect the recruitment of NifA-RpoN-dependent promoters, which are highly upregulated in bacteroids, for late symbiotic functions. RpoN is not essential for bacterial survival, but controls genes whose products have a wide range of different functions [6]. Apparently, there has been a selection in bacteroids for the NifA-RpoN-type promoter, not only for nitrogen fixation but for other symbiotic functions as well. This is illustrated by the high number of non-nif/fix genes on pNGR234a preceded by a -24/-12-type promoter and preferentially expressed in bacteroids (see Additional data files) [42]. The symbiotic island of M. loti R7A is about 109 kb smaller than that of MAFF303099 and the two areas share only a conserved backbone of 248 kb [30]. This distinction seems not to be reflected by the RpoN-regulon, as the majority of possible RpoN-dependent loci (75%) on the two islands is conserved (see Additional data files).

Although an A. tumefaciens rpoN mutant is not affected in tumorigenesis, expression of vir genes, chemotaxis or flagellum formation, it does not grow on nitrate as sole nitrogen source nor on C4-dicarboxylates as sole carbon source, and grows only very poorly on arginine as sole nitrogen source [35]. We were able to identify possible RpoN-binding sites in the promoter regions of the dctA gene, the gInK-amtB, glnBA, glnII and nrtABC-nirBD loci (see Additional data files, pages 1,2,10,12,16), providing genetic evidence for the previously observed phenotypes. As could be expected, no virulence, tumorigenesis, chemotaxis or flagellar genes were picked up in screening, although a good match was found upstream of motB, which encodes a flagellar motor protein. Strikingly, in Agrobacterium the total number of gene products whose genes are preceded by a possible -24/-12-type promoter is approximately equal to that of Sinorhizobium (Table 2). S. meliloti and A. tumefaciens are closely related at the level of protein identity as well as nucleotide colinearity and conservation of gene order [33]. This relatedness appears to be somewhat conserved at the regulatory level, as S. meliloti and A. tumefaciens share seven predicted RpoN-binding sites upstream of the same gene loci: dctA, glnK-amtB, glnBA, glnll, nrtABC, atu1754-smc01000 and atu4739-smb20103 (see Additional data files, pages 1,2,10,12,18,21,24). The relatively high number of positive matches in A. tumefaciens and the lack of tumorigenesis-related phenotype in the rpoN mutant hints to other, so far unidentified, RpoN-controlled functions. Indicative of this is the number (22) of possible RpoN-binding sites upstream of hypothetical genes in A. tumefaciens (see Additional data files, pages 21,26). Of the predicted proteins of A. tumefaciens, 16% are not present in M. loti or S. meliloti and 89% of these are hypothetical gene products [33], leaving room for genetic diversity between rhizobia and A. tumefaciens.

B. melitensis is an intracellular pathogen of humans and animals, and is closely related to rhizobia and A. tumefaciens [55]. Only 12 matches resulted from the in silico analysis for possible RpoN-binding sites (Table 2), two of which were found in the upstream region of amtB and glnBA (see Additional data files, pages 1,2). At present, no report of a B. melitensis rpoN mutant is available. A good candidate function for RpoN to regulate in B. melitensis would be pathogenesis, as is the case for several other pathogens [6]. Six possible -24/-12-type promoters were found in the B. melitensis genome upstream of hypothetical CDSs (see Additional data files, pages 21,22,26).

E. coli was estimated to have around 30 -24/-12-type promoters, controlling mainly functions related to nitrogen metabolism [7]. Although there is some overlap of RpoN-controlled functions between E. coli and members of the Rhizobiales (for example, ammonium assimilation and hydrogen oxidation), RpoN seems to control a more diverse set of functions in the latter. This appears to somewhat contradict the observation that E. coli possesses a higher number of EBPs (see 'Signals controlling RpoN-dependent gene expression in the Rhizobiales').

Additional control mechanisms involving RpoN

Expression studies of rpoN in wild-type and rpoN- backgrounds revealed that RpoN negatively controls its own expression in B. japonicum, Acinetobacter calcoaceticus, R. etli and R. leguminosarum bv. viciae [14,56,57,58]. The presence of a putative oppositely oriented RpoN-binding site on the template strand of rpoN promoter regions, overlapping with the -10 promoter region and the transcription start site, was proposed [58]. It was stated that, as RpoN is able to bind to -24/-12-type promoters in the absence of core-RNAP, the negative autoregulation of the rpoN genes occurs by direct interference of RpoN with the binding of σ70-holo-RNAP complex to the -10 promoter region of the rpoN gene. Site-directed mutagenesis of the highly conserved GG to TT in the putative RpoN-binding site of the rpoN promoter relieved the negative autoregulation, giving strong support to the above hypothesis [14].

The in silico screening revealed the presence of oppositely oriented possible RpoN-binding sites upstream of the rpoN genes of M. loti, S. meliloti, A. tumefaciens and B. melitensis (Table 2). A comparison with the rpoN coding sequences of R. etli, R. leguminosarum, Rhizobium sp. NGR234, B. japonicum, M. loti and S. meliloti revealed that the rpoN genes of A. tumefaciens and B. melitensis were incorrectly annotated. Their coding sequences should be 63 bp longer and shorter, respectively. An alignment of the rpoN promoter regions of these species shows the strong conservation of these promoters (Figure 1). In addition, the screening of the lower strand of the intergenic sequences revealed a slightly lower number of matches than that of the correctly oriented RpoN-binding sites (Table 2). It is thus not unconceivable that RpoN could alter the expression of these genes in a way similar to its own autoregulation, that is, by interference with the binding of the holo-RNAP or a regulatory protein to the promoter. This would significantly broaden the working domain of RpoN.

Figure 1
figure 1

Manual alignment of rpoN promoter sequences. At (Agrobacterium tumefaciens, GI: 17738659); Bj (Bradyrhizobium japonicum, GI: 152137); Bm (Brucella melitensis, GI: 17983821); Ml (Mesorhizobium loti, GI: 14023393); NGR (Rhizobium sp. NGR234, GI: 152431); Re (Rhizobium etli, GI: 1046228), Rl (Rhizobium leguminosarum bv. viciae, GI: 5759116), Sm (Sinorhizobium meliloti, GI: 152389). Upper line, -35 and -10 consensus sequences of Escherichia coli and the transcription start site (*) as determined for S. meliloti [69]. Lower line, consensus sequence of -24/-12-type promoter [1]. Nucleotides are shaded in black (100% conserved) or gray (75% conserved). The numbers represent the distance (in bp) from the end of the alignment to the start codon of the downstream rpoN gene.

Conclusions

A highly specific in silico screening method was applied to predict members of the RpoN-regulon in eight different species of the Rhizobiales. The matches obtained were individually checked and classified according to protein function, resulting in a highly annotated and manually curated dataset. This dataset was complemented with available literature data on members of the RpoN-regulon in Rhizobiales. In addition, a screening was carried out to identify possible EBPs controlling the expression of RpoN-dependent genes. Together, these data serve as a source of exhaustive information on the (possible) roles of RpoN in symbiotic and non-symbiotic processes.

RpoN-binding sites were found upstream of genes involved in common RpoN-dependent functions, such as assimilation of ammonium and uptake of C4-dicarboxylic acids. The symbiotic members of the Rhizobiales seem to have recruited RpoN for the expression of nitrogen fixation and other symbiotic genes. This is illustrated by the high number of possible RpoN-binding sites in the symbiotic regions of these bacteria. Other RpoN-dependent symbiotic functions might include detoxification and transport or might be controlled indirectly through other regulatory proteins. Whereas an A. tumefaciens rpoN mutant only displays common RpoN-dependent phenotypes, the relatively high number of possible RpoN-binding sites present in its genome points to several other, yet unidentified, RpoN-dependent functions. So far, no reports are present on RpoN-dependent phenotypes in B. melitensis. This animal pathogen has a significantly lower number of possible RpoN-binding sites than the other members of the Rhizobiales. B. melitensis RpoN might be required for infection of the host organism, as is the case for other pathogens. Furthermore, the species screened seem to have recruited RpoN independently, in a species-specific manner, for the transcription of different gene sets. The high percentage of hypothetical conserved and non-conserved CDSs preceded by a possible RpoN-binding site opens up ample opportunities for future research. Several uncharacterized EBPs were identified besides the 'classic' EBPs such as NtrC, NtrX, NifA and DctD. This implies that signals other than nitrogen, oxygen and C4-dicarboxylates control the expression of RpoN-dependent genes in species of the Rhizobiales. Identification of these signals will give better insight into yet uncharacterized RpoN-dependent functions. Finally, a similar number of possible RpoN-binding sites were found on the lower strand of the upstream intergenic sequences. RpoN might thus significantly extend its working domain by blocking the binding of transcription regulatory factors. Such is the case, for instance, in the negative autoregulation of rpoN.

Although much consideration was given in our analysis to the design and, to some extent, the experimental validation of the approach, experimental confirmation will ultimately be required to establish the biological meaning of the predicted -24/-12-type promoters, as is the case with all computer predictions.

In conclusion, a highly efficient method was applied to predict the RpoN-regulon of different members of the Rhizobiales group. The same approach might be used for the prediction of RpoN-dependent genes in other bacterial species.

Materials and methods

Retrieval of intergenic sequences

Complete DNA sequences from A. tumefaciens C58 (circular chromosome: NC_003304, linear chromosome NC_003305, plasmid AT: NC_003306, plasmid Ti: NC_003308), B. japonicum USDA110 (symbiotic chromosomal region: AF322012 and AF322013), B. melitensis 16 M (chromosome I: NC_003317, chromosome II: NC_003318), M. loti MAFF303099 (chromosome: NC_002678, plasmid pMla: NC_002679, plasmid pMlb: NC_002682), M. loti R7A (symbiotic island: AL672111), Rhizobium sp. NGR234 (plasmid pNGR234a: NC_000914), R. etli CFN42 (plasmid p42d: NC_004041) and S. meliloti 1021 (chromosome: NC_003047, plasmid pSymA: NC_003037, plasmid pSymB: NC_003078) were extracted from GenBank. Intergenic sequences were identified by automatically parsing the corresponding GenBank files using the modules of INCLUsive [59,60]. An intergenic region is defined as the non-coding region between two genes. Intergenic regions smaller than 10 nucleotides were discarded, as the corresponding genes are likely to belong to an operon.

Prediction of possible RpoN-binding sites

The intergenic regions were screened with the PATSER module of the Regulatory Sequence Analysis Tools (RSAT) [61,62,63] for the presence of the -24/-12 promoter consensus sequence. PATSER scores N-mers (in this case 16-mers) from a sequence against a given weight matrix. A set of 186 RpoN-dependent promoters from different bacterial species [1] was used to generate the weight matrix (Table 1). Initially, this matrix was trained against 67 -24/-12-type promoters with a known transcriptional start site. From the distribution of these scores (Figure 2), it was decided to retain all matches with a score higher than or equal to the fifth percentile (8.9). PATSER was run with the GC content of the intergenic sequences as a measure for the a priori probabilities of the nucleotides. The intergenic GC content differs markedly from the total GC content of the genomes (Table 2).

Figure 2
figure 2

Distribution of PATSER scores (see Material and methods) of 67 -24/-12-type promoters with mapped transcriptional start sites. (GenBank GI number: 2979503, 141885, 141892, 38664, 38679, 1769418, 142336, 142378, 142326, 39977, 550310, 408911, 39532, 39516, 39526, 152106, 152283, 152315, 39548, 312974, 12620419, 152100, 152280, 152317, 2316081, 144194, 896457, 144223, 262651, 3493239, 7208421, 262651, 145911, 556890, 41774, 146158, 1004098, 41568, 42538, 149241, 43802, 149252, 149256, 149246, 43857, 43864, 149273, 149275, 150095, 950651, 150993, 490170, 6492432, 151643, 6636054, 46254, 152305, 152230, 340664, 46285, 46324, 46324, 550144, 550144, 664946, 453435, 1649033). Cumulative percentage: black line.

Prediction of possible EBPs

An estimate of the number of possible EBPs in the respective proteomes was obtained by looking for the presence of the Pfam Sigma54_activat domain [64]. Sigma54_activat is a conserved domain present in EBPs that is involved in the ATP-dependent interaction with RpoN. The predicted proteome sequences of A. tumefaciens, B. melitensis, S. meliloti and M. loti MAFF303099 were downloaded from Proteome Analysis @ EBI [65] and the protein sequences of the symbiotic regions of B. japonicum, M. loti R7A, Rhizobium sp. NGR234 and R. etli were obtained from the NCBI protein database [66]. The protein sequences were subsequently queried with the PF00158 HMM motif using HMMER 2.2 g [67]. Matches with an E-value less than or equal to 10-25 were retained.

Additional data files

An additional data file listing all genes and proteins included in this analysis is available. Each protein is accompanied by a functional description, the species from which it comes, GenBank GI number and transcription unit; additional references for each protein are also included in the file. Information on the gene's regulation is provided where available. The data were obtained from an in silico analysis (see Materials and methods) and the literature [12,13,14,17,18,19,20,21,22,23,24,25,26,27,29,30,31,32,3334,39,40,41,42,43,44,46,50,51,56,58,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88].

References

  1. Barrios H, Valderrama B, Morett E: Compilation and analysis of sigma 54-dependent promoter sequences. Nucleic Acids Res. 1999, 27: 4305-4313. 10.1093/nar/27.22.4305.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  2. Buck M, Gallegos MT, Studholme DJ, Guo YL, Gralla JD: The bacterial enhancer-dependent sigma(54) (sigma(N)) transcription factor. J Bacteriol. 2000, 182: 4129-4136. 10.1128/JB.182.15.4129-4136.2000.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  3. Kustu S, Santero E, Keener J, Popham D, Weiss D: Expression of sigma 54 (ntrA)-dependent genes is probably united by a common mechanism. Microbiol Rev. 1989, 53: 367-376.

    PubMed  CAS  PubMed Central  Google Scholar 

  4. Morett E, Segovia L: The sigma 54 bacterial enhancer-binding protein family: mechanism of action and phylogenetic relationship of their functional domains. J Bacteriol. 1993, 175: 6067-6074.

    PubMed  CAS  PubMed Central  Google Scholar 

  5. Shingler V: Signal sensing by sigma 54-dependent regulators: derepression as a control mechanism. Mol Microbiol. 1996, 19: 409-416. 10.1046/j.1365-2958.1996.388920.x.

    Article  PubMed  CAS  Google Scholar 

  6. Studholme DJ, Buck M: The biology of enhancer-dependent transcriptional regulation in bacteria: insights from genome sequences. FEMS Microbiol Lett. 2000, 186: 1-9. 10.1016/S0378-1097(00)00082-3.

    Article  PubMed  CAS  Google Scholar 

  7. Reitzer L, Schneider BL: Metabolic context and possible physiological themes of sigma(54)-dependent genes in Escherichia coli. Microbiol Mol Biol Rev. 2001, 65: 422-444. 10.1128/MMBR.65.3.422-444.2001.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  8. Keseler IM, Kaiser D: Sigma(54), a vital protein for Myxococcus xanthus. Proc Natl Acad Sci USA. 1997, 94: 1979-1984. 10.1073/pnas.94.5.1979.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  9. Arcondéguy T, Jack R, Merrick M: P-II signal transduction proteins, pivotal players in microbial nitrogen control. Microbiol Mol Biol Rev. 2001, 65: 80-105. 10.1128/MMBR.65.1.80-105.2001.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Ronson CW, Nixon BT, Albright LM, Ausubel FM: Rhizobium meliloti ntrA (rpoN) gene is required for diverse metabolic functions. J Bacteriol. 1987, 169: 2424-2431.

    PubMed  CAS  PubMed Central  Google Scholar 

  11. van Slooten JC, Cervantes E, Broughton WJ, Wong CH, Stanley J: Sequence and analysis of the rpoN sigma factor gene of Rhizobium sp. strain NGR234, a primary coregulator of symbiosis. J Bacteriol. 1990, 172: 5563-5574.

    PubMed  CAS  PubMed Central  Google Scholar 

  12. Stigter J, Schneider M, de Bruijn FJ: Azorhizobium caulinodans nitrogen fixation (nif/fix) gene regulation: mutagenesis of the nifA -24/-12 promoter element, characterization of a ntrA(rpoN) gene, and derivation of a model. Mol Plant-Microbe Interact. 1993, 6: 238-252.

    Article  PubMed  CAS  Google Scholar 

  13. Michiels J, Moris M, Dombrecht B, Verreth C, Vanderleyden J: Differential regulation of Rhizobium etli rpoN2 gene expression during symbiosis and free-living growth. J Bacteriol. 1998, 180: 3620-3628.

    PubMed  CAS  PubMed Central  Google Scholar 

  14. Clark SR, Oresnik IJ, Hynes MF: RpoN of Rhizobium leguminosarum bv. viciae strain VF39SM plays a central role in FnrN-dependent microaerobic regulation of genes involved in nitrogen fixation. Mol Gen Genet. 2001, 264: 623-633. 10.1007/s004380000348.

    Article  PubMed  CAS  Google Scholar 

  15. Fischer HM: Genetic regulation of nitrogen fixation in rhizobia. Microbiol Rev. 1994, 58: 352-386.

    PubMed  CAS  PubMed Central  Google Scholar 

  16. Kaminski PA, Batut J, Boistard P: A survey of symbiotic nitrogen fixation by rhizobia. In The Rhizobiaceae, Molecular Biology of Model Plant-Associated Bacteria. Edited by: Spaink HP, Kondorosi A, Hooykaas PJJ. 1998, Dordrecht: Kluwer, 431-460.

    Google Scholar 

  17. Hawkins FK, Johnston AW: Transcription of a Rhizobium leguminosarum biovar phaseoli gene needed for melanin synthesis is activated by NifA of Rhizobium and Klebsiella pneumoniae. Mol Microbiol. 1988, 2: 331-337.

    Article  PubMed  CAS  Google Scholar 

  18. Soto MJ, Zorzano A, Mercado-Blanco J, Lepek V, Olivares J, Toro N: Nucleotide sequence and characterization of Rhizobium meliloti nodulation competitiveness genes nfe. J Mol Biol. 1993, 229: 570-576. 10.1006/jmbi.1993.1060.

    Article  PubMed  CAS  Google Scholar 

  19. Chun JY, Sexton GL, Roth LE, Stacey G: Identification and characterization of a novel Bradyrhizobium japonicum gene involved in host-specific nitrogen fixation. J Bacteriol. 1994, 176: 6717-6729.

    PubMed  CAS  PubMed Central  Google Scholar 

  20. Chun JY, Stacey G: A Bradyrhizobium japonicum gene essential for nodulation competitiveness is differentially regulated from two promoters. Mol Plant Microbe Interact. 1994, 7: 248-255.

    Article  PubMed  CAS  Google Scholar 

  21. Brito B, Martinez M, Fernandez D, Rey L, Cabrera E, Palacios JM, Imperial J, Ruiz-Argueso T: Hydrogenase genes from Rhizobium leguminosarum bv. viciae are controlled by the nitrogen fixation regulatory protein NifA. Proc Natl Acad Sci USA. 1997, 94: 6019-6024. 10.1073/pnas.94.12.6019.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  22. Fellay R, Hanin M, Montorzi G, Frey J, Freiberg C, Golinowski W, Staehelin C, Broughton WJ, Jabbouri S: nodD2 of Rhizobium sp NGR234 is involved in the repression of the nodABC operon. Mol Microbiol. 1998, 27: 1039-1050. 10.1046/j.1365-2958.1998.00761.x.

    Article  PubMed  CAS  Google Scholar 

  23. Vlassak KM, de Wilde P, Snoeck C, Luyten E, van Rhijn P, Vanderleyden J: The Rhizobium sp. BR816 nodD3 gene is regulated by a transcriptional regulator of the AraC/XylS family. Mol Gen Genet. 1998, 258: 558-561. 10.1007/s004380050768.

    Article  PubMed  CAS  Google Scholar 

  24. Dusha I, Austin S, Dixon R: The upstream region of the nodD3 gene of Sinorhizobium meliloti carries enhancer sequences for the transcriptional activator NtrC. FEMS Microbiol Lett. 1999, 179: 491-499. 10.1016/S0378-1097(99)00450-4.

    Article  PubMed  CAS  Google Scholar 

  25. Garcia-Rodríguez FM, Toro N: Sinorhizobium meliloti nfe (nodulation formation efficiency) genes exhibit temporal and spatial expression patterns similar to those of genes involved in symbiotic nitrogen fixation. Mol Plant Microbe Interact. 2000, 13: 583-591.

    Article  PubMed  Google Scholar 

  26. Gao M, D'Haeze W, De Rycke R, Holsters M: Dual control of the nodA operon of Azorhizobium caulinodans ORS571 by a nod box and a NifA-sigma54-type promoter. Mol Genet Genomics. 2001, 265: 1050-1059. 10.1007/s004380100500.

    Article  PubMed  CAS  Google Scholar 

  27. Luyten E, Swinnen E, Vlassak K, Verreth C, Dombrecht B, Vanderleyden J: Analysis of a symbiosis-specific cytochrome P450 homolog in Rhizobium sp. BR816. Mol Plant Microbe Interact. 2001, 14: 918-924.

    Article  PubMed  CAS  Google Scholar 

  28. Freiberg C, Fellay R, Bairoch A, Broughton WJ, Rosenthal A, Perret X: Molecular basis of symbiosis between Rhizobium and legumes. Nature. 1997, 387: 394-401. 10.1038/387394a0.

    Article  PubMed  CAS  Google Scholar 

  29. Göttfert M, Rothlisberger S, Kundig C, Beck C, Marty R, Hennecke H: Potential symbiosis-specific genes uncovered by sequencing a 410-kilobase DNA region of the Bradyrhizobium japonicum chromosome. J Bacteriol. 2001, 183: 1405-1412. 10.1128/JB.183.4.1405-1412.2001.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Sullivan JT, Trzebiatowski JR, Cruickshank RW, Gouzy J, Brown SD, Elliot RM, Fleetwood DJ, McCallum NG, Rossbach U, Stuart GS, et al: Comparative sequence analysis of the symbiosis island of Mesorhizobium loti strain R7A. J Bacteriol. 2002, 184: 3086-3095. 10.1128/JB.184.11.3086-3095.2002.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  31. Kaneko T, Nakamura Y, Sato S, Asamizu E, Kato T, Sasamoto S, Watanabe A, Idesawa K, Ishikawa A, Kawashima K, et al: Complete genome structure of the nitrogen-fixing symbiotic bacterium Mesorhizobium loti. DNA Res. 2000, 7: 331-338.

    Article  PubMed  CAS  Google Scholar 

  32. Galibert F, Finan TM, Long SR, Puhler A, Abola P, Ampe F, Barloy-Hubler F, Barnett MJ, Becker A, Boistard P, et al: The composite genome of the legume symbiont Sinorhizobium meliloti. Science. 2001, 293: 668-672.

    Article  PubMed  CAS  Google Scholar 

  33. Wood DW, Setubal JC, Kaul R, Monks DE, Kitajima JP, Okura VK, Zhou Y, Chen L, Wood GE, Almeida NF, et al: The genome of the natural genetic engineer Agrobacterium tumefaciens C58. Science. 2001, 294: 2317-2323. 10.1126/science.1066804.

    Article  PubMed  CAS  Google Scholar 

  34. DelVecchio VG, Kapatral V, Redkar RJ, Patra G, Mujer C, Los T, Ivanova N, Anderson I, Bhattacharyya A, Lykidis A, et al: The genome sequence of the facultative intracellular pathogen Brucella melitensis. Proc Natl Acad Sci USA. 2002, 99: 443-448. 10.1073/pnas.221575398.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  35. Wu ZL, Charles TC, Wang H, Nester EW: The ntrA gene of Agrobacterium tumefaciens: identification, cloning, and phenotype of a site-directed mutant. J Bacteriol. 1992, 174: 2720-2723.

    PubMed  CAS  PubMed Central  Google Scholar 

  36. Benítez-Bellón E, Moreno-Hagelsieb G, Collado-Vides J: Evaluation of thresholds for the detection of binding sites for regulatory proteins in Escherichia coli K12 DNA. Genome Biol. 2002, 3: research0013.1-0013.16. 10.1186/gb-2002-3-3-research0013.

    Article  Google Scholar 

  37. Gubler M, Hennecke H: Regulation of the fixA gene and fixBC operon in Bradyrhizobium japonicum. J Bacteriol. 1988, 170: 1205-1214.

    PubMed  CAS  PubMed Central  Google Scholar 

  38. Michiels J, D'hooghe I, Verretch C, Pelemans H, Vanderleyden J: Characterization of the Rhizobium leguminosarum biovar phaseoli nifA gene, a positive regulator of nif gene expression. Arch Microbiol. 1994, 161: 404-408. 10.1007/s002030050073.

    Article  PubMed  CAS  Google Scholar 

  39. Dombrecht B, Tesfay MZ, Verreth C, Heusdens C, Nápoles MC, Vanderleyden J, Michiels J: The Rhizobium etli gene iscN is highly expressed in bacteroids and required for nitrogen fixation. Mol Gen Genomics. 2002, 267: 820-828. 10.1007/s00438-002-0715-0.

    Article  CAS  Google Scholar 

  40. Dombrecht B: The complex regulation and role of the Rhizobium etli RpoN regulon in the symbiotic interaction with the common bean plant (Phaseolus vulgaris L.). 2001, Leuven: Katholieke Universiteit Leuven

    Google Scholar 

  41. Michiels J, Dirix G, Vanderleyden J, Xi CW: Processing and export of peptide pheromones and bacteriocins in gram-negative bacteria. Trends Microbiol. 2001, 9: 164-168. 10.1016/S0966-842X(01)01979-5.

    Article  PubMed  CAS  Google Scholar 

  42. Perret X, Freiberg C, Rosenthal A, Broughton WJ, Fellay R: High-resolution transcriptional analysis of the symbiotic plasmid of Rhizobium sp. NGR234. Mol Microbiol. 1999, 32: 415-425. 10.1046/j.1365-2958.1999.01361.x.

    Article  PubMed  CAS  Google Scholar 

  43. Van Soom C, de Wilde P, Vanderleyden J: HoxA is a transcriptional regulator for expression of the hup structural genes in free-living Bradyrhizobium japonicum. Mol Microbiol. 1997, 23: 967-977. 10.1046/j.1365-2958.1997.2781648.x.

    Article  PubMed  CAS  Google Scholar 

  44. Durmowicz MC, Maier RJ: The FixK(2) protein is involved in regulation of symbiotic hydrogenase expression in Bradyrhizobium japonicum. J Bacteriol. 1998, 180: 3253-3256.

    PubMed  CAS  PubMed Central  Google Scholar 

  45. Becana M, Dalton DA, Moran JF, Iturbe-Ormaetxe I, Matamoros MA, Rubio MC: Reactive oxygen species and antioxidants in legume nodules. Physiol Plant. 2000, 109: 372-381. 10.1034/j.1399-3054.2000.100402.x.

    Article  CAS  Google Scholar 

  46. Nienaber A, Huber A, Gottfert M, Hennecke H, Fischer HM: Three new NifA-regulated genes in the Bradyrhizobium japonicum symbiotic gene region discovered by competitive DNA-RNA hybridization. J Bacteriol. 2000, 182: 1472-1480. 10.1128/JB.182.6.1472-1480.2000.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  47. Downie JA: Functions of rhizobial nodulation genes. In The Rhizobiaceae, Molecular Biology of Model Plant-Associated Bacteria. Edited by: Spaink HP, Kondorosi A, Hooykaas PJJ. 1998, Dordrecht: Kluwer, 403-416.

    Google Scholar 

  48. Rosenblueth M, Hynes MF, Martinez-Romero E: Rhizobium tropici teu genes involved in specific uptake of Phaseolus vulgaris bean exudate compounds. Mol Gen Genet. 1998, 258: 587-598. 10.1007/s004380050772.

    Article  PubMed  CAS  Google Scholar 

  49. Marie C, Broughton WJ, Deakin WJ: Rhizobium type III secretion systems: legume charmers or alarmers?. Curr Opin Plant Biol. 2001, 4: 336-342. 10.1016/S1369-5266(00)00182-5.

    Article  PubMed  CAS  Google Scholar 

  50. Snoeck C: Host specificity determinants of Sinorhizobium sp. B816 for early signaling during symbiotic interactions. 2001, Leuven: Katholieke Universiteit Leuven

    Google Scholar 

  51. Taté R, Riccio A, Merrick M, Patriarca EJ: The Rhizobium etli amtB gene coding for an NH4+ transporter is down-regulated early during bacteroid differentiation. Mol Plant Microbe Interact. 1998, 11: 188-198.

    Article  PubMed  Google Scholar 

  52. Ercolano E, Mirabella R, Merrick M, Chiurazzi M: The Rhizobium leguminosarum glnB gene is down-regulated during symbiosis. Mol Gen Genet. 2001, 264: 555-564. 10.1007/s004380000333.

    Article  PubMed  CAS  Google Scholar 

  53. Taté R, Cermola M, Riccio A, Iaccarino M, Merrick M, Favre R, Patriarca EJ: Ectopic expression of the Rhizobium etli amtB gene affects the symbiosome differentiation process and nodule development. Mol Plant Microbe Interact. 1999, 12: 515-525.

    Article  Google Scholar 

  54. Soupène E, Foussard M, Boistard P, Truchet G, Batut J: Oxygen as a key developmental regulator of Rhizobium meliloti N2-fixation gene expression within the alfalfa root nodule. Proc Natl Acad Sci USA. 1995, 92: 3759-3763.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Gándara B, Merino AL, Rogel MA, Martinez-Romero E: Limited genetic diversity of Brucella spp. J Clin Microbiol. 2001, 39: 235-240. 10.1128/JCM.39.1.235-240.2001.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Kullik I, Fritsche S, Knobel H, Sanjuan J, Hennecke H, Fischer HM: Bradyrhizobium japonicum has two differentially regulated, functional homologs of the sigma 54 gene (rpoN). J Bacteriol. 1991, 173: 1125-1138.

    PubMed  CAS  PubMed Central  Google Scholar 

  57. Ehrt S, Ornston LN, Hillen W: RpoN (sigma 54) is required for conversion of phenol to catechol in Acinetobacter calcoaceticus. J Bacteriol. 1994, 176: 3493-3499.

    PubMed  CAS  PubMed Central  Google Scholar 

  58. Michiels J, Van Soom T, D'hooghe I, Dombrecht B, Benhassine T, de Wilde P, Vanderleyden J: The Rhizobium etli rpoN locus: DNA sequence analysis and phenotypical characterization of rpoN, ptsN, and ptsA mutants. J Bacteriol. 1998, 180: 1729-1740.

    PubMed  CAS  PubMed Central  Google Scholar 

  59. Thijs G, Moreau Y, De Smet F, Mathys J, Lescot M, Rombauts S, Rouze P, De Moor B, Marchal K: INCLUSive: Integrated Clustering, Upstream sequence retrieval and motif Sampling. Bioinformatics. 2002, 18: 331-332. 10.1093/bioinformatics/18.2.331.

    Article  PubMed  CAS  Google Scholar 

  60. BioI@Sista: Software. [http://www.esat.kuleuven.ac.be/~dna/Biol/Software.html]

  61. Hertz GZ, Hartzell III GW, Stormo GD: Identification of consensus patterns in unaligned DNA sequences known to be functionally regulated. Comput Applic Biosci. 1990, 6: 81-92.

    CAS  Google Scholar 

  62. Hertz GZ, Stormo GD: Identification of consensus patterns in unaligned DNA and protein sequences: a large-deviation statistical byasis for penalizing gaps. In Bioinformatics and Genome Research. Edited by: Lim HA, Cantor CR. 1995, Singapore: World Scientific, 201-216.

    Google Scholar 

  63. RSA-tools. [http://rsat.ulb.ac.be/rsat/]

  64. Pfam: Sigma54_activat. [http://www.sanger.ac.uk/cgi-bin/Pfam/getacc?PF00158]

  65. Proteome analysis @ EBI. [http://www.ebi.ac.uk/proteome/index.html]

  66. Entrez-Protein. [http://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/entrez/query.fcgi?db=Protein]

  67. Sean Eddy's Lab:HMMER 2.2. [http://hmmer.wustl.edu/]

  68. FramePlot 2.3.2. [http://www.nih.go.jp/~jun/cgi-bin/frameplot.pl]

  69. Albright LM, Ronson CW, Nixon BT, Ausubel FM: Identification of a gene linked to Rhizobium meliloti ntrA whose product is homologous to a family to ATP-binding proteins. J Bacteriol. 1989, 171: 1932-1941.

    PubMed  CAS  PubMed Central  Google Scholar 

  70. Michel-Reydellet N, Desnoues N, de Zamaroczy M, Elmerich C, Kaminski PA: Characterisation of the glnK-amtB operon and the involvement of AmtB in methylammonium uptake in Azorhizobium caulinodans. Mol Gen Genet. 1998, 258: 671-677. 10.1007/s004380050781.

    Article  PubMed  CAS  Google Scholar 

  71. Michel-Reydellet N, Desnoues N, Elmerich C, Kaminski PA: Characterization of Azorhizobium caulinodans glnB and glnA genes: involvement of the P(II) protein in symbiotic nitrogen fixation. J Bacteriol. 1997, 179: 3580-3587.

    PubMed  CAS  PubMed Central  Google Scholar 

  72. Martin GB, Thomashow MF, Chelm BK: Bradyrhizobium japonicum glnB, a putative nitrogen-regulatory gene, is regulated by NtrC at tandem promoters. J Bacteriol. 1989, 171: 5638-5645.

    PubMed  CAS  PubMed Central  Google Scholar 

  73. Chiurazzi M, Iaccarino M: Transcriptional analysis of the glnB-glnA region of Rhizobium leguminosarum biovar viciae. Mol Microbiol. 1990, 4: 1727-1735.

    Article  PubMed  CAS  Google Scholar 

  74. Arcondéguy T, Huez I, Fourment J, Kahn D: Symbiotic nitrogen fixation does not require adenylylation of glutamine synthetase I in Rhizobium meliloti. FEMS Microbiol Lett. 1996, 145: 33-40. 10.1016/0378-1097(96)00381-3.

    Article  PubMed  Google Scholar 

  75. Martin GB, Chapman KA, Chelm BK: Role of the Bradyrhizobium japonicum ntrC gene product in differential regulation of the glutamine synthetase II gene (glnII). J Bacteriol. 1988, 170: 5452-5459.

    PubMed  CAS  PubMed Central  Google Scholar 

  76. Patriarca EJ, Chiurazzi M, Manco G, Riccio A, Lamberti A, De Paolis A, Rossi M, Defez R, Iaccarino M: Activation of the Rhizobium leguminosarum glnII gene by NtrC is dependent on upstream DNA sequences. Mol Gen Genet. 1992, 234: 337-345.

    Article  PubMed  CAS  Google Scholar 

  77. Szeto WW, Nixon BT, Ronson CW, Ausubel FM: Identification and characterization of the Rhizobium meliloti ntrC gene: R. meliloti has separate regulatory pathways for activation of nitrogen fixation genes in free-living and symbiotic cells. J Bacteriol. 1987, 169: 1423-1432.

    PubMed  CAS  PubMed Central  Google Scholar 

  78. Michiels J, Vanderleyden J: Cloning and sequence of the Rhizobium leguminosarum biovar phaseoli fixA gene. Biochim Biophys Acta. 1993, 1144: 232-233. 10.1016/0005-2728(93)90179-J.

    Article  PubMed  CAS  Google Scholar 

  79. Hontelez JG, Lankhorst RK, Katinakis P, van den Bos RC, van Kammen A: Characterization and nucleotide sequence of a novel gene fixW upstream of the fixABC operon in Rhizobium leguminosarum. Mol Gen Genet. 1989, 218: 536-544.

    Article  PubMed  CAS  Google Scholar 

  80. Bauer E, Kaspar T, Fischer HM, Hennecke H: Expression of the fixR-nifA operon in Bradyrhizobium japonicum depends on a new response regulator, RegR. J Bacteriol. 1998, 180: 3853-3863.

    PubMed  CAS  PubMed Central  Google Scholar 

  81. Quinto C, Delavega H, Flores M, Leemans J, Cevallos MA, Pardo MA, Azpiroz R, Girard MD, Calva E, Palacios R: Nitrogenase reductase: a functional multigene family in Rhizobium phaseoli. Proc Natl Acad Sci USA. 1985, 82: 1170-1174.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  82. Szeto WW, Zimmerman JL, Sundaresan V, Ausubel FM: A Rhizobium meliloti symbiotic regulatory gene. Cell. 1984, 36: 1035-1043.

    Article  PubMed  CAS  Google Scholar 

  83. Aguilar OM, Reilander H, Arnold W, Puhler A: Rhizobium meliloti nifN (fixF) gene is part of an operon regulated by a nifA-dependent promoter and codes for a polypeptide homologous to the nifK gene product. J Bacteriol. 1987, 169: 5393-5400.

    PubMed  CAS  PubMed Central  Google Scholar 

  84. Reid CJ, Poole PS: Roles of DctA and DctB in signal detection by the dicarboxylic acid transport system of Rhizobium leguminosarum. J Bacteriol. 1998, 180: 2660-2669.

    PubMed  CAS  PubMed Central  Google Scholar 

  85. Jiang J, Gu BH, Albright LM, Nixon BT: Conservation between coding and regulatory elements of Rhizobium meliloti and Rhizobium leguminosarum dct genes. J Bacteriol. 1989, 171: 5244-5253.

    PubMed  CAS  PubMed Central  Google Scholar 

  86. Murphy PJ, Heycke N, Trenz SP, Ratet P, de Bruijn FJ, Schell J: Synthesis of an opine-like compound, a rhizopine, in alfalfa nodules is symbiotically regulated. Proc Natl Acad Sci USA. 1988, 85: 9133-9137.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  87. Fischer HM, Babst M, Kaspar T, Acuna G, Arigoni F, Hennecke H: One member of a groESL-like chaperonin multigene family in Bradyrhizobium japonicum is co-regulated with symbiotic nitrogen fixation genes. EMBO J. 1993, 12: 2901-2912.

    PubMed  CAS  PubMed Central  Google Scholar 

  88. Weidenhaupt M, Fischer HM, Acuna G, Sanjuan J, Hennecke H: Use of a promoter-probe vector system in the cloning of a new NifA-dependent promoter (ndp) from Bradyrhizobium japonicum. Gene. 1993, 129: 33-40. 10.1016/0378-1119(93)90693-W.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

We thank I. Van Cauteren and G. Thijs for computational assistance. B.D. is a recipient of a postdoctoral fellowship of the Onderzoeksfonds Katholieke Universiteit Leuven. K.M. is a postdoctoral fellow of the Fund for Scientific Research (Flanders). We acknowledge financial support from the Flemish government (GOA98 to J.V.).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jan Michiels.

Electronic supplementary material

Rights and permissions

Reprints and permissions

About this article

Cite this article

Dombrecht, B., Marchal, K., Vanderleyden, J. et al. Prediction and overview of the RpoN-regulon in closely related species of the Rhizobiales. Genome Biol 3, research0076.1 (2002). https://0-doi-org.brum.beds.ac.uk/10.1186/gb-2002-3-12-research0076

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/gb-2002-3-12-research0076

Keywords