Acessibilidade / Reportar erro

Bradyrhizobium elkanii nod regulon: insights through genomic analysis

Abstract

A successful symbiotic relationship between soybean [Glycine max (L.) Merr.] and Bradyrhizobium species requires expression of the bacterial structural nod genes that encode for the synthesis of lipochitooligosaccharide nodulation signal molecules, known as Nod factors (NFs). Bradyrhizobium diazoefficiens USDA 110 possesses a wide nodulation gene repertoire that allows NF assembly and modification, with transcription of the nodYABCSUIJnolMNOnodZ operon depending upon specific activators, i.e., products of regulatory nod genes that are responsive to signaling molecules such as flavonoid compounds exuded by host plant roots. Central to this regulatory circuit of nod gene expression are NodD proteins, members of the LysR-type regulator family. In this study, publicly available Bradyrhizobium elkanii sequenced genomes were compared with the closely related B. diazoefficiens USDA 110 reference genome to determine the similarities between those genomes, especially with regards to the nod operon and nod regulon. Bioinformatics analyses revealed a correlation between functional mechanisms and key elements that play an essential role in the regulation of nod gene expression. These analyses also revealed new genomic features that had not been clearly explored before, some of which were unique for some B. elkanii genomes.

Keywords:
Bradyrhizobium; NodD1 protein; nod box; nod genes

Introduction

Soybean symbiotic partners mainly belong to the genus Bradyrhizobium, initially proposed as a group of slow-growing, alkaline-producing root nodule nitrogen-fixing bacteria (Somasegaran and Hoben, 1994Somasegaran P and Hoben HJ (1994) Handbook for Rhizobia: Methods in Legume-Rhizobium Technology. Springer-Verlag, New York, 450 p.). Genetic and physiological information, including biochemical profile, DNA homology, and phylogenomics, is essential for clarifying the differences among isolates and for supporting the taxonomic classification of over 25 species in the genus Bradyrhizobium (Jordan, 1982Jordan DC (1982) Transfer of Rhizobium japonicum Buchanan 1980 to Bradyrhizobium gen. nov., a genus of slow-growing, root nodule bacteria from leguminous plants. Int J Sys Evol Microb 32:136-139.; Rivas et al., 2009Rivas R, Martens M, de Lajudie P and Willems A (2009) Multilocus sequence analysis of the genus Bradyrhizobium. Syst Appl Microbiol 32:101-110.). Among the species of this genus, considerable research efforts have focused on B. japonicum and B. elkanii because of their commercial use as a source for inoculant formulations, with many effective long-term programs for elite strain identification and selection being undertaken in different countries (Campos et al., 2001Campos BC, Hungria M and Tedesco V (2001) Eficiência da fixação biológica de N2 por estirpes de Bradyrhizobium na soja em plantio direto. Rev Bras Ciênc Sol 25:583-592.; Kober et al., 2004Kober MdV, de Sá ELS, Freire JRJ and Giongo A (2004) Characterization of variants of Bradyrhizobium elkanii and B. japonicum and symbiotic behaviour in soybeans. Ciênc Rural 34:1459-1464.; Melchiorre et al., 2011Melchiorre M, de Luca M, Gonzalez Anta G, Suarez P, Lopez C, Lascano R and Racca R (2011) Evaluation of bradyrhizobia strains isolated from field-grown soybean plants in Argentina as improved inoculants. Biol Fertil Soils 47:81-89.).

The characterization of symbiosis itself is not an easy task because of the high degree of specialization involved in this phenomenon, particularly since a restricted group of rhizobial species/strains interacts with only a select range of plant species/varieties and vice versa. Moreover, specificity may occur at distinct stages of the interaction, from the very initial bacterial contact with the plant root to late nodule development and nitrogen fixation, resulting in biological nitrogen fixation that varies according to the host-microsymbiont combination (Schumpp and Deakin, 2010Schumpp O and Deakin WJ (2010) How inefficient rhizobia prolong their existence within nodules. Trends Plant Sci 15:189-195.; Wang et al., 2012Wang D, Yang S, Tang F and Zhu H (2012) Symbiosis specificity in the legume – Rhizobial mutualism. Cell Microbiol 14:334-342.). In view of this, symbiosis can be conceived as a complex framework promoted by strong evolutionary forces that involves a stringent initial molecular dialogue and signal exchange between the symbiotic partners. The establishment of a successful mutualistic relationship initially requires products of the bacterial structural nod genes that encode for the synthesis of lipochitooligosaccharide (LCO) nodulation signaling molecules, also known as Nod factors (NFs). Subsequently, the recognition of symbiotic NFs by the plant triggers a signaling cascade that ultimately allows bacterial infection and induces de novo organogenesis of the nodule to accommodate the symbiont and further support nitrogen fixation (Tóth and Stacey, 2015Tóth K and Stacey G (2015) Does plant immunity play a critical role during initiation of the legume-rhizobium symbiosis? Front Plant Sci 6:401.). Bradyrhizobium diazoefficiens USDA 110 (formerly B. japonicum USDA 110) has a wide nodulation gene repertoire involved in NF assembly and modification.

In several species of rhizobia nod genes are frequently organized in an operon (nod operon), which suggests that regulation of their expression involves common mechanisms (Denarie et al., 1993Denarie J, Debelle F, Truchet G and Prome J-C (1993) Rhizobium and legume nodulation: A molecular dialogue. In: Palacios R, Mora J and Newton W (eds) New Horizons in Nitrogen Fixation. Vol 17. Current Plant Science and Biotechnology in Agriculture. Springer, Dordrecht, pp 19-30.). Indeed, transcription of the nodYABCSUIJnolMNO operon in B. diazoefficiens USDA 110 depends upon transcriptional activators, i.e., products of regulatory nod genes responsive to signaling molecules, such as flavonoid compounds, exuded by host plant roots (Luka et al., 1993Luka S, Sanjuan J, Carlson RW and Stacey G (1993) nolMNO genes of Bradyrhizobium japonicum are co-transcribed with nodYABCSUIJ, and nolO is involved in the synthesis of the lipo-oligosaccharide nodulation signals. J Biol Chem 268:27053-27059.; Loh and Stacey, 2003Loh J and Stacey G (2003) Nodulation gene regulation in Bradyrhizobium japonicum: A unique integration of global regulatory circuits. Appl Environ Microbiol 69:10-17.). Central to this regulatory circuit of nod gene expression are NodD proteins, members of the LysR-type regulator family. Upon activation by a particular flavonoid ligand NodD proteins can bind to specific DNA motifs upstream of the nod operon, the so called nod boxes, and selectively control the expression of structural nod genes in the early stages of plant-bacteria interaction (Hong et al., 1987Hong GF, Burn JE and Johnston AW (1987) Evidence that DNA involved in the expression of nodulation (nod) genes in Rhizobium binds to the product of the regulatory gene nodD. Nucleic Acids Res 15:9677-9690.; Henikoff et al., 1988Henikoff S, Haughn GW, Calvo JM and Wallace JC (1988) A large family of bacterial activator proteins. Proc Natl Acad Sci U S A 85:6602-6606.; Goethals et al., 1992Goethals K, Van Montagu M and Holsters M (1992) Conserved motifs in a divergent nod box of Azorhizobium caulinodans ORS571 reveal a common structure in promoters regulated by LysR-type proteins. Proc Natl Acad Sci U S A 89:1646-1650.). Although multiple isoforms of NodD proteins have been identified in distinct rhizobial species, perhaps indicative of a role in expanding the plant host spectrum of these symbionts, only two were found in the B. diazoefficiens USDA 110 genome, namely, NodD1 and NodD2, products of the nodD1 and nodD2 genes, respectively (Gottfert et al., 1989Gottfert M, Lamb JW, Gasser R, Semenza J and Hennecke H (1989) Mutational analysis of the Bradyrhizobium japonicum common nod genes and further nod box-linked genomic DNA regions. Mol Gen Genet 215:407-415.). These two proteins show distinct expression patterns and play different functional roles in regulating the expression of structural nod genes. Active NodD1, i.e., in the presence of a flavonoid ligand molecule such as genistein in soybean root exudates, operates as a positive transcriptional regulator of the nod operon in B. diazoefficiens. Unique to this organism when compared to other known rhizobial species, NodD1 is not constitutively expressed; instead, it is induced by flavonoid compounds and shows autoregulation (Banfalvi et al., 1988Banfalvi Z, Nieuwkoop A, Schell M, Besl L and Stacey G (1988) Regulation of nod gene expression in Bradyrhizobium japonicum. Mol Gen Genet 214:420-424.). Although nodulation requires nod gene induction by flavonoids in most diazotrophs, the efficiency of this process depends on appropriate spatial and temporal expression of these genes. Hence, it is not surprising that, in addition to positive transcriptional regulators, nod genes are also controlled by repressor elements, as is the case for NodD2 protein, that acts as a negative regulator of the nod operon (Loh and Stacey, 2003Loh J and Stacey G (2003) Nodulation gene regulation in Bradyrhizobium japonicum: A unique integration of global regulatory circuits. Appl Environ Microbiol 69:10-17.).

In B. diazoefficiens USDA 110, the core regulatory mechanism involving NodD1 and NodD2 is extended with additional regulators that act synergistically with NodD proteins to modulate the expression of nod genes. Of these, the roles of NolA, a MerR-type regulator encoded by nolA, and NodVW, the product of nodVW, that form a two-component regulatory system are particularly noteworthy. Initially identified as a soybean genotype-specific nodulation factor, NolA was later shown to be an activator of nodD2 involved in negative regulation with NodD2, with both involved in the feedback and quorum regulation of nod genes (Sadowsky et al., 1991Sadowsky MJ, Cregan PB, Gottfert M, Sharma A, Gerhold D, Rodriguez-Quinones F, Keyser HH, Hennecke H and Stacey G (1991) The Bradyrhizobium japonicum nolA gene and its involvement in the genotype-specific nodulation of soybeans. Proc Natl Acad Sci U S A 88:637-641.; Garcia et al., 1996Garcia M, Dunlap J, Loh J and Stacey G (1996) Phenotypic characterization and regulation of the nolA gene of Bradyrhizobium japonicum. Mol Plant Microbe Interact 9:625-636.; Loh and Stacey, 2003Loh J and Stacey G (2003) Nodulation gene regulation in Bradyrhizobium japonicum: A unique integration of global regulatory circuits. Appl Environ Microbiol 69:10-17.). On the other hand, the NodVW two-component regulatory system provides an alternative flavonoid responsive pathway for nod gene activation, which explains the residual nodulation of soybean plants in NodD1 mutants (Göttfert et al., 1990Göttfert M, Grob P and Hennecke H (1990) Proposed regulatory pathway encoded by the nodV and nodW genes, determinants of host specificity in Bradyrhizobium japonicum. Proc Natl Acad Sci U S A 87:2680-2684.; Loh et al., 1997Loh J, Garcia M and Stacey G (1997) NodV and NodW, a second flavonoid recognition system regulating nod gene expression in Bradyrhizobium japonicum. J Bacteriol 179:3013-3020.).

As the major source for the soybean inoculant industry, strains of B. japonicum and B. elkanii differ markedly in their physiology and in their competitive fitness (Minamisawa, 1989Minamisawa K (1989) Comparison of extracellular polysaccharide composition, rhizobitoxine production, and hydrogenase phenotype among various strains of Bradyrhizobium japonicum. Plant Cell Physiol 30:877-884.; Vasilas and Furhmann, 1993Vasilas BL and Furhmann JJ (1993) Field response of soybean to nodulation by a rhizobitoxine-producing strain of Bradyrhizobium. Agron J 85:302-305.). Such differences indicate the need to identify and characterize the genetic nature of specificity in the symbiotic relationship as a crucial step in developing guided strategies to enhance the effectiveness of soybean inoculants, given that biological nitrogen fixation ultimately provides a resource for more sustainable agricultural systems.

While there is a considerable amount of knowledge regarding the genetics and molecular mechanisms of the soybean symbiont B. diazoefficiens, including the complete genome sequence, information on the genetics of B. elkanii is restricted mainly to research focused on specific features. Although genomic data and a few drafts of the genome are available for B. elkanii, comparative genomic analyses of these two species have not been reported. In this study, publicly available B. elkanii genomes were compared with the closely related B. diazoefficiens USDA 110 reference genome (Kaneko et al., 2002Kaneko T, Nakamura Y, Sato S, Minamisawa K, Uchiumi T, Sasamoto S, Watanabe A, Idesawa K, Iriguchi M, Kawashima K, et al. (2002) Complete genomic sequence of nitrogen-fixing symbiotic bacterium Bradyrhizobium japonicum USDA110. DNA Res 9:189-197.) to gain some insights into the mechanisms of nod gene expression in B. elkanii, especially for the nod operon and nod regulon. The results of this analysis should provide a more comprehensive understanding of the molecular dynamics and complexity of mechanisms involved in fine-tuning signal communication between this symbiont and its host plants.

Material and Methods

Strains and genomic data

Genomic data consisting of the complete genome of B. diazoefficiens USDA 110 (reference genome) and draft genomes of B. elkanii strains SEMIA 587, CCBAU 05737, CCBAU 43297, USDA 94, USDA 3254 and USDA 3259 were obtained from the publicly available database of The National Center for Biotechnology Information (NCBI; http://www.ncbi.nlm.nih.gov). Genome features and accession numbers are described in Table 1. The type-strain taxonomy of each organisms genome was confirmed by 16S rRNA sequence analysis using the RDP SeqMatch k-nearest-neighbor (k-NN) classifier (Wang et al., 2007Wang Q, Garrity GM, Tiedje JM and Cole JR (2007) Naïve Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl Environ Microbiol 73:5261-5267.) and checked with Basic Local Alignment Search Tool (BLAST) searches and pairwise global sequence alignments implemented in the EzTaxon server database (Kim et al., 2012Kim O-S, Cho Y-J, Lee K, Yoon S-H, Kim M, Na H, Park S-C, Jeon YS, Lee J-H, Yi H, et al. (2012) Introducing EzTaxon-e: A prokaryotic 16S rRNA gene sequence database with phylotypes that represent uncultured species. Int J Sys Evol Microb 62:716-721.).

Table 1
Characteristics of the genomes analysed.

Bioinformatics analyses

Functional annotation was done with the Rapid Annotation using Subsystem Technology (RAST) server (Aziz et al., 2008Aziz RK, Bartels D, Best AA, DeJongh M, Disz T, Edwards RA, Formsma K, Gerdes S, Glass E, Kubal M, et al. (2008) The RAST Server: Rapid annotations using subsystems technology. BMC Genomics 9:75.), with Glimmer set for gene calling (Salzberg et al., 1998Salzberg SL, Delcher AL, Kasif S and White O (1998) Microbial gene identification using interpolated Markov models. Nucleic Acids Res 26:544-548.), frameshift correction, backfilling of gaps and automatic fixing errors. Assigned functional features were triple-checked with InterProScan (Zdobnov and Apweiler, 2001Zdobnov EM and Apweiler R (2001) InterProScan – An integration platform for the signature-recognition methods in InterPro. Bioinformatics 17:847-848.) by the signature-recognition method in the InterPro database (Hunter et al., 2009Hunter S, Apweiler R, Attwood TK, Bairoch A, Bateman A, Binns D, Bork P, Das U, Daugherty L, Duquenne L, et al. (2009) InterPro: The integrative protein signature database. Nucleic Acids Res 37:211-215.), ScanProsite (de Castro et al., 2006de Castro E, Sigrist CA, Gattiker A, Bulliard V, Langendijk-Genevaux PS, Gasteiger E, Bairoch A and Hulol N (2006) ScanProsite: Detection of PROSITE signature matches and ProRule-associated functional and structural residues in proteins. Nucleic Acids Res 34:362-365.) for protein signature matches in the PROSITE database (Sigrist et al., 2010Sigrist CJA, Cerutti L, de Castro E, Langendijk-Genevaux PS, Bulliard V, Bairoch A and Hulo N (2010) PROSITE, a protein domain database for functional characterization and annotation. Nucleic Acids Res 38:161-166.), and BLASTp against the UniProtKB database (Magrane and Consortium, 2011Magrane M and Consortium U (2011) UniProt Knowledgebase: A hub of integrated protein data. Database (Oxford) 2011:bar009.). An inventory of genes involved in nodulation (structural- and regulatory-nod/nol genes) for each genome in this study is provided in Table S1. The B. diazoefficiens USDA 110 genome (Kaneko et al., 2002Kaneko T, Nakamura Y, Sato S, Minamisawa K, Uchiumi T, Sasamoto S, Watanabe A, Idesawa K, Iriguchi M, Kawashima K, et al. (2002) Complete genomic sequence of nitrogen-fixing symbiotic bacterium Bradyrhizobium japonicum USDA110. DNA Res 9:189-197.) was used as the reference genome and missing nod/nol genes were searched for in other genomes with BLASTn using homologous nucleotide sequences of the closely-related reference species. Possible frameshift annotation errors in assigned genes were corrected and Open Reading Frames (ORFs) were checked with the Expert Protein Analysis System (ExPASy) translate tool (Gasteiger et al., 2003Gasteiger E, Gattiker A, Hoogland C, Ivanyi I, Appel RD and Bairoch A (2003) ExPASy: The proteomics server for in-depth protein knowledge and analysis. Nucleic Acids Res 31:3784-3788.) by comparing with the respective reference.

The subcellular localization of proteins was predicted using sequence-based tools in a coordinated fashion (Emanuelsson et al., 2007Emanuelsson O, Brunak S, von Heijne G and Nielsen H (2007) Locating proteins in the cell using TargetP, SignalP and related tools. Nat Protocols 2:953-971.). Initially, SignalP 4.1 software (Petersen et al., 2011Petersen TN, Brunak S, von Heijne G and Nielsen H (2011) SignalP 4.0: Discriminating signal peptides from transmembrane regions. Nat Methods 8:785-786.) was used to screen amino acid sequences for the presence and location of signal peptide cleavage sites characteristic of secretory proteins and this was followed by the ab initio prediction of non-classical protein secretion, i.e., secretion not triggered by a signal peptide, with SecretomeP 2.0 (Bendtsen et al., 2004Bendtsen JD, Jensen LJ, Blom N, von Heijne G and Brunak S (2004) Feature-based prediction of non-classical and leaderless protein secretion. Prot Eng Des Select 17:349-356.). Lipoprotein signal peptides and N-terminal membrane helix prediction was done using LipoP 1.0 (Rahman et al., 2008Rahman O, Cummings S, Harrington D and Sutcliffe I (2008) Methods for the bioinformatic identification of bacterial lipoproteins encoded in the genomes of Gram-positive bacteria. World J Microbiol Biotechnol 24:2377-2382.), while the presence and location of potential twin-arginine translocation signal peptide cleavage sites was verified with TatP 1.0 (Bendtsen et al., 2005Bendtsen J, Nielsen H, Widdick D, Palmer T and Brunak S (2005) Prediction of twin-arginine signal peptides. BMC Bioinformatics 6:167.). Finally, the prediction of transmembrane topology in proteins was assessed using a combined approach based on a hidden Markov model algorithm implemented in the TMHMM 2.0 server (Krogh et al., 2001Krogh A, Larsson B, von Heijne G and Sonnhammer ELL (2001) Predicting transmembrane protein topology with a hidden Markov model: Application to complete genomes. J Mol Biol 305:567-580.) and Phobius (Käll et al., 2004Käll L, Krogh A and Sonnhammer ELL (2004) A combined transmembrane topology and signal peptide prediction. Meth J Mol Biol 338:1027-1036.).

All bioinformatics analyses used the default parameters of the respective software.

Phylogenetic and sequence analyses

Multiple sequence alignment of NodD proteins was done by distance estimation using kmer counting and progressive alignment with log-expectation scores, followed by refinement using tree-dependent restricted partitioning implemented by MUSCLE (Edgar, 2004Edgar RC (2004) MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32:1792-1797.). The algorithm was implemented in the Molecular Evolutionary Genetics Analysis - MEGA 6.0 package (Tamura et al., 2013Tamura K, Stecher G, Peterson D, Filipski A and Kumar S (2013) MEGA6: Molecular Evolutionary Genetics Analysis Version 6.0. Mol Biol Evol 30:2725-2729.) and the parameters were set for the Neighbor-Joining clustering method in all interactions, with −2.9 and 0 for gap opening and gap extension penalties, respectively (center specified as 0). Subsequent phylogenetic analysis and tree reconstruction were done using the Neighbor-Joining method in the same package, with the molecular distances of the aligned sequences computed based on p-distance parameters and 1,000 bootstrap replicates and pairwise deletion treatment for gaps. Point accepted mutation (PAM) 250 calculations were used as a substitution matrix model for scoring sequence alignments (Dayhoff and Schwartz, 1978Dayhoff MO and Schwartz RM (1978) A model of evolutionary change in proteins. In: Dayhoff MO (ed) Atlas of Protein Sequence and Structure. Natl Biomed Res Found, Washington DC, pp 345-352.).

Protein structure prediction and alignment

Protein 3D structure prediction was done using the SWISS-MODEL web server (Biasini et al., 2014Biasini, M, Bienert S, Waterhouse A, Arnold K, Studer G, Schmidt T, Kiefer F, Cassarino TG, Bertoni M, Bordoli L, et al. (2014) SWISS-MODEL: Modelling protein tertiary and quaternary structure using evolutionary information. Nucleic Acids Res 42:252-258.) based on evolutionarily-related structures, amino acid sequences and protein structure homologies. The technique uses hidden Markov model-sensitive searches run against the SWISS-MODEL template library (SMTL) to generate a structural model of the protein of interest. A QMEAN potential to assess model quality is then generated with an independent accuracy evaluation by the Continuous Automated Model EvaluatiOn project – CAMEO (Haas et al., 2013Haas J, Roth S, Arnold K, Kiefer F, Schmidt T, Bordoli L and Schwede T (2013) The Protein Model Portal - A comprehensive resource for protein structure and model information. Database 2013. doi:10.1093/database/bat031.
https://doi.org/10.1093/database/bat031...
) based on target sequences pre-released by the Protein Data Bank (PDB) (Berman et al., 2007Berman H, Henrick K, Nakamura H and Markley JL (2007) The worldwide Protein Data Bank (wwPDB): Ensuring a single, uniform archive of PDB data. Nucleic Acids Res 35:301-303.). The predicted protein structures were aligned with a combinatorial extension (CE) algorithm (Shindyalov and Bourne, 1998Shindyalov IN and Bourne PE (1998) Protein structure alignment by incremental combinatorial extension (CE) of the optimal path. Protein Eng 11:739-747.) implemented in the Research Collaboratory for Structural Bioinformatics (RCSB) PDB Protein Comparison Tool at www.rcsb.org (Bernstein et al., 1977Bernstein FC, Koetzle TF, Williams GJB, Meyer Jr EF, Brice MD, Rodgers JR, Kennard O, Shimanouchi T and Tasumi M (1977) The Protein Data Bank: A computer-based archival file for macromolecular structures. J Mol Biol 112:535-542.; Berman et al., 2007Berman H, Henrick K, Nakamura H and Markley JL (2007) The worldwide Protein Data Bank (wwPDB): Ensuring a single, uniform archive of PDB data. Nucleic Acids Res 35:301-303.; Goodsell et al., 2015Goodsell DS, Dutta S, Zardecki C, Voigt M, Berman HM and Burley SK (2015) The RCSB PDB “Molecule of the Month”: Inspiring a molecular view of biology. PLoS Biol 13:e1002140.). The parameters were set as 30 for the maximum gap size allowed during aligned fragment pairs (AFP) extension in fragment size m = 8, with gap open and gap extension penalties of 5.0 and 0.5, respectively.

Results

nod operon and regulatory genes

Annotation of the B. elkanii genome followed by manual curation revealed a conserved operon structure and organization of the nodKABCSUIJnolOnodZ genes in strains CCBAU 05737, CCBAU 43297, USDA 94, USDA 3254 and USDA 3259 similar to that in the B. diazoefficiens USDA 110 nod operon, with the substitution of nodY for the corresponding nodK and the lack of nolMN genes. Curiously, in B. elkanii SEMIA 587, a distinct pattern for this “canonical” gene organization was observed, with these genes scattered in the chromosome and organized as nodKABCS, a second sparse block containing nodI, nodJ, nolO and nodZ, and then nodU, which was separate from the other genes of this operon. As in the case of nolMN, nolZ was not identified in the genomes of any B. elkanii strain; likewise, nolY was not detected in the genomes of strains USDA 94, USDA 3254, and USDA 3259 (Figure 1).

Figure 1
Genetic map showing organization of the nod/nol genes in B. diazoefficiens USDA 110 (reference genome) and six B. elkanii genomes. Letters indicate the name of each gene in the genome. Gene classes are color coded and transcriptional orientation is indicated by the arrows. Relative gene distances are indicated or follow the scale presented. Gene location within the genome is represented by a straight line below the gene map of the respective chromosome (B. diazoefficiens) or contig (B. elkanii) ID.

Annotation also highlighted the presence of nodD1 and nodD2 regulatory genes in all six B. elkanii genomes and in the B. diazoefficiens USDA 110 genome. The organization of these two genes in the B. elkanii genomes followed the pattern observed for B. diazoefficiens USDA 110, i.e., they were positioned close to each other and close to the nod operon, although in opposite orientation. Similarly, the presence of nodVW that coded for the two-component regulatory elements was ubiquitous in all six B. elkanii genomes, with its location relative to the nod operon varying according to each genome. Conversely, the nolA gene showed a conserved location among the B. elkanii genomes, close to nodD2; the exception was for strain SEMIA 587, in which this gene was located at a position distant from nodD2 (Figure 1).

Analysis of nodD promoter regions

Analysis of a 250-bp region upstream of the nodD1 ORFs revealed that all six B. elkanii genomes and the B. diazoefficiens USDA 110 genome contained one −10/-35 σ70 potential promoter (TTGCTA-N17-TGGTAAAAT) located 46 bp upstream from the nodD1 CDS start site. Additionally, all sequences showed two 47-bp nod boxes, one of which corresponded to a consensus nod box sequence with the respective palindromic structure, located 11 bp from the nodD1 CDS start site that controls transcription of the nod operon. The second one was a presumptive nod box-like sequence located in the upstream region (84 bp) of the nodD1 CDS start site (Figure 2A). Interestingly, the consensus nod box sequence (Box1) overlapped the sequence of the −10/-35 σ70 putative nodD1 promoter, while the presumptive nod box-like sequence was located just 6 bp upstream to it (Figure 2A).

Figure 2
Alignment of the ~250-bp region upstream from nodD ORFs in Bradyrhizobium genomes. (a) The nodD1 nod box sequences are boxed and show a consensus nod box sequence near the nodD1 ORF start codon (Box 1) and a presumptive nod box-like sequence (Box 2) upstream, with conserved residues highlighted in grey, and (b) nodD2 partial nod box sequence structure (boxed) found in some B. elkanii genomes. Smaller boxes denote the putative nod box motifs (ATC-N9-GAT) with their respective palindromic structure. Yellow shaded sequences show a −10/-35 σ70 potential promoter. Sequences read 5' to 3' from right to left, with the first nucleotide in the sequence representing that immediately prior to ATG from the nodD1 (a) or nodD2 (b) CDS start site

On the other hand, analysis of the 250-bp region upstream of nodD2 revealed that only the B. elkanii SEMIA 587, CCBAU 05737, CCBAU 43297 and USDA 94 genomes exhibited a −10/-35 σ70 potential promoter with the sequence TTGTCG-N13-CTGTAAAAG, along with a partial nod box-like sequence located 152 bp upstream from the putative σ70 promoter. In the B. diazoefficiens USDA 110 and B. elkanii USDA 3254 and USDA 3259 genomes, neither of these elements could be identified in the nodD2 promoter region, i.e., they apparently had no potential −10/-35 σ70 promoters nor a partial (or complete) nod box-like sequence (Figure 2B).

The nolA promoter regions (250 bp) of all genomes in this study exhibited a −10/-35 σ70 potential promoter sequence (TTGAAT-N16-TTGTAGGCT), except for B. elkanii SEMIA 587. Additionally, apart from B. elkanii SEMIA 587, the nolA promoter regions displayed a high degree of sequence similarity among genomes (Figure S1A). Potential −10/-35 σ70 promoter sequences for the nodVW regulatory gene were found in all genomes, although there were differences in the structure and location of these promoters. In contrast to nolA, the promoter regions of the nodVW genes differed considerably in sequence conservation among genomes (Figure S1B). A search for the nod box in the nolA and nodVW promoter regions revealed the absence of this regulatory motif in the 250-bp upstream gene sequences in all genomes (Figure S1A,B).

Sequence analysis of regulatory Nod proteins

All the B. elkanii genomes displayed NodD1 ORFs of 314 amino acids, in agreement with the size of the NodD1 protein of B. diazoefficiens USDA 110. In contrast, there was discrete variation in the size of the NodD2 ORFs among genomes, with ORFs of 330 amino acids in B. diazoefficiens USDA 110, 331 amino acids in B. elkanii strains SEMIA 587, CCBAU 05737, CCBAU 43297 and USDA 94, and 329 residues in B. elkanii strains USDA 3254 and USDA 3259 (Table S1).

NodD1 and NodD2 showed a relatively high degree of conservation, although some dissimilarity among organisms ultimately clustered each B. elkanii protein into one of two groups (Figure S2). Indeed, a PAM250 matrix showed that B. elkanii strains SEMIA 587, CCBAU 05737, CCBAU 43297 and USDA 94 shared 100% global sequence similarity among themselves for both NodD1 and NodD2; consequently, each protein set for these strains grouped together. The same situation occurred with USDA 3254 and USDA 3259, although sequence similarity decreased by 7% for NodD1 and by 12% for NodD2 compared to their respective homologs from other strains within the same species. NodD regulator proteins from the B. elkanii strains were still conserved when compared to B. diazoefficiens USDA 110 orthologs, especially NodD1, which exhibited at least 93% sequence similarity, while for NodD2 no less than 73% similarity was observed (Figure S3A-B). Additionally, conservation plots for NodD1 and NodD2 proteins from the B. elkanii genomes revealed lower conservation at the carboxy-terminus region of the proteins, especially for NodD2, when compared to the reference genome of B. diazoefficiens USDA 110 (Figure S4A,B).

All the sequences analysed belonged to the LysR family of transcriptional regulators, and exhibited a 57-amino acid helix-turn-helix (HTH) LysR-type domain located at position 6-63. These domains also contained a 20-amino acid HTH DNA-binding motif from residues 23-42 in the amino-terminus region of NodD1 and NodD2 proteins. The NodD1 DNA-binding motif from B. elkanii strains SEMIA 587, CCBAU 05737, CCBAU 43297 and USDA 94 was highly conserved and showed 100% identity with the sequence LTAAARQINLSQPAMSAAIA (Figure 3A). Curiously, B. elkanii USDA 3254 and USDA 3259 showed 100% identity with the B. diazoefficiens USDA 110 NodD1 DNA-binding motif LTAAARKINLSQPAMSAAIA, with a single amino acid substitution occurring at position 29, in which glutamine (Q) was replaced by lysine (K) (Figure 3A).

Figure 3
Multiple sequence alignment (MSA) of the B. elkanii NodD HTH LysR-type domain and sequence logo showing conservation of the HTH DNA-binding motif. The HTH DNA-binding sequence conserved among organisms is boxed in red. The sequences are for the (a) NodD1 and (b) NodD2 domains.

The B. elkanii NodD2 DNA-binding motif also split into two groups, with strains SEMIA 587, CCBAU 05737, CCBAU 43297 and USDA 94 showing 100% identity for the sequence LTAAARKINLSQPAMSAAVA, while for strains USDA 3254 and USDA 3259 a serine (S) replaced the threonine (T) at position 24 of NodD2, resulting in the DNA-binding motif LSAAARKINLSQPAMSAAVA (Figure 3B). A comparison of NodD2 among Bradyrhizobium species showed that B. diazoefficiens USDA 110 shared higher similarity with B. elkanii strains USDA 3254 and USDA 3259, although an additional amino acid residue divergence at position 25 was observed, with replacement of an alanine (A) in B. elkanii for a serine (S) in B. diazoefficiens, providing the sequence LSSAARKINLSQPAMSAAVA (Figure 3B).

In silico analysis of the subcellular location of NodD1 and NodD2 proteins did not predict the presence of any significant ordinary signal peptide cleavage site in their amino acid sequences or any twin-arginine signal peptide cleavage site. Lipoprotein signal peptides and non-classical protein secretory patterns were also not identified. Additional protein topology and signal peptide examination did not identify any transmembrane helices. Finally, calculations using an integrative approach algorithm indicated that NodD1 and NodD2 were more likely to be cytoplasmic components rather than membrane-bound proteins.

Structural analysis of regulatory Nod proteins

Global protein structure alignment of NodD1 from the B. elkanii strains revealed high identity and similarity to NodD1 protein from the reference genome of B. diazoefficiens USDA 110. In particular, the genomes of B. elkanii strains USDA 3254 and USDA 3259 showed 92.62% and 97.65% structural identity and similarity for NodD1 compared to the orthologous protein from B. diazoefficiens USDA 110, respectively. Although slightly lower, the corresponding values for NodD1 structural identity and similarity from B. elkanii strains SEMIA 587, CCBAU 05737, CCBAU 43297 and USDA 94 were 91.61% and 95.97%, respectively. Analysis of the NodD1 structural alignments indicated that the higher identity and similarity of B. elkanii strains USDA 3254 and USDA 3259 with B. diazoefficiens USDA 110 resided mainly in the HTH motif of the DNA-binding domain (Figure S5A-F). In contrast, NodD2 global protein structure alignments for the B. elkanii SEMIA 587, CCBAU 05737, CCBAU 43297 and USDA 94 genomes showed higher identity (75.17%) and similarity (84.56%) with the NodD2 ortholog from B. diazoefficiens USDA 110 than with the B. elkanii USDA 3254 and USDA 3259 genomes, for which the identity and similarity were no higher than 74.16% and 82.89%, respectively (Figure S6A-F). Additionally, even though substantial identity and similarity were observed in the structural alignment of NodD2 between Bradyrhizobium species, the values were lower compared to the structural alignments for NodD1 proteins.

When the global structural alignment of NodD1 vs. NodD2 proteins was compared within each genome, the identity and similarity were still relatively significant. In the B. elkanii USDA 3254 and USDA 3259 genomes for example, the NodD1 vs. NodD2 structural alignment showed 70.47% identity and 81.88% similarity, while for the B. elkanii SEMIA 587, CCBAU 05737, CCBAU 43297 and USDA 94 genomes, the identity and similarity reached values of 69.46% and 81.54%, respectively. The B. diazoefficiens USDA 110 reference genome showed the most dissimilar structural alignment of NodD1 vs. NodD2 and consequently had the lowest identity (63.76%) and similarity (80.54%) values (Figure S7A-F). However, considering only the 57-residue protein segment corresponding to the DNA-binding domain containing the HTH motif, the NodD1 vs. NodD2 structural alignment values were higher in all genomes, being identical in the B. elkanii USDA 3254 and USDA 3259 genomes, i.e., 100% identity and similarity. The B. elkanii SEMIA 587, CCBAU 05737, CCBAU 43297 and USDA 94 genomes also showed better structural alignment of the amino-terminus region that accommodated the DNA-binding domain for the NodD1 vs. NodD2 comparison, with 91.38% identity and 98.28% similarity, compared to B. diazoefficiens USDA 110, which showed values of 81.03% and 93.10% for identity and similarity, respectively (Figure 4).

Figure 4
Structural alignment of the NodD1 and NodD2 57-residue HTH domain within each genome using the Combinatorial Extension (CE) algorithm. The alignments are for B. diazoefficiens USDA 110 (a), B. elkanii SEMIA 587 (b), B. elkanii CCBAU 05737 (c), B. elkanii CCBAU 43297 (d), B. elkanii USDA 94 (e), B. elkanii USDA 3254 (f) and B. elkanii USDA 3259 (g). Orange residues represent NodD1 superimposed on light-blue residues representing NodD2.

Discussion

nod operon/regulon in Bradyrhizobium genomes

Lipochitooligosaccharides (LCOs) produced by the action of bacterial nodulation (nod, nol, noe) gene products are key signaling molecules in establishing the Bradyrhizobium-legume symbiosis that triggers the formation of a new organ (the nodule) in which biological nitrogen fixation occurs (Day et al., 2000Day RB, Loh J, Cohn JR and Stacey G (2000) Signal exchange involved in the establishment of the Bradyrhizobium-legume symbiosis. In: Triplett E (ed) Prokaryotic Nitrogen Fixation, a Model System for the Analysis of a Biological Process. Horizon Scientific Press, Norfolk, pp 385-414.). Bradyrhizobium diazoefficiens and B. elkanii can nodulate soybean plants. Despite this common ability, detailed genetic and biochemical studies involving different Bradyrhizobium species have shown distinct physiological profiles that ultimately affect their symbiotic behavior, including nodulation capacity, nitrogen fixation efficiency and competitive ability for nodule formation (Kober et al., 2004Kober MdV, de Sá ELS, Freire JRJ and Giongo A (2004) Characterization of variants of Bradyrhizobium elkanii and B. japonicum and symbiotic behaviour in soybeans. Ciênc Rural 34:1459-1464.). Most of the structural and regulatory nod/nol genes with known function seem to be conserved between the B. diazoefficiens USDA 110 and B. elkanii genomes. Moreover, the presence of the structural nodY/KABCSUIJnolOnodZ genes, arranged in an operon in the B. diazoefficiens and most B. elkanii genomes, may indicate that similar patterns of LCO formation can be expected for these organisms, depending primarily upon a particular operon regulation in each genome.

The organization of nod genes in clusters/operons is a common and often conserved feature among various rhizobia genomes, many of them with a chromosomal location, as in the case of B. diazoefficiens, thereby facilitating coordination of the expression of these respective genes (Schlaman et al., 1998Schlaman HR, Phillips DA and Kondorosi E (1998) Genetic organization and transcriptional regulationof rhizobial nodulation genes. In: Spaink HP, Kondorosi A and Hooykaas P (eds) The Rhizobiaceae: Molecular Biology of Model Plant-Associated Bacteria. Springer, Dordrecht, pp 361-386.). Although nod genes in Bradyrhizobium are frequently organized in operons, this is not a fixed pattern, as seen in the B. elkanii SEMIA 587 genome which showed a reduced nodKABCS operon with additional nod/nol genes scattered along the chromosome. Interestingly, separation of the genes from this operon may result in the differential regulation of nod genes, ultimately providing an altered pattern of LCO formation for this strain compared to other B. elkanii genomes, even though this is not mandatory for divergent nod gene arrangement (Vázquez et al., 1991Vázquez M, Dávalos A, de las Peñas A, Sánchez F and Quinto C (1991) Novel organization of the common nodulation genes in Rhizobium leguminosarum bv. phaseoli strains. J Bacteriol 173:1250-1258.). As in most rhizobia, expression of the nod/nol operon in Bradyrhizobium is dependent on distinct flavonoid compounds produced and secreted by each leguminous host plant, with the quantity and spectrum of these molecules varying according to the age and physiological state of the plant (Phillips and Streit, 1996Phillips DA and Streit WR (1996) Legume signals to rhizobial symbionts: A new approach for defining rhizosphere colonization. In: Stacey G and Keen NT (eds) Plant-Microbe Interactions. Vol 1. Springer Science & Business Media, Springer, Berlin, pp 236-271.; Loh and Stacey, 2003Loh J and Stacey G (2003) Nodulation gene regulation in Bradyrhizobium japonicum: A unique integration of global regulatory circuits. Appl Environ Microbiol 69:10-17.; Hassan and Mathesius, 2012Hassan S and Mathesius U (2012) The role of flavonoids in root-rhizosphere signalling: Opportunities and challenges for improving plant-microbe interactions. J Exp Bot 63:3429-3444.).

Regulation of the nod/nol operon in B. diazoefficiens and B. elkanii is essentially under control of the same elements, namely, the nod regulon that consists of nodD1, nodD2, nolA and nodVW genes, all of them present in the genomes of both species. The pattern of gene organization observed for nodD1 and nodD2, both of which were invariably located close to each other and in opposite orientation to the nod operon in Bradyrhizobium genomes, reinforced their importance as key components in the core regulatory mechanism of operonic nod gene expression. Indeed, the products of nodD genes are assumed to be fundamental in coordinating structural nod genes in many symbionts, including Bradyrhizobium species (Mulligan and Long, 1989Mulligan JT and Long SR (1989) A family of activator genes regulates expression of Rhizobium meliloti nodulation genes. Genetics 122:7-18.; Honma et al., 1990Honma MA, Asomaning M and Ausubel FM (1990) Rhizobium meliloti nodD genes mediate host-specific activation of nodABC. J Bacteriol 172:901-911.; Loh and Stacey, 2003Loh J and Stacey G (2003) Nodulation gene regulation in Bradyrhizobium japonicum: A unique integration of global regulatory circuits. Appl Environ Microbiol 69:10-17.; del Cerro et al., 2015del Cerro P, Rolla-Santos AAP, Gomes DF, Marks BB, Pérez-Montano F, Rodríguez-Carvajal MA, Nakatani AS, Gil-Serrano A, Megías M, Ollero FJ, et al. (2015) Regulatory nodD1 and nodD2 genes of Rhizobium tropici strain CIAT 899 and their roles in the early stages of molecular signaling and host-legume nodulation. BMC Genomics 16:251.). For instance, in B. diazoefficiens products of the nodD1 and nodD2 regulatory genes orchestrate nodYABCSUIJnolMNOnodZ operon expression by activation or repression, respectively (Loh and Stacey, 2003Loh J and Stacey G (2003) Nodulation gene regulation in Bradyrhizobium japonicum: A unique integration of global regulatory circuits. Appl Environ Microbiol 69:10-17.). Such a regulatory system suggests a common mechanism that may apply for B. elkanii genomes with conserved nod gene organization.

Besides their divergent transcriptional orientation, components of the LysR-type transcriptional regulator (LTTR) family, such as NodD1 and NodD2, also exhibit another characteristic feature, namely, autoregulation (Maddocks and Oyston, 2008Maddocks SE and Oyston PCF (2008) Structure and function of the LysR-type transcriptional regulator (LTTR) family proteins. Microbiology 154:3609-3623.), which ultimately indicates that expression of the associated operon largely depends on the regulation of these regulatory genes themselves. Since B. diazoefficiens nodD mutants retain a large marginal ability to nodulate soybean plants, alternative transcriptional activators were proposed to take part in this process. Indeed, the identification of a two-component regulatory system consisting of NodV and NodW, products of nodVW, sheds some light on this paradox and provides a suitable explanation. In such a system, the sensor kinase component NodV can detect the environmental stimulus (such as a specific isoflavone) resulting in its autophosphorylation and subsequent signal transduction by transfer of the phosphoryl group to its cognate response regulator protein NodW, which in turn is then able to activate its target nod operon (Loh et al., 1997Loh J, Garcia M and Stacey G (1997) NodV and NodW, a second flavonoid recognition system regulating nod gene expression in Bradyrhizobium japonicum. J Bacteriol 179:3013-3020.). Further evidence that Bradyrhizobium nodVW products are essential for efficient nodulation of mung bean [Vigna radiata (L.) Wilezek], cowpea [Vigna unguiculata (L.) Walp.] and siratro [Macroptilium atropurpureum (Moc. & Sessé ex DC.) Urb.], but are not required for soybean, suggests a host-specific role for these proteins, probably through recognition of specific flavonoid inducers produced by the host plants in response to the correlate sensor component NodV (Göttfert et al., 1990Göttfert M, Grob P and Hennecke H (1990) Proposed regulatory pathway encoded by the nodV and nodW genes, determinants of host specificity in Bradyrhizobium japonicum. Proc Natl Acad Sci U S A 87:2680-2684.; Sanjuan et al., 1994Sanjuan J, Grob P, Göttfert M, Hennecke H and Stacey G (1994) NodW is essential for full expression of the common nodulation genes in Bradyrhizobium japonicum. Mol Plant Microbe Interact 7:364-369.). As shown here, we identified this same system in B. elkanii genomes, suggesting that both species possess an alternative pathway for nod operon activation and possibly a strategy for broadening their respective host ranges.

Whereas successful soybean nodulation depends on timely expression of the structural nod gene at the right place and in suitable amounts, it is reasonable to assume that negative regulation also occurs, in addition to activation. Although this negative control in B. diazoefficiens is mediated by the nodD2 product that operates as a nod operon repressor as mentioned before, this process is assisted by the nolA gene product, which establishes an additional level of regulation by inducing nodD2 expression under appropriate conditions. In other words, the product of nolA acts as an transcriptional inducer of the repressor nodD2 (Garcia et al., 1996Garcia M, Dunlap J, Loh J and Stacey G (1996) Phenotypic characterization and regulation of the nolA gene of Bradyrhizobium japonicum. Mol Plant Microbe Interact 9:625-636.). As already mentioned, NolA is a member of the MerR family of transcriptional regulators that activates transcription upon binding to specific DNA motifs and consequently induces DNA binding leading to its appropriate alignment for RNA polymerase positioning and subsequent transcriptional activation (Ansari et al., 1992Ansari AZ, Chael ML and O'Halloran TV (1992) Allosteric underwinding of DNA is a critical step in positive control of transcription by Hg-MerR. Nature 355:87-89.; Philips et al., 2015Philips SJ, Canalizo-Hernandez M, Yildirim I, Schatz GC, Mondragon A and O'Halloran TV (2015) TRANSCRIPTION. Allosteric transcriptional regulation via changes in the overall topology of the core promoter. Science 349:877-881.). Moreover, nolA has the peculiar feature of encoding for the three proteins NolA1, NolA2 and NolA3 that originate from three in-frame ATG start codons, with NolA1 controlling the expression of the other two and being involved in the activation of nodD2 (Loh et al., 1999Loh J, Stacey MG, Sadowsky MJ and Stacey G (1999) The Bradyrhizobium japonicum nolA gene encodes three functionally distinct proteins. J Bacteriol 181:1544-1554.). This mechanism is also shared with high homology by the B. elkanii genomes and, in most cases, shows a very similar gene organization and transcription to nodD1 and nodD2, as previously demonstrated (Dobert et al., 1994Dobert RC, Breil BT and Triplett EW (1994) DNA sequence of the common nodulation genes of Bradyrhizobium elkanii and their phylogenetic relationship to those of other nodulating bacteria. Mol Plant Microbe Interact 7:564-572.). This finding indicates tight nod gene regulation in this species as well.

cis-regulatorynod elements and trans-acting factors

Considering the nature of genomic elements present in the nod operon and especially in the nod regulon of B. diazoefficiens and their marked similarity with those in the B. elkanii genomes, both in terms of structure and organization, we hypothesized that major phenotypic differences in LCO production and secretory patterns in these two species could occur at some transcriptional or post-transcriptional regulatory level. Accordingly, in the presence of the corresponding flavonoid compounds, activated NodD proteins would specifically bind to conserved cis-regulatory elements on bacterial DNA, namely, the nod boxes. These regulatory structures basically consist of a 47-bp conserved region containing the nod box consensus sequence with the palindromic ATC-N9-GAT motif and are located upstream to the nod operon and control its expression (Rostas et al., 1986Rostas K, Kondorosi E, Horvath B, Simoncsits A and Kondorosi A (1986) Conservation of extended promoter regions of nodulation genes in Rhizobium. Proc Natl Acad Sci U S A 83:1757-1761.; Nieuwkoop et al., 1987Nieuwkoop AJ, Banfalvi Z, Deshmane N, Gerhold D, Schell MG, Sirotkin KM and Stacey G (1987) A locus encoding host range is linked to the common nodulation genes of Bradyrhizobium japonicum. J Bacteriol 169:2631-2638.; Goethals et al., 1992Goethals K, Van Montagu M and Holsters M (1992) Conserved motifs in a divergent nod box of Azorhizobium caulinodans ORS571 reveal a common structure in promoters regulated by LysR-type proteins. Proc Natl Acad Sci U S A 89:1646-1650.). The conservation of such cis-regulatory elements and their essentiality in many rhizobia species studied so far strongly suggests that NodD1 is central in nod operon gene expression in these species, including Bradyrhizobium (Spaink et al., 1987Spaink HP, Wijffelman CA, Pees E, Okker RJH and Lugtenberg BJJ (1987) Rhizobium nodulation gene nodD as a determinant of host specificity. Nature 328:337-340.). The nod operon activation by NodD1 in B. elkanii genomes is no exception to this rule and seems to be under a similar, if not identical, regulatory mechanism since the same genetic features, highly conserved in relation to B. diazoefficiens USDA 110, are observed in this context.

Besides the activation of nod gene expression, an autoregulatory function unique to Bradyrhizobium is assigned to nodD1 by its own product NodD1 upon binding the same isoflavone molecules that activate the nod operon. Due to divergent transcriptional orientation, the nod box sequences are located upstream from the nod operon containing the structural nod genes and the nodD1/nodD2 regulatory genes. This location of nod boxes in between the nod operon and nodD1/nodD2 within the genome complicates the analysis of their regulation. Although autoregulation by NodD1 is also accomplished by a DNA-binding mechanism, this is reportedly achieved by binding to an alternative presumptive nod box-like sequence located upstream to the consensus nod box in B. diazoefficiens USDA 110 (Banfalvi et al., 1988Banfalvi Z, Nieuwkoop A, Schell M, Besl L and Stacey G (1988) Regulation of nod gene expression in Bradyrhizobium japonicum. Mol Gen Genet 214:420-424.; Wang and Stacey, 1991Wang SP and Stacey G (1991) Studies of the Bradyrhizobium japonicum nodD1 promoter: A repeated structure for the nod box. J Bacteriol 173:3356-3365.). Indeed, the conservation of this regulatory sequence in B. elkanii genomes reinforces its proposed function, a conclusion supported by the close proximity of this additional cis-regulatory region to a −10/-35 σ70 potential promoter. Despite similarities such as high sequence conservation and location in the genome, a few differences still exist among the presumed nod box-like sequences in distinct Bradyrhizobium species, e.g., the slightly smaller size and some nucleotide divergence in B. elkanii genomes compared to B. diazoefficiens, that ultimately may affect the DNA-binding process itself.

Based on the nodD1 and nodD2 sequence-homology and their resulting description as members of LysR-type regulators, it seems reasonable to also consider a potential autoregulatory role for NodD2, akin to what can be observed for NodD1. Previous screening of Rhizobium japonicum USDA 191 DNA sequences for the presence of a nod box revealed that only the nodD1 promoter region showed high sequence homology to other nod boxes, while no extensive similar motifs were identified upstream to nodD2, even though several alignments of up to four homologous base pairs were observed (Appelbaum et al., 1988Appelbaum ER, Thompson DV, Idler K and Chartrain N (1988) Rhizobium japonicum USDA 191 has two nodD genes that differ in primary structure and function. J Bacteriol 170:12-20.). Although no complete nod box has been found in the promoter region of nodD2 in Bradyrhizobium genomes, a remarkably conserved sequence containing only one copy of the palindrome ATC-N9-GAT, characteristic of a nod box sequence structure, was identified in some B. elkanii genomes. Curiously, the same genomes that contained this palindromic sequence also displayed a potential −10/-35 σ70 binding site. Since this relationship has not been reported before, the precise biological significance of this findings is still unclear and may represent another mechanism (in addition to NolA) by which nodD2 transcription is regulated in this species; this question deserves further investigation.

The current model for nodD2 expression in B. diazoefficiens essentially considers the induction of transcription by NolA upon binding to a putative NolA binding-site present upstream to nodD2 (Garcia et al., 1996Garcia M, Dunlap J, Loh J and Stacey G (1996) Phenotypic characterization and regulation of the nolA gene of Bradyrhizobium japonicum. Mol Plant Microbe Interact 9:625-636.; Loh and Stacey, 2003Loh J and Stacey G (2003) Nodulation gene regulation in Bradyrhizobium japonicum: A unique integration of global regulatory circuits. Appl Environ Microbiol 69:10-17.). Sequence homology has shown that NolA is a member of the MerR transcriptional regulator protein family that is known to activate suboptimal σ70-dependent promoters through protein-dependent DNA distortion that ultimately provides an appropriate alignment of −10/-35 σ70 and the correct positioning of RNA polymerase in relation to the respective promoter (Brown et al., 2003Brown NL, Stoyanov JV, Kidd SP and Hobman JL (2003) The MerR family of transcriptional regulators. FEMS Microb Rev 27:145-163.). Surprisingly, as shown here, preliminary screening for a potential NolA binding site upstream to nodD2 found no corresponding conserved cis-element in this region, contrary to current literature.

NodD protein sequences and structure conservation

In addition to cis-regulatory regions in the bacterial genome, trans-acting factors can also affect transcriptional regulation. The high degree of protein conservation observed for NodD1 and NodD2 among Bradyrhizobium genomes highlighted the evolutionary importance of the mechanism by these regulators operate. Indeed, the NodD amino acid sequence showed extensive conservation that ranged from > 90% sequence similarity among NodD proteins from organisms of the same genus to < 50% for NodD sequences from distantly related organisms (Göttfert et al., 1992Göttfert M, Holzhäuser D, Bäni D and Hennecke H (1992) Structural and functional analysis of two different nodD genes in Bradyrhizobium japonicum USDA110. Mol Plant Microbe Interact 5:257-265.). Although the global sequence similarity for NodD can vary considerably among different organisms, a general feature is the high-to-low conservation in the amino acid sequence within the protein from the amino-terminus to the carboxy-terminus of the polypeptide. This structural pattern occurs in many rhizobia species and is consistent with the function attributed to each domain of the protein (Burn et al., 1987Burn J, Rossen L and Johnston AWB (1987) Four classes of mutations in the nodD gene of Rhizobium leguminosarum biovar. viciae that affect its ability to autoregulate and/or activate other nod genes in the presence of flavonoid inducers. Genes Dev 1:456-464., 1989Burn JE, Hamilton WD, Wootton JC and Johnston AW (1989) Single and multiple mutations affecting properties of the regulatory gene nodD of Rhizobium. Mol Microbiol 3:1567-1577.; Horvath et al., 1987Horvath B, Bachem CW, Schell J and Kondorosi A (1987) Host-specific regulation of nodulation genes in Rhizobium is mediated by a plant-signal, interacting with the nodD gene product. EMBO J 6:841-848.).

The amino-terminal region, the most conserved part of the protein, anchors the HTH DNA-binding motif responsible for recognition of the nod box cis-elements in bacterial DNA. Despite this higher conservation in sequence similarity, subtle amino acid substitutions nevertheless occur among B. elkanii genomes and may ultimately modify the affinity of this motif for the DNA, thereby altering the pattern of binding and consequent activity of these regulators. On the other hand, the NodD carboxy-terminus region is believed to be primarily involved in protein oligomerization, an important aspect since it is generally accepted that NodD1 is active in a tetrameric form that arises from the cognate homodimers (Peck et al., 2013Peck MC, Fisher RF, Bliss R and Long SR (2013) Isolation and characterization of mutant Sinorhizobium meliloti NodD1 proteins with altered responses to luteolin. J Bacteriol 195:3714-3723.). Multiple weak interactions occur among non-specific amino acids in this region of NodD proteins and may contribute to oligomerization (Ezezika et al., 2007Ezezika OC, Haddad S, Neidle EL and Momany C (2007) Oligomerization of BenM, a LysR-type transcriptional regulator: Structural basis for the aggregation of proteins in this family. Acta Crystal Sect F 63:361-368.; Peck et al., 2013Peck MC, Fisher RF, Bliss R and Long SR (2013) Isolation and characterization of mutant Sinorhizobium meliloti NodD1 proteins with altered responses to luteolin. J Bacteriol 195:3714-3723.). As observed in other rhizobia, the NodD carboxy-terminus region was clearly less conserved in the B. elkanii genomes. Even though this lower sequence similarity did not markedly affect protein structure, it may play a role in oligomerization.

Based on the information available for the B. diazoefficiens USDA 110 reference genome, the similarities between B. diazoefficiens and B. elkanii, and the results obtained from this comparative genomic analysis, an extension in the model of nod gene regulation proposed for B. diazoefficiens can be applied to B. elkanii with some new added features (Figure 5). Besides the activity of all known elements in the core mechanism of nod operon regulation by the already discussed nod regulon, the presence of a partial nod box sequence upstream to nodD2 suggests a putative NodD1 binding site in this region, with a possible regulatory effect on nodD2 as a transcriptional activation. Such a mechanism may possibly function under higher concentrations of NodD1 as a consequence of the autoregulation of nodD1 expression by its respective product. Under these conditions, a surplus of NodD1 may be “trapped” by this partial nod box sequence; NodD1 may show lower affinity for the regulator protein relative to the conserved consensus nod box sequence and ultimately activate transcription of the repressor protein NodD2 to balance expression of the nod operon.

Figure 5
A model illustrating the modulation of nod gene expression proposed for B. elkanii. Expression of the nod operon containing the structural nod genes is regulated by regulatory nod genes (nod regulon) in the presence of the respective flavonoid inducer. Transcriptional activation of the nod operon is mediated by NodD1 and NodVW, resulting in the biosynthesis of lipochitooligosaccharides. Negative regulation of the nod operon is due to the action of NodD2 and indirectly by NolA. Based on Loh and Stacey (2003)Loh J and Stacey G (2003) Nodulation gene regulation in Bradyrhizobium japonicum: A unique integration of global regulatory circuits. Appl Environ Microbiol 69:10-17..

The prediction of highly conserved structures between NodD1 and NodD2 and their structural alignment suggested the possible formation of NodD1-NodD2 heterodimers since their protein structures are very similar, but no specific amino acid residues responsible for oligomerization have yet been identified. If NodD1-NodD2 heterodimers are formed in Bradyrhizobium, they may well affect the NodD1-driven regulation in this symbiont, in addition to interfering with nod operon gene expression. Although the data are suggestive of these events, there is still no conclusive evidence for this hypothesis and experimental confirmation of such mechanisms is required to prove the validity of this model.

Our knowledge of the genetic control of nodulation and our understanding of the mechanisms involved in the early events of a new symbiotic relationship have undoubtedly expanded considerably in the last two decades. Although a general framework for nodulation and symbiosis has not yet been developed for all diazotrophic microorganisms, given their genetic and phenotypic diversity, comparative studies may nevertheless help to clarify these processes and provide some useful preliminary evidence for future investigations. As microbial genome sequencing projects are quickly delivering an ever-increasing amount of data, comparative analyses of cis-elements and their cognate trans-acting factors could provide new insights into the activation and/or repression of nod gene expression.

This work has identified several similarities between B. elkanii genomes and the closely related B. diazoefficiens. Based on these similarities, it was possible to identify and correlate functional mechanisms and key elements that play an essential role in regulating nod gene expression. In addition, new genomic features that had not been clearly explored before, some of which were unique to certain B. elkanii genomes, have raised new questions for future research.

Acknowledgments

This work was financed by a grant and fellowships from the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, Brazil) and the Instituto Nacional de Ciência e Tecnologia (INCT) da Fixação Biológica de Nitrogênio (Brazil).

References

  • Ansari AZ, Chael ML and O'Halloran TV (1992) Allosteric underwinding of DNA is a critical step in positive control of transcription by Hg-MerR. Nature 355:87-89.
  • Appelbaum ER, Thompson DV, Idler K and Chartrain N (1988) Rhizobium japonicum USDA 191 has two nodD genes that differ in primary structure and function. J Bacteriol 170:12-20.
  • Aziz RK, Bartels D, Best AA, DeJongh M, Disz T, Edwards RA, Formsma K, Gerdes S, Glass E, Kubal M, et al. (2008) The RAST Server: Rapid annotations using subsystems technology. BMC Genomics 9:75.
  • Banfalvi Z, Nieuwkoop A, Schell M, Besl L and Stacey G (1988) Regulation of nod gene expression in Bradyrhizobium japonicum Mol Gen Genet 214:420-424.
  • Bendtsen JD, Jensen LJ, Blom N, von Heijne G and Brunak S (2004) Feature-based prediction of non-classical and leaderless protein secretion. Prot Eng Des Select 17:349-356.
  • Bendtsen J, Nielsen H, Widdick D, Palmer T and Brunak S (2005) Prediction of twin-arginine signal peptides. BMC Bioinformatics 6:167.
  • Berman H, Henrick K, Nakamura H and Markley JL (2007) The worldwide Protein Data Bank (wwPDB): Ensuring a single, uniform archive of PDB data. Nucleic Acids Res 35:301-303.
  • Bernstein FC, Koetzle TF, Williams GJB, Meyer Jr EF, Brice MD, Rodgers JR, Kennard O, Shimanouchi T and Tasumi M (1977) The Protein Data Bank: A computer-based archival file for macromolecular structures. J Mol Biol 112:535-542.
  • Biasini, M, Bienert S, Waterhouse A, Arnold K, Studer G, Schmidt T, Kiefer F, Cassarino TG, Bertoni M, Bordoli L, et al. (2014) SWISS-MODEL: Modelling protein tertiary and quaternary structure using evolutionary information. Nucleic Acids Res 42:252-258.
  • Brown NL, Stoyanov JV, Kidd SP and Hobman JL (2003) The MerR family of transcriptional regulators. FEMS Microb Rev 27:145-163.
  • Burn J, Rossen L and Johnston AWB (1987) Four classes of mutations in the nodD gene of Rhizobium leguminosarum biovar. viciae that affect its ability to autoregulate and/or activate other nod genes in the presence of flavonoid inducers. Genes Dev 1:456-464.
  • Burn JE, Hamilton WD, Wootton JC and Johnston AW (1989) Single and multiple mutations affecting properties of the regulatory gene nodD of Rhizobium. Mol Microbiol 3:1567-1577.
  • Campos BC, Hungria M and Tedesco V (2001) Eficiência da fixação biológica de N2 por estirpes de Bradyrhizobium na soja em plantio direto. Rev Bras Ciênc Sol 25:583-592.
  • Day RB, Loh J, Cohn JR and Stacey G (2000) Signal exchange involved in the establishment of the Bradyrhizobium-legume symbiosis. In: Triplett E (ed) Prokaryotic Nitrogen Fixation, a Model System for the Analysis of a Biological Process. Horizon Scientific Press, Norfolk, pp 385-414.
  • Dayhoff MO and Schwartz RM (1978) A model of evolutionary change in proteins. In: Dayhoff MO (ed) Atlas of Protein Sequence and Structure. Natl Biomed Res Found, Washington DC, pp 345-352.
  • de Castro E, Sigrist CA, Gattiker A, Bulliard V, Langendijk-Genevaux PS, Gasteiger E, Bairoch A and Hulol N (2006) ScanProsite: Detection of PROSITE signature matches and ProRule-associated functional and structural residues in proteins. Nucleic Acids Res 34:362-365.
  • del Cerro P, Rolla-Santos AAP, Gomes DF, Marks BB, Pérez-Montano F, Rodríguez-Carvajal MA, Nakatani AS, Gil-Serrano A, Megías M, Ollero FJ, et al. (2015) Regulatory nodD1 and nodD2 genes of Rhizobium tropici strain CIAT 899 and their roles in the early stages of molecular signaling and host-legume nodulation. BMC Genomics 16:251.
  • Denarie J, Debelle F, Truchet G and Prome J-C (1993) Rhizobium and legume nodulation: A molecular dialogue. In: Palacios R, Mora J and Newton W (eds) New Horizons in Nitrogen Fixation. Vol 17. Current Plant Science and Biotechnology in Agriculture. Springer, Dordrecht, pp 19-30.
  • Dobert RC, Breil BT and Triplett EW (1994) DNA sequence of the common nodulation genes of Bradyrhizobium elkanii and their phylogenetic relationship to those of other nodulating bacteria. Mol Plant Microbe Interact 7:564-572.
  • Edgar RC (2004) MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32:1792-1797.
  • Emanuelsson O, Brunak S, von Heijne G and Nielsen H (2007) Locating proteins in the cell using TargetP, SignalP and related tools. Nat Protocols 2:953-971.
  • Ezezika OC, Haddad S, Neidle EL and Momany C (2007) Oligomerization of BenM, a LysR-type transcriptional regulator: Structural basis for the aggregation of proteins in this family. Acta Crystal Sect F 63:361-368.
  • Garcia M, Dunlap J, Loh J and Stacey G (1996) Phenotypic characterization and regulation of the nolA gene of Bradyrhizobium japonicum Mol Plant Microbe Interact 9:625-636.
  • Gasteiger E, Gattiker A, Hoogland C, Ivanyi I, Appel RD and Bairoch A (2003) ExPASy: The proteomics server for in-depth protein knowledge and analysis. Nucleic Acids Res 31:3784-3788.
  • Goethals K, Van Montagu M and Holsters M (1992) Conserved motifs in a divergent nod box of Azorhizobium caulinodans ORS571 reveal a common structure in promoters regulated by LysR-type proteins. Proc Natl Acad Sci U S A 89:1646-1650.
  • Goodsell DS, Dutta S, Zardecki C, Voigt M, Berman HM and Burley SK (2015) The RCSB PDB “Molecule of the Month”: Inspiring a molecular view of biology. PLoS Biol 13:e1002140.
  • Gottfert M, Lamb JW, Gasser R, Semenza J and Hennecke H (1989) Mutational analysis of the Bradyrhizobium japonicum common nod genes and further nod box-linked genomic DNA regions. Mol Gen Genet 215:407-415.
  • Göttfert M, Grob P and Hennecke H (1990) Proposed regulatory pathway encoded by the nodV and nodW genes, determinants of host specificity in Bradyrhizobium japonicum. Proc Natl Acad Sci U S A 87:2680-2684.
  • Göttfert M, Holzhäuser D, Bäni D and Hennecke H (1992) Structural and functional analysis of two different nodD genes in Bradyrhizobium japonicum USDA110. Mol Plant Microbe Interact 5:257-265.
  • Haas J, Roth S, Arnold K, Kiefer F, Schmidt T, Bordoli L and Schwede T (2013) The Protein Model Portal - A comprehensive resource for protein structure and model information. Database 2013. doi:10.1093/database/bat031.
    » https://doi.org/10.1093/database/bat031
  • Hassan S and Mathesius U (2012) The role of flavonoids in root-rhizosphere signalling: Opportunities and challenges for improving plant-microbe interactions. J Exp Bot 63:3429-3444.
  • Henikoff S, Haughn GW, Calvo JM and Wallace JC (1988) A large family of bacterial activator proteins. Proc Natl Acad Sci U S A 85:6602-6606.
  • Hong GF, Burn JE and Johnston AW (1987) Evidence that DNA involved in the expression of nodulation (nod) genes in Rhizobium binds to the product of the regulatory gene nodD. Nucleic Acids Res 15:9677-9690.
  • Honma MA, Asomaning M and Ausubel FM (1990) Rhizobium meliloti nodD genes mediate host-specific activation of nodABC. J Bacteriol 172:901-911.
  • Horvath B, Bachem CW, Schell J and Kondorosi A (1987) Host-specific regulation of nodulation genes in Rhizobium is mediated by a plant-signal, interacting with the nodD gene product. EMBO J 6:841-848.
  • Hunter S, Apweiler R, Attwood TK, Bairoch A, Bateman A, Binns D, Bork P, Das U, Daugherty L, Duquenne L, et al. (2009) InterPro: The integrative protein signature database. Nucleic Acids Res 37:211-215.
  • Jordan DC (1982) Transfer of Rhizobium japonicum Buchanan 1980 to Bradyrhizobium gen. nov., a genus of slow-growing, root nodule bacteria from leguminous plants. Int J Sys Evol Microb 32:136-139.
  • Käll L, Krogh A and Sonnhammer ELL (2004) A combined transmembrane topology and signal peptide prediction. Meth J Mol Biol 338:1027-1036.
  • Kaneko T, Nakamura Y, Sato S, Minamisawa K, Uchiumi T, Sasamoto S, Watanabe A, Idesawa K, Iriguchi M, Kawashima K, et al. (2002) Complete genomic sequence of nitrogen-fixing symbiotic bacterium Bradyrhizobium japonicum USDA110. DNA Res 9:189-197.
  • Kim O-S, Cho Y-J, Lee K, Yoon S-H, Kim M, Na H, Park S-C, Jeon YS, Lee J-H, Yi H, et al. (2012) Introducing EzTaxon-e: A prokaryotic 16S rRNA gene sequence database with phylotypes that represent uncultured species. Int J Sys Evol Microb 62:716-721.
  • Kober MdV, de Sá ELS, Freire JRJ and Giongo A (2004) Characterization of variants of Bradyrhizobium elkanii and B. japonicum and symbiotic behaviour in soybeans. Ciênc Rural 34:1459-1464.
  • Krogh A, Larsson B, von Heijne G and Sonnhammer ELL (2001) Predicting transmembrane protein topology with a hidden Markov model: Application to complete genomes. J Mol Biol 305:567-580.
  • Loh J and Stacey G (2003) Nodulation gene regulation in Bradyrhizobium japonicum: A unique integration of global regulatory circuits. Appl Environ Microbiol 69:10-17.
  • Loh J, Garcia M and Stacey G (1997) NodV and NodW, a second flavonoid recognition system regulating nod gene expression in Bradyrhizobium japonicum J Bacteriol 179:3013-3020.
  • Loh J, Stacey MG, Sadowsky MJ and Stacey G (1999) The Bradyrhizobium japonicum nolA gene encodes three functionally distinct proteins. J Bacteriol 181:1544-1554.
  • Luka S, Sanjuan J, Carlson RW and Stacey G (1993) nolMNO genes of Bradyrhizobium japonicum are co-transcribed with nodYABCSUIJ, and nolO is involved in the synthesis of the lipo-oligosaccharide nodulation signals. J Biol Chem 268:27053-27059.
  • Maddocks SE and Oyston PCF (2008) Structure and function of the LysR-type transcriptional regulator (LTTR) family proteins. Microbiology 154:3609-3623.
  • Magrane M and Consortium U (2011) UniProt Knowledgebase: A hub of integrated protein data. Database (Oxford) 2011:bar009.
  • Melchiorre M, de Luca M, Gonzalez Anta G, Suarez P, Lopez C, Lascano R and Racca R (2011) Evaluation of bradyrhizobia strains isolated from field-grown soybean plants in Argentina as improved inoculants. Biol Fertil Soils 47:81-89.
  • Minamisawa K (1989) Comparison of extracellular polysaccharide composition, rhizobitoxine production, and hydrogenase phenotype among various strains of Bradyrhizobium japonicum Plant Cell Physiol 30:877-884.
  • Mulligan JT and Long SR (1989) A family of activator genes regulates expression of Rhizobium meliloti nodulation genes. Genetics 122:7-18.
  • Nieuwkoop AJ, Banfalvi Z, Deshmane N, Gerhold D, Schell MG, Sirotkin KM and Stacey G (1987) A locus encoding host range is linked to the common nodulation genes of Bradyrhizobium japonicum J Bacteriol 169:2631-2638.
  • Peck MC, Fisher RF, Bliss R and Long SR (2013) Isolation and characterization of mutant Sinorhizobium meliloti NodD1 proteins with altered responses to luteolin. J Bacteriol 195:3714-3723.
  • Petersen TN, Brunak S, von Heijne G and Nielsen H (2011) SignalP 4.0: Discriminating signal peptides from transmembrane regions. Nat Methods 8:785-786.
  • Philips SJ, Canalizo-Hernandez M, Yildirim I, Schatz GC, Mondragon A and O'Halloran TV (2015) TRANSCRIPTION. Allosteric transcriptional regulation via changes in the overall topology of the core promoter. Science 349:877-881.
  • Phillips DA and Streit WR (1996) Legume signals to rhizobial symbionts: A new approach for defining rhizosphere colonization. In: Stacey G and Keen NT (eds) Plant-Microbe Interactions. Vol 1. Springer Science & Business Media, Springer, Berlin, pp 236-271.
  • Rahman O, Cummings S, Harrington D and Sutcliffe I (2008) Methods for the bioinformatic identification of bacterial lipoproteins encoded in the genomes of Gram-positive bacteria. World J Microbiol Biotechnol 24:2377-2382.
  • Rivas R, Martens M, de Lajudie P and Willems A (2009) Multilocus sequence analysis of the genus Bradyrhizobium. Syst Appl Microbiol 32:101-110.
  • Rostas K, Kondorosi E, Horvath B, Simoncsits A and Kondorosi A (1986) Conservation of extended promoter regions of nodulation genes in Rhizobium. Proc Natl Acad Sci U S A 83:1757-1761.
  • Sadowsky MJ, Cregan PB, Gottfert M, Sharma A, Gerhold D, Rodriguez-Quinones F, Keyser HH, Hennecke H and Stacey G (1991) The Bradyrhizobium japonicum nolA gene and its involvement in the genotype-specific nodulation of soybeans. Proc Natl Acad Sci U S A 88:637-641.
  • Salzberg SL, Delcher AL, Kasif S and White O (1998) Microbial gene identification using interpolated Markov models. Nucleic Acids Res 26:544-548.
  • Sanjuan J, Grob P, Göttfert M, Hennecke H and Stacey G (1994) NodW is essential for full expression of the common nodulation genes in Bradyrhizobium japonicum Mol Plant Microbe Interact 7:364-369.
  • Schlaman HR, Phillips DA and Kondorosi E (1998) Genetic organization and transcriptional regulationof rhizobial nodulation genes. In: Spaink HP, Kondorosi A and Hooykaas P (eds) The Rhizobiaceae: Molecular Biology of Model Plant-Associated Bacteria. Springer, Dordrecht, pp 361-386.
  • Schumpp O and Deakin WJ (2010) How inefficient rhizobia prolong their existence within nodules. Trends Plant Sci 15:189-195.
  • Shindyalov IN and Bourne PE (1998) Protein structure alignment by incremental combinatorial extension (CE) of the optimal path. Protein Eng 11:739-747.
  • Sigrist CJA, Cerutti L, de Castro E, Langendijk-Genevaux PS, Bulliard V, Bairoch A and Hulo N (2010) PROSITE, a protein domain database for functional characterization and annotation. Nucleic Acids Res 38:161-166.
  • Somasegaran P and Hoben HJ (1994) Handbook for Rhizobia: Methods in Legume-Rhizobium Technology. Springer-Verlag, New York, 450 p.
  • Spaink HP, Wijffelman CA, Pees E, Okker RJH and Lugtenberg BJJ (1987) Rhizobium nodulation gene nodD as a determinant of host specificity. Nature 328:337-340.
  • Tamura K, Stecher G, Peterson D, Filipski A and Kumar S (2013) MEGA6: Molecular Evolutionary Genetics Analysis Version 6.0. Mol Biol Evol 30:2725-2729.
  • Tóth K and Stacey G (2015) Does plant immunity play a critical role during initiation of the legume-rhizobium symbiosis? Front Plant Sci 6:401.
  • Vasilas BL and Furhmann JJ (1993) Field response of soybean to nodulation by a rhizobitoxine-producing strain of Bradyrhizobium. Agron J 85:302-305.
  • Vázquez M, Dávalos A, de las Peñas A, Sánchez F and Quinto C (1991) Novel organization of the common nodulation genes in Rhizobium leguminosarum bv. phaseoli strains. J Bacteriol 173:1250-1258.
  • Wang D, Yang S, Tang F and Zhu H (2012) Symbiosis specificity in the legume – Rhizobial mutualism. Cell Microbiol 14:334-342.
  • Wang Q, Garrity GM, Tiedje JM and Cole JR (2007) Naïve Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl Environ Microbiol 73:5261-5267.
  • Wang SP and Stacey G (1991) Studies of the Bradyrhizobium japonicum nodD1 promoter: A repeated structure for the nod box. J Bacteriol 173:3356-3365.
  • Zdobnov EM and Apweiler R (2001) InterProScan – An integration platform for the signature-recognition methods in InterPro. Bioinformatics 17:847-848.
  • Associate Editor: Ana Tereza R. Vasconcelos

Publication Dates

  • Publication in this collection
    31 July 2017
  • Date of issue
    Jul-Sep 2017

History

  • Received
    30 Jan 2017
  • Accepted
    16 Feb 2017
Sociedade Brasileira de Genética Rua Cap. Adelmio Norberto da Silva, 736, 14025-670 Ribeirão Preto SP Brazil, Tel.: (55 16) 3911-4130 / Fax.: (55 16) 3621-3552 - Ribeirão Preto - SP - Brazil
E-mail: editor@gmb.org.br