Next Article in Journal
YfiB: An Outer Membrane Protein Involved in the Virulence of Shigella flexneri
Previous Article in Journal
Virological Characterization of Pigs with Erythema Multiforme
 
 
Correction published on 11 November 2022, see Microorganisms 2022, 10(11), 2229.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Proteogenome of Symbiotic Frankia alni in Alnus glutinosa Nodules

1
Ecologie Microbienne, CNRS, UMR5557, Université Lyon 1, Université de Lyon; INRA, UMR1418, 7330 Villeurbanne, France
2
Département Médicaments et Technologies pour la Santé (DMTS), CEA, INRAE, Université Paris-Saclay, SPI, 30200 Bagnols sur Cèze, France
*
Authors to whom correspondence should be addressed.
Submission received: 24 February 2022 / Revised: 14 March 2022 / Accepted: 15 March 2022 / Published: 18 March 2022 / Corrected: 11 November 2022
(This article belongs to the Section Plant Microbe Interactions)

Abstract

:
Omics are the most promising approaches to investigate microbes for which no genetic tools exist such as the nitrogen-fixing symbiotic Frankia. A proteogenomic analysis of symbiotic Frankia alni was done by comparing those proteins more and less abundant in Alnus glutinosa nodules relative to N-replete pure cultures with propionate as the carbon source and ammonium as the nitrogen-source. There were 250 proteins that were significantly overabundant in nodules at a fold change (FC) ≥ 2 threshold, and 1429 with the same characteristics in in vitro nitrogen-replete pure culture. Nitrogenase, SuF (Fe–Su biogenesis) and hopanoid lipids synthesis determinants were the most overabundant proteins in symbiosis. Nitrogenase was found to constitute 3% of all Frankia proteins in nodules. Sod (superoxide dismutase) was overabundant, indicating a continued oxidative stress, while Kats (catalase) were not. Several transporters were overabundant including one for dicarboxylates and one for branched amino acids. The present results confirm the centrality of nitrogenase in the actinorhizal symbiosis.

1. Introduction

Frankia is a genus containing soil actinobacteria that establish a nitrogen-replete root nodule symbiosis with 24 genera and hundreds of species of dicotyledons [1]. The best known of these is Alnus glutinosa L., the designated type, which is representative of the genus both for morphological characters and evolutionary rate [2].
Hundreds of Frankia strains have been isolated from nodules of the different species and classified according to their physiology [3], host specificity [4] or phylogeny [5]. More recently, many species have been described—the first of which was Frankia alni [6]—that can nodulate Alnus spp. and Myrica spp. Its genome has been deciphered and compared to other species [5] and found to contain no symbiosis island, with all genes related to symbiosis present in several unlinked clusters [7].
This symbiosis between F. alni and Alnus spp. is initiated by Ca-spiking [8], followed by the synthesis of a root-hair-deforming factor [9] that allows the deformation, entrapment of hyphae and penetration into the cortical cells [10]; the formation of a pre-nodule and the emergence of a mature nitrogen-fixing nodule. On the plant side, the mechanisms involved appear similar to those present in legumes, with several elements of the common symbiotic cascade involved, such as the nuclear transcription factor NIN (nodule inception) [11] and the SYMR (symbiosis) kinase [12]. The genomics of symbiotic and non-symbiotic plant phylogenetic neighbors has shown that the common symbiotic determinants were present in all leguminous and actinorhizal lineages and that loss of symbiotic capacity was accompanied by the loss of NIN and RPG (rhizobium polar growth) [13].
On the microbial side, less is known since no genetic tools could be developed despite several attempts, hampering mutagenesis and complementation studies. The genomes revealed no canonical nod genes [5] except in two cl2 strains [14,15] and one cl3 strain [16] where two gene clusters are found, one with nodAB1 and the other with nodB2CIJ; however, these were not detected by proteomics of F. soli exposed to root exudates from compatible Elaeagnus angustifolia and incompatible hosts [17] even though the mRNAs had been detected in Datisca nodules [14]. Transcriptomics of Frankia in Alnus nodules [7] has identified nif (nitrogenase), hop (hopanoid synthesis), suf (FeS cluster synthesis) and hup (hydrogenase uptake) as upregulated but has shed no light on the molecular dialogue between the two partners. The determinants and structure of the root deforming factor [10], thus, remain elusive. We also know through omics [18] of early rhizospheric interactions that a cellulase/cellulose synthase cluster is upregulated even though F. alni cannot grow using glucose as a carbon source, which is evocative of a local weakening of the hair cell wall and a hardening of the hyphal tip to facilitate entry into root tissues. We also know that the auxin PAA is synthesized in nodules and in pure culture—an auxin that when applied onto roots at 10−5 M causes emergence of stunted swollen secondary roots that are similar to nodules [19]. It is known that the plant synthesizes peptides that bind to and modify the porosity of vesicles [20] and a non-comparative survey of field nodules has shown nitrogenase, the protein that reduces dinitrogen into ammonium and tricarboxylate cycle (TCA) proteins to be abundant [21].
Other in vitro proteomic studies have been done of Frankia with/without ammonium [22] or under osmotic stress [23] and have shown TCA involvement, stress determinants and various regulators.
We undertook the present study to characterize the proteomic response of Frankia alni as it forms mature 21 dpi nodules on A. glutinosa roots.

2. Materials and Methods

2.1. Plant and Bacterial Material

Frankia alni strain ACN14a [24] cells were maintained in BAP medium [25] with 5 mM propionate as the carbon source and 5 mM NH4+ as the nitrogen source buffered to pH 6.5.
Alnus glutinosa seeds were harvested from a tree growing in Lyon, France, used previously [7]. They were grown as before with some modifications: seedlings were transferred to Fahraeus’s solution [26] in opaque plastic pots (eight seedlings/pot) and grown for four weeks with 0.5 g·L−1 KNO3, followed by one week without KNO3 before inoculation with F. alni [24]. To inoculate seedlings, Frankia cells were grown in BAP-PCM medium (4 × 250 mL) until log-phase [27]. Cells were collected by centrifugation, washed twice with sterile ultra-pure water and resuspended in 500 mL of Fahraeus’s solution without KNO3. Cell cultures were homogenized by syringing though a 21G needle. The Frankia cell suspensions were applied onto plant roots (symbiotic condition). After 21 days, mature nodules were harvested and ground in liquid nitrogen.
As a reference, F. alni cells were inoculated after syringing with a series of needles (21G, 23G, 25G, 27G) and grown for 10 days (corresponding to the end of the exponential phase) in 250 mL of BAP medium with ammonium (5 mM) in agitated 500 mL Erlenmeyer flasks [25] buffered to pH 6.5. No vesicles could be found.

2.2. Proteome Characterization

Each sample was dissolved in LDS 1X buffer (Invitrogen, Carlsbad, CA, USA) with 100 µL of LDS 1X per 30 mg of pellet. The solutions were warmed at 99 °C for 5 min and subjected to sonication in an ultrasonic bath for 5 min. Each sample was transferred into a 2 mL Precellys (Bertin Technologies, Montigny-le-Bretonneux, France) tube containing 200 mg of glass beads and subjected to three cycles of grinding for 20 s, followed by 30 s pauses. Samples were centrifuged for 40 s at 16,000× g. The resulting supernatants were transferred into Eppendorf tubes and heated for 10 min at 99 °C. Samples were subjected to a short SDS-PAGE migration and processed as previously described [28]. Briefly, the whole proteome was extracted as a single polyacrylamide band, reduced with dithiothreitol, treated with iodoacetamide and proteolyzed with sequencing-grade trypsin (Roche, Basel, Switzerland) in the presence of 0.01% of proteaseMAX detergent (Promega, Madison, WI, USA). The resulting peptides were analyzed with an ESI-Q Exactive HF mass spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) coupled to an Ultimate 3000 176 RSL Nano LC System (Thermo). A volume of 10 µL of peptides was injected onto a reverse-phase Acclaim PepMap 100 C18 column (3 µm, 100 Å, 75 µm id × 500 mm) and resolved at a flow rate of 0.2 µL/min with a 60 min gradient of CH3CN (2.5% to 40%) in the presence of 0.1% HCOOH. The tandem mass spectrometer was operated with a Top20 strategy in data-dependent mode. Only peptide molecular ions with double or triple positive charges were selected for fragmentation with a dynamic exclusion of 10 s as previously described [29]. Tandem mass spectrometry spectra were interpreted using the MASCOT 2.2.04 software (Matrix Science, London, UK) with standard parameters. Proteins were quantified based on their spectral counts, and their abundances were normalized with total spectral counts of all proteins identified as belonging to Frankia for comparing distinct conditions. Proteome comparison between conditions was done according to the TFold module from the PatternLab software (www.patternlabforproteomics.org/; last access 7 January 2022).

2.3. Proteome Data

The mass spectrometry proteomics data were deposited in the ProteomeXchange Consortium via the PRIDE partner repository (www.ebi.ac.uk/pride/archive/; last access 16 December 2021) with the dataset identifier PXD030468 and Project DOI 10.6019/PXD030468.

3. Results

The three biological replicates of symbiotic Frankia alni overproduced at a fold change of ≥2250 proteins (Supplementary Table S1) using a nitrogen-replete propionate-fed pure culture as reference, of which 100 had an FC ≥ 4.38 (Table 1). These 100 proteins signing for the specificity of symbiotic state account for 17.9% of the total Frankia proteins in nodules, based on their cumulated normalized spectral abundance factors. Conversely, there were 1489 under-produced proteins at an FC ≤ 0.5. NifH (FRAAL6813) was the most overabundant protein in this condition with an FC of 291.3.
The most underabundant was a cAMP-binding membrane-bound transcriptional regulator (FRAAL6506) at an FC of 0.01 (Supplementary Table S1) in synton (set of genes with a conserved order) with a geosmin synthase gene. There were 2983 proteins detected that belonged to Frankia alni. There were also 68 overabundant proteins in nodules and 30 overabundant ones in pure cultures that did not meet the p-value criterion (≤0.05). There were 729 proteins that were not certified by mass spectrometry with the validation of at least two distinct peptides in this dataset. The overabundant protein coding genes were scattered around the genome (Figure 1).
Among F. alni proteins, the nitrogenase proteins were the most overabundant with 7 among the 10 highest using as reference a nitrogen-replete pure culture. In addition to NifH, NifK (FRAAL6811) at an FC of 279 and NifD (FRAAL6812) at an FC of 86.17, the nif cluster comprises 17 genes from FRAAL6798 (korB) to FRAAL6814 (nifV), 13 of which were identified accounting for 3% of the cumulated NSAF in nodules.
Proteins associated to symbiotic life such as the Suf cluster (FRAAL4558-FRAAL4564, FC = 5.4 − 3.39) and the squalene lipids cluster (FRAAL1427-FRAAL1435, FC = 8.4) were among those detected as overproduced; however, the Hup clusters were not detected (FRAAL1822-FRAAL1832; FRAAL2388-FRAAL2402). Proteins involved in the upstream non-mevalonate pathway for hopanoids biosynthesis such as Dxs (FRAAL2088, FC = 2.34), Dxr (FRAAL5774, FC = 10.54), IdsA (FRAAL1431, FC = 7.75) and Idi (FRAAL6504, FC = 2.21) were also overproduced.
Other lipid metabolism proteins were overabundant such as several unlinked acetyl-CoA acetyl transferases (FRAAL4934, FC = 12.91; FRAAL2618, FC = 11.83; FRAAL0409; FC = 4.53; 3973, FC = 2.95), a fatty acid-CoA ligase (FRAAL3361, FC = 7.43), an acyl-CoA dehydrogenase (FRAAL2835, FC = 6.71) and a diacylglycerol O-transferase (FRAAL3613, FC = 5.67).
Oxidative stress proteins such as SodF (FRAAL4337, FC = 3.85) and its bacterioferritin neighbor (FRAAL4338, FC = 8.49), glutathione peroxidase (FRAAL1783, FC = 3.55), two nitroreductases (FRAAL4025, FC = 44.33; FRAAL3236, FC = 18.83) and rubrerythrin (FRAAL5116, FC = 3.08) were overabundant. Other stress proteins were underabundant such as KatA (FRAAL3167, FC = 0.09) and glutathione synthase (FRAAL4656, FC = 0.03) or were not identified such as KatB (FRAAL3889).
Several TCA cycle proteins such as succinyl-CoA synthase (FRAAL1156, FC = 6.14), aconitate hydratase (FRAAL2064, FC = 5.44), malate dehydrogenase (FRAAL5612, FC = 4.37) or fumarase (FRAAL6152, FC = 3.96) were more abundant in nodules, but several others were not detected. There were 15 ribosomal proteins overabundant at an FC > 2 (FRAAL1069, FRAAL1068, FRAAL1097, etc.) as well as other ancillary proteins such as a ribosome-recycling factor (FRAAL5778, FC = 3.35).
Other carbohydrate metabolism proteins were overabundant such as glycosyl transferases (FRAAL2049, FC = 15.5; FRAAL4552, FC = 3.5; FRAAL5400, FC = 2.85), a sugar epimerase (FRAAL6795, FC = 10.67), a ribose isomerase (FRAAL1862, FC = 9.07), a sugar-binding protein (FRAAL5914, FC = 3.49) and an enolase (FRAAL6233, FC = 3.46).
Ammonia assimilation proteins were overabundant such as the nitrogen regulatory protein P-II (FRAAL0904, FC = 3.27), glutamine-synthetase adenylyltransferase (FRAAL5164; FC = 2.13), glutamate-1-semialdehyde aminotransferase FRAAL0997, FC = 5.93) and a GS (FRAAL6426; FC = 1.8), but several others such as GSII were not detected or were underabundant such as GSI (FRAAL5161, FC = 0.01) and GOGAT (FRAAL4964-65, FC = 0.08 − 0.05).
A cell-division control ATPase was overabundant FRAAL5516 (FC = 89), while other proteins with related functions were less abundant (FRAAL6588, cell division protein FtsH FC = 0.22) or not detected.
Chaperones were overabundant such as a disaggregation chaperone (FRAAL2601, FC = 4.22) and several GroEL homologs (FRAAL6640, FC = 2.64; FRAAL1699, FC = 2.16; FRAAL1133, FC = 2.2; FRAAL6486, FC = 1.94). Other overabundant proteins with related functions were FRAAL4269 (Clp Protease), FRAAL6701 (GroEL) and FRAAL5348 (cold shock), with the corresponding genes scattered on the chromosome.
Some transporters were overabundant in nodules such as a molybdenum transporter (FRAAL1673, FC = 64.17), a BCA transporter (FRAAL3827, FC = 5.18) and a C4-dicarboxylate transporter (FRAAL1390, FC = 3.42). A phosphate transporter (FRAAL6535) was also overabundant (FC = 2.09), while many were underabundant.
Some regulators were overabundant in nodules such as a Tet-R (FRAAL0868, FC = 33.09) and a Sarp-R (FRAAL0301, FC = 12.38). A Bhl-R (FRAAL5827, FC = 9.76), a MarR (FRAAL0711, FC = 8.76) and a luxR (FRAAL4898, FC = 7.67) were also overabundant, while many were underabundant. An anti-sigma factor (FRAAL5991, FC = 5.12) in synton with a sporulation-specific sigma factor was overabundant.
Most secondary metabolite synthesis proteins were not detected or underabundant except two PKS (FRAAL5096, FC = 7.22; FRAAL4072, FC = 4.56) and a bacteriocin (FRAAL4503, FC = 2.16).
A gamma-aminobutyraldehyde dehydrogenase (FRAAL6022, FC = 3.73) was overabundant. This enzyme transforms 4-aminobutanal into 4-aminobutanoate (GABA).
The Kdp proteins (FRAAL5299-FRAAL5305), the cellulase synton proteins (FRAAL4954-FRAAL4959) and the acyl-dehydrogenases (FRAAL1659-FRAAL1670) were not identified.
Almost all carboxylases were underabundant such as three acetyl-CoA carboxylases (FRAAL3159, FC = 0.55; FRAAL1213, FC = 0.43; FRAAL3158, FC = 0.23), a methyl-crotonoyl-CoA carboxylase (FRAAL2343, FC = 0.15) and two propionyl-CoA carboxylases (FRAAL5672, FC = 0.05; FRAAL1210, FC = 0.68). The only overabundant one was a phosphoribosylaminoimidazole carboxylase (FRAAL6664, FC = 4.9) involved in the synthesis of purines.
A COG analysis (Figure 2) revealed that “C” (energy production) and “J” (translation) were the most represented in the nodule overabundant proteins, while “E” (amino acid transport and metabolism) and “K” (transcription) were the most represented in the underabundant proteins (Figure 2). Many COGs were very rare in the nodule (N, cell motility; U, intracellular trafficking; V, defense mechanisms and Q, secondary metabolites).
A drawing with the major overabundant proteins in nodules permits to show nitrogenase and other energy-generating, transporter and stress-coping determinants (Figure 3).

4. Discussion

Proteomics has the potential to decipher the physiological changes occurring in cells upon ecological transitions [32]. Establishment of the actinorhizal symbiosis is a drastic change for both partners that cannot be analyzed through negative genetics. Omics thus offer an interesting approach to pinpoint the key molecular players from both organisms. Proteomics in particular has been used to study Frankia either in early interaction with Alnus [18]—or with Elaeagnus angustifolia, Ceanothus thyrsiflorus and Coriaria myrtifolia [17], showing, among others, cellulose synthase and a potassium transporter over-detected upon contact with Alnus [18] and nitrogen fixation and assimilation proteins upon contact with Elaeagnus [17], as well as stress response and respiration proteins—or in the mature field nodules of a range of actinorhizal plants [21].
Nitrogenase is the most abundant protein complex in symbiotic Frankia alni. It comprises 17 genes that had previously been shown to be the most transcribed in mature nodules [7]. The present study confirms our vision that symbiotic Frankia is essentially a nitrogenase machine. Associated proteins were seen here, such as the SuF cluster assembly proteins and the hopanoid biosynthesis proteins (that protect nitrogenase from oxygen), but not the hydrogenase proteins even though they were among those highly expressed [33]. Frankia hydrogenase proteins seen previously in field nodules [21] were not detected.
The high overabundance for 2-oxoglutarate ferredoxin oxidoreductase (KorAB) confirms a role initially hypothesized based on its position next to nif genes of 2-oxoglutarate as a primary electron source for nitrogenase [21]. The position of the korAB genes next to nif genes and their duplication in symbiotic strains only are probably the result of a specialization on the one hand for symbiosis (FRAAL6798-6790, FC = 91.5) and on the other hand for saprophytic life (FRAAL1050-1051, FC = 0.80) as seen before for Hup [34].
Nitrogen fixation is an energy-demanding process that consumes eight ATP molecules per molecule of NH4+ produced. It is, thus, expected that the TCA cycle will be running at full force, transforming the photosynthates given by the plant into energy. The transcriptome of symbiotic F. alni was seen to contain several TCA genes upregulated at a high level [7] as was the case for the transcriptome of F. soli upon early contact [17].
There is an important specialization in nodules with only a twentieth of the genome identified (250 proteins with an FC ≥ 2) relative to a pure culture where five times as many proteins were identified at the same threshold. A similar specialization was noted in S. meliloti nodules [35].
The oxidative stress proteins catalase and superoxide dismutase are essential for the nodulation of Sinorhizobium [36] to cope with the oxidative burst induced as an early plant defense response against avirulent pathogens [37]. Nitroreductases play a similar role in Sinorhizobium [38]. In actinorhizal nodules, there is also an oxidative burst with which Frankia must cope [39]. Pure cultures of F. alni have been shown to have a basic expression of its two catalases that are then markedly upregulated upon contact with H2O2 or with methyl viologen [40]. However, it appears that this upregulation is not maintained in mature nodules.
GABA is considered a nutrient as well as an effector to trigger plant responses to heat, salt, herbivory and other stresses [41]. GABA has been detected in the metabolome of Alnus and Casuarina nodules and roots at very high levels [42,43]. GABA was also seen to improve nitrogenase and respiration in pure culture [43]. In leguminous nodules, GABA has been suggested to be a nutrient fed to the symbiont since it accumulated labeled 15N2 and contributed to the N-nutrition [44]. In Alnus, labeled 15NH4+ was recovered as alanine (Ala), γ-amino butyrate GABA), glutamine (Gln), glutamate (Glu), citrulline (Cit) and arginine (Arg) [45]. GABA was late to appear, but then, it continued at a fast pace. Since GABA synthesis involves the utilization of protons and releases CO2, it has been suggested as a means to reduce acidity [46].
The photosynthates fed by the plant to the bacterium have been the subject of several hypotheses. It has long been known that Alnus-infective strains could not use sugars [3], which is why organic acids are routinely used to grow most strains. The identification of a dicarboxylate transporter among those genes specific to nodules [47] indicated that dicarboxylates were good candidates. Several of these were measured in nodules and roots, and fumarate, succinate, malate and 2-oxoglutarate were found to be overabundant in nodules, although this approach is biased by the presence of mitochondria where citrate is also very abundant, yet it is toxic to Frankia [48]. The overabundance in the present study of the C4-dicarboxylate transporter, FRAAL1390 confirms that dicarboxylates play a role in trophic exchanges between symbionts, although its modest overabundance (FC = 3.42) points either to other transporters or to other photosynthates involved.
Chaperones such as GroEL have been shown to be essential to nodulation in Sinorhizobium meliloti [49]. They are also involved in response to other stresses such as heat [50] or salt [51] or simply for the assembly of very complex proteins such as nitrogenase [52]. Such chaperones were previously identified in the transcriptome of symbiotic Frankia alni [7].
Proteomics permits complementation of the vision achieved by transcriptomics. The two approaches have different strengths and weaknesses. Shotgun proteomics carried out in the standard mode (i.e., based on trypsin proteolysis) may miss small proteins, membrane proteins and polypeptides with rare combinations of arginine and lysine residues [53]. Transcriptomics misses unstable messengers or those with secondary structures. Furthermore, proteomics and transcriptomics address molecular entities with different half-lives and dynamics, thus they are highly complementary. The present study reinforces the picture of a microbe highly specialized for the energy-demanding nitrogenase and the accompanying stresses. It sheds no light, however, on the molecular processes involved in communication between the two partners.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/microorganisms10030651/s1: Table S1: List of Frankia proteins identified in nodules and in a nitrogen-replete pure culture and their spectral counts.

Author Contributions

P.P., N.A. and P.N. conceived the experiment; P.P., N.A. and P.F. grew the plant and the bacterium; G.M. and J.A. extracted the proteins and ran the proteomic analysis; D.A. and P.N. did the analyses and P.P. and P.N. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by ANR funding (BugsInACell ANR-13-BSV7-0013-03) to P.N.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The mass spectrometry proteomics data were deposited in the ProteomeXchange Consortium via the PRIDE partner repository (www.ebi.ac.uk/pride/archive/; 16 December 2021) with the dataset identifier PXD030468 and Project DOI 10.6019/PXD030468.

Acknowledgments

Thanks are expressed to Elise Lacroix for Greenhouse management (Université de Lyon) and to Aude Herrera-Belaroussi (Université de Lyon) for the nitrogen-replete Frankia cells.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Normand, P.; Fernandez, M.P. Frankia Brunchorst 1886, 174AL. Bergey’s Manual of Systematics of Archaea and Bacteria; Whitman, W., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2019. [Google Scholar] [CrossRef]
  2. Navarro, E.; Bousquet, J.; Moiroud, A.; Munive, A.; Piou, D.; Normand, P. Molecular phylogeny of Alnus (Betulaceae), inferred from nuclear ribosomal DNA ITS sequences. Plant Soil 2003, 254, 207–217. [Google Scholar] [CrossRef]
  3. Lechevalier, M. Taxonomy of the genus Frankia (Actinomycetales). Int. J. Syst. Bacteriol. 1994, 44, 1–8. [Google Scholar] [CrossRef]
  4. Baker, D.D. Relationships among pure-cultured strains of Frankia based on host specificity. Physiol. Plant 1987, 70, 245–248. [Google Scholar] [CrossRef]
  5. Normand, P.; Lapierre, P.; Tisa, L.S.; Gogarten, J.P.; Alloisio, N.; Bagnarol, E.; Bassi, C.A.; Berry, A.M.; Bickhart, D.M.; Choisne, N.; et al. Genome characteristics of facultatively symbiotic Frankia sp. strains reflect host range and host plant biogeography. Genome Res. 2007, 17, 7–15. [Google Scholar] [CrossRef] [PubMed]
  6. Nouioui, I.; Ghodhbane-Gtari, F.; del Carmen Montero-Calasanz, M.; Göker, M.; Meier-Kolthoff, J.P.; Schumann, P.; Rohde, M.; Goodfellow, M.; Fernandez, M.P.; Normand, P.; et al. Proposal of a type strain for Frankia alni (Woronin 1866) Von Tubeuf 1895, emended description of Frankia alni, and recognition of Frankia casuarinae sp. nov. and Frankia elaeagni sp. nov. Int. J. Syst. Evol. Microbiol. 2016, 66, 5201–5210. [Google Scholar] [CrossRef] [PubMed]
  7. Alloisio, N.; Queiroux, C.; Fournier, P.; Pujic, P.; Normand, P.; Vallenet, D.; Médigue, C.; Yamaura, M.; Kakoi, K.; Kucho, K.I. The Frankia alni symbiotic transcriptome. Mol. Plant Microbe Interact. 2010, 23, 593–607. [Google Scholar] [CrossRef] [PubMed]
  8. Granqvist, E.; Sun, J.; den Camp, R.O.; Pujic, P.; Hill, L.; Normand, P.; Morris, R.J.; Downie, J.A.; Geurts, R.; Oldroyd, G.E. Bacterial-induced calcium oscillations are common to nitrogen-fixing associations of nodulating legumes and nonlegumes. New Phytol. 2015, 207, 551–558. [Google Scholar] [CrossRef]
  9. Cissoko, M.; Hocher, V.; Gherbi, H.; Gully, D.; Carré-Mlouka, A.; Sane, S.; Pignoly, S.; Champion, A.; Ngom, M.; Pujic, P.; et al. Actinorhizal signaling molecules: Frankia Root Hair Deforming Factor shares properties with NIN inducing factor. Front Plant Sci. 2018, 9, 1494. [Google Scholar] [CrossRef]
  10. Van Ghelue, M.; Lovaas, E.; Ringo, E.; Solheim, B. Early interactions between Alnus glutinosa and Frankia strain ArI3. Production and specificity of root hair deformation factor(s). Physiol. Plant 1997, 99, 579–587. [Google Scholar] [CrossRef]
  11. Clavijo, F.; Diedhiou, I.; Vaissayre, V.; Brottier, L.; Acolatse, J.; Moukouanga, D.; Crabos, A.; Auguy, F.; Franche, C.; Gherbi, H.; et al. The Casuarina NIN gene is transcriptionally activated throughout Frankia root infection as well as in response to bacterial diffusible signals. New Phytol. 2015, 208, 887–903. [Google Scholar] [CrossRef]
  12. Gherbi, H.; Markmann, K.; Svistoonoff, S.; Estevan, J.; Autran, D.; Giczey, G.; Auguy, F.; Péret, B.; Laplaze, L.; Franche, C.; et al. SymRK defines a common genetic basis for plant root endosymbioses with arbuscular mycorrhiza fungi, rhizobia, and Frankia bacteria. Proc. Natl. Acad. Sci. USA 2008, 105, 4928–4932. [Google Scholar] [CrossRef] [PubMed]
  13. Griesmann, M.; Chang, Y.; Liu, X.; Song, Y.; Haberer, G.; Crook, M.B.; Billault-Penneteau, B.; Lauressergues, D.; Keller, J.; Imanishi, L.; et al. Phylogenomics reveals multiple independent losses of the nitrogen-fixing root nodule symbiosis. Science 2018, 361, 6398. [Google Scholar] [CrossRef] [PubMed]
  14. Persson, T.; Battenberg, K.; Demina, I.V.; Vigil-Stenman, T.; Vanden Heuvel, B.; Pujic, P.; Facciotti, M.T.; Wilbanks, E.G.; O’Brien, A.; Fournier, P.; et al. Candidatus Frankia datiscae Dg1, the actinobacterial microsymbiont of Datisca glomerata, expresses the canonical nod genes nodABC in symbiosis with its host plant. PLoS ONE 2015, 10, e0127630. [Google Scholar] [CrossRef]
  15. Nguyen, T.V.; Wibberg, D.; Vigil-Stenman, T.; Berckx, F.; Battenberg, K.; Demchenko, K.N.; Blom, J.; Fernandez, M.P.; Yamanaka, T.; Berry, A.M.; et al. Frankia-enriched metagenomes from the earliest diverging symbiotic Frankia cluster: They come in teams. Genome Biol. Evol. 2019, 11, 2273–2291. [Google Scholar] [CrossRef] [PubMed]
  16. Gtari, M.; Ghodhbane-Gtari, F.; Nouioui, I. Frankia soli sp. nov., an actinobacterium isolated from soil beneath Ceanothus jepsonii. Int. J. Syst. Evol. Microbiol. 2019, 70, 1203–1209. [Google Scholar] [CrossRef]
  17. Gueddou, A.; Sarker, I.; Sen, A.; Ghodhbane-Gtari, F.; Benson, D.R.; Armengaud, J.; Gtari, M. Effect of actinorhizal root exudates on the proteomes of Frankia soli NRRL B-16219, a strain colonizing the root tissues of its actinorhizal host via intercellular pathway. Res. Microbiol. 2021, 103900. [Google Scholar] [CrossRef]
  18. Pujic, P.; Alloisio, N.; Fournier, P.; Roche, D.; Sghaier, H.; Miotello, G.; Armengaud, J.; Berry, A.M.; Normand, P. Omics of the early molecular dialogue between Frankia alni and Alnus glutinosa and the cellulase synton. Environ. Microbiol. 2019, 21, 3328–3345. [Google Scholar] [CrossRef]
  19. Hammad, Y.; Nalin, R.; Marechal, J.; Fiasson, K.; Pepin, R.; Berry, A.M.; Normand, P.; Domenach, A.M. A possible role for phenylacetic acid (PAA) in Alnus glutinosa nodulation by Frankia. Plant Soil 2003, 254, 193–205. [Google Scholar] [CrossRef]
  20. Carro, L.; Pujic, P.; Alloisio, N.; Fournier, P.; Boubakri, H.; Hay, A.E.; Poly, F.; Francois, P.; Hocher, V.; Mergaert, P.; et al. Alnus peptides modify membrane porosity and induce the release of nitrogen-rich metabolites from nitrogen-fixing Frankia. ISME J. 2015, 9, 1723–1733. [Google Scholar] [CrossRef]
  21. Mastronunzio, J.E.; Benson, D.R. Wild nodules can be broken: Proteomics of Frankia in field-collected root nodules. Symbiosis 2010, 50, 13–26. [Google Scholar] [CrossRef]
  22. Alloisio, N.; Félix, S.; Maréchal, J.; Pujic, P.; Rouy, Z.; Vallenet, D.; Medigue, C.; Normand, P. Frankia alni proteome under nitrogen-fixing and nitrogen-replete conditions. Physiol. Plant 2007, 13, 440–453. [Google Scholar] [CrossRef]
  23. Ghedira, K.; Harigua-Souiai, E.; Ben Hamda, C.; Fournier, P.; Pujic, P.; Guesmi, S.; Guizani, I.; Miotello, G.; Armengaud, J.; Normand, P.; et al. The PEG-responding desiccome of the alder microsymbiont Frankia alni. Sci. Rep. 2018, 8, 759. [Google Scholar] [CrossRef] [PubMed]
  24. Normand, P.; Lalonde, M. Evaluation of Frankia strains isolated from provenances of two Alnus species. Can. J. Microbiol. 1982, 28, 1133–1142. [Google Scholar] [CrossRef]
  25. Murry, M.A.; Fontaine, M.S.; Torrey, J.G. Growth kinetics and nitrogenase induction in Frankia sp. HFPArI3 grown in batch culture. Plant Soil 1984, 78, 61–78. [Google Scholar] [CrossRef]
  26. Fåhraeus, G. The infection of clover root hairs by nodule bacteria studied by a simple glass slide technique. J. Gen. Microbiol. 1957, 16, 374–381. [Google Scholar] [CrossRef]
  27. Schwenke, J. Rapid, exponential growth and increased biomass yield of some Frankia strains in buffered and stirred mineral medium (BAP) with phosphatidyl choline. Plant Soil 1991, 137, 37–41. [Google Scholar] [CrossRef]
  28. Hartmann, E.M.; Allain, F.; Gaillard, J.C.; Pible, O.; Armengaud, J. Taking the shortcut for high-throughput shotgun proteomic analysis of bacteria. Methods Mol. Biol. 2014, 1197, 275–285. [Google Scholar]
  29. Klein, G.; Mathé, C.; Biola-Clier, M.; Devineau, S.; Drouineau, E.; Hatem, E.; Marichal, L.; Alonso, B.; Gaillard, J.C.; Lagniel, G.; et al. RNA-binding proteins are a major target of silica nanoparticles in cell extracts. Nanotoxicology 2016, 10, 1555–1564. [Google Scholar] [CrossRef]
  30. Gouveia, D.; Grenga, L.; Pible, O.; Armengaud, J. Quick microbial molecular phenotyping by differential shotgun proteomics. Environ. Microbiol. 2020, 22, 2996–3004. [Google Scholar] [CrossRef]
  31. Leul, M.; Normand, P.; Sellstedt, A. The organization, regulation and phylogeny of uptake hydrogenase genes in Frankia. Physiol. Plant 2007, 130, 464–470. [Google Scholar] [CrossRef]
  32. Richau, K.H.; Kudahettige, R.L.; Pujic, P.; Kudahettige, N.P.; Sellstedt, A. Structural and gene expression analyses of uptake hydrogenases and other proteins involved in nitrogenase protection in Frankia. J. Biosci. 2013, 38, 703–712. [Google Scholar] [CrossRef] [PubMed]
  33. Yurgel, S.N.; Qu, Y.; Rice, J.T.; Ajeethan, N.; Zink, E.M.; Brown, J.M.; Purvine, S.; Lipton, M.S.; Kahn, M.L. Specialization in a nitrogen-fixing symbiosis: Proteome differences between Sinorhizobium medicae bacteria and bacteroids. Mol. Plant-Microbe Interact. 2021, 34, 1409–1422. [Google Scholar] [CrossRef] [PubMed]
  34. Santos, R.; Herouart, D.; Puppo, A.; Touati, D. Critical protective role of bacterial superoxide dismutase in rhizobium-legume symbiosis. Mol. Microbiol. 2000, 38, 750–759. [Google Scholar] [CrossRef]
  35. Santos, R.; Herouart, D.; Sigaud, S.; Touati, D.; Puppo, A. Oxidative burst in alfalfa-Sinorhizobium meliloti symbiotic interaction. Mol. Plant-Microbe Interact. 2001, 14, 86–89. [Google Scholar] [CrossRef]
  36. Berger, A.; Boscari, A.; Puppo, A.; Brouquisse, R. Nitrate reductases and hemoglobins control nitrogen-fixing symbiosis by regulating nitric oxide accumulation. J. Exp. Bot. 2021, 72, 873–884. [Google Scholar] [CrossRef] [PubMed]
  37. Tavares, F.; Santos, C.L.; Sellstedt, A. Reactive oxygen species in legume and actinorhizal nitrogen-fixing symbioses: The microsymbiont’s responses to an unfriendly reception. Physiol. Plant 2007, 130, 344–356. [Google Scholar] [CrossRef]
  38. Santos, C.L.; Vieira, J.; Sellstedt, A.; Normand, P.; Moradas-Ferreira, P.; Tavares, F. Modulation of Frankia alni ACN14a oxidative stress response: Activity, expression and phylogeny of catalases. Physiol. Plant 2007, 130, 454–463. [Google Scholar] [CrossRef]
  39. Bown, A.W.; Shelp, B.J. Plant GABA: Not just a metabolite. Trends Plant Sci. 2016, 21, 811–813. [Google Scholar] [CrossRef]
  40. Brooks, J.M.; Benson, D.R. Comparative metabolomics of root nodules infected with Frankia sp. strains and uninfected roots from Alnus glutinosa and Casuarina cunninghamiana reflects physiological integration. Symbiosis 2016, 70, 87–96. [Google Scholar] [CrossRef]
  41. Hay, A.E.; Herrera-Belaroussi, A.; Rey, M.; Fournier, P.; Normand, P.; Boubakri, H. Feed-back regulation of N-fixation in Frankia-Alnus symbiosis through amino acids contents in field and greenhouse nodules. Mol. Plant-Microbe Interact. 2019, 33, 499–508. [Google Scholar] [CrossRef]
  42. Prell, J.; Bourdes, A.; Karunakaran, R.; Lopez-Gomez, M.; Poole, P. Pathway of gamma-aminobutyrate metabolism in Rhizobium leguminosarum 3841 and its role in symbiosis. J. Bacteriol. 2009, 191, 2177–2186. [Google Scholar] [CrossRef] [PubMed]
  43. Lundberg, P.; Lundquist, P.O. Primary metabolism in N2-fixing Alnus incana-Frankia symbiotic root nodules studied with 15N and 31P nuclear magnetic resonance spectroscopy. Planta 2004, 219, 661–672. [Google Scholar] [CrossRef]
  44. Bown, A.W.; Shelp, B.J. The metabolism and functions of gamma-aminobutyric acid. Plant Physiol. 1997, 115, 1–5. [Google Scholar] [CrossRef] [PubMed]
  45. Jeong, J.; Suh, S.; Guan, C.; Tsay, Y.F.; Moran, N.; Oh, C.J.; An, C.S.; Demchenko, K.N.; Pawlowski, K.; Lee, Y. A nodule-specific dicarboxylate transporter from alder is a member of the peptide transporter family. Plant Physiol. 2004, 134, 969–978. [Google Scholar] [CrossRef] [PubMed]
  46. Carro, L.; Persson, T.; Pujic, P.; Alloisio, N.; Fournier, P.; Boubakri, H.; Pawlowski, K.; Normand, P. Organic acids metabolism in Frankia alni. Symbiosis 2016, 70, 37–48. [Google Scholar] [CrossRef]
  47. Farkas, A.; Maróti, G.; Dürgő, H.; Györgypál, Z.; Lima, R.M.; Medzihradszky, K.F.; Kereszt, A.; Mergaert, P.; Kondorosi, É. Medicago truncatula symbiotic peptide NCR247 contributes to bacteroid differentiation through multiple mechanisms. Proc. Natl. Acad. Sci. USA 2014, 111, 5183–5188. [Google Scholar] [CrossRef]
  48. Alexandre, A.; Laranjo, M.; Oliveira, S. Global transcriptional response to heat shock of the legume symbiont Mesorhizobium loti MAFF303099 comprises extensive gene downregulation. DNA Res. 2014, 21, 195–206. [Google Scholar] [CrossRef]
  49. Brígido, C.; Alexandre, A.; Oliveira, S. Transcriptional analysis of major chaperone genes in salt-tolerant and salt-sensitive mesorhizobia. Microbiol. Res. 2012, 167, 623–629. [Google Scholar] [CrossRef]
  50. Ribbe, M.W.; Burgess, B.K. The chaperone GroEL is required for the final assembly of the molybdenum-iron protein of nitrogenase. Proc. Natl. Acad. Sci. USA 2001, 98, 5521–5525. [Google Scholar] [CrossRef]
  51. Durighello, E.; Christie-Oleza, J.A.; Armengaud, J. Assessing the exoproteome of marine bacteria, lesson from a RTX-Toxin abundantly secreted by Phaeobacter strain DSM 17395. PLoS ONE 2014, 9, e89691. [Google Scholar] [CrossRef]
  52. Tatusov, R.L.; Natale, D.A.; Garkavtsev, I.V.; Tatusova, T.A.; Shankavaram, U.T.; Rao, B.S.; Kiryutin, B.; Galperin, M.Y.; Fedorova, N.D.; Koonin, E.V. The COG database: New developments in phylogenetic classification of proteins from complete genomes. Nucleic Acids Res. 2001, 29, 22–28. [Google Scholar] [CrossRef] [PubMed]
  53. Vallenet, D.; Calteau, A.; Cruveiller, S.; Gachet, M.; Lajus, A.; Josso, A.; Mercier, J.; Renaux, A.; Rollin, J.; Rouy, Z.; et al. MicroScope in 2017: An expanding and evolving integrated resource for community expertise of microbial genomes. Nucleic Acids Res. 2017, 45, D517–D528. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Circular map of the genome of Frankia alni with the proteins over-abundant in nodules relative to a nitrogen-replete pure-culture (FC ≥ 2) positioned along the genome. The nif genes are at the top close to the origin of replication.
Figure 1. Circular map of the genome of Frankia alni with the proteins over-abundant in nodules relative to a nitrogen-replete pure-culture (FC ≥ 2) positioned along the genome. The nif genes are at the top close to the origin of replication.
Microorganisms 10 00651 g001
Figure 2. Histogram of over- and underabundant proteins coding genes (FC > 2 hatched; FC < 0.5 dotted) assigned to COGs. Undefined categories and those with one or less occurrences were omitted.
Figure 2. Histogram of over- and underabundant proteins coding genes (FC > 2 hatched; FC < 0.5 dotted) assigned to COGs. Undefined categories and those with one or less occurrences were omitted.
Microorganisms 10 00651 g002
Figure 3. Schematics of a Frankia cell in its symbiotic host cell showing the likely function and localization of the 100 most abundant proteins related to nitrogen fixation, lipid metabolism, Krebs cycle (TCA), cell signaling, electron transfer, transport of anions and cations, enzymes involved in nucleotides biosynthesis, DNA metabolism and regulation and storage and stress proteins. The identities of the proteins and specific genes are listed in Table 1.
Figure 3. Schematics of a Frankia cell in its symbiotic host cell showing the likely function and localization of the 100 most abundant proteins related to nitrogen fixation, lipid metabolism, Krebs cycle (TCA), cell signaling, electron transfer, transport of anions and cations, enzymes involved in nucleotides biosynthesis, DNA metabolism and regulation and storage and stress proteins. The identities of the proteins and specific genes are listed in Table 1.
Microorganisms 10 00651 g003
Table 1. List of the 100 most over-abundant Frankia proteins in nodules with the NCBI accession, the FRAAL Id, gene names, the protein name, the fold change, the COG (cluster of orthologous genes), the phylogenetic clusters distribution and the NSAF (normalized spectral abundance factors). The COGs are according to [30]. The distribution code A is present in all Frankia and in other actinomycetes, F is present in all Frankia strains, f is present in some Frankia strains and S is present in symbiotic Frankia strains (C1, C1c, C2 and C3 but not in C4). C1, C1c, C2, C3 and C4 indicate it is present in this cluster based on a threshold of 35% Id as seen on Mage [31]. Hypothetical genes have been removed from the list; they are listed in Table S1.
Table 1. List of the 100 most over-abundant Frankia proteins in nodules with the NCBI accession, the FRAAL Id, gene names, the protein name, the fold change, the COG (cluster of orthologous genes), the phylogenetic clusters distribution and the NSAF (normalized spectral abundance factors). The COGs are according to [30]. The distribution code A is present in all Frankia and in other actinomycetes, F is present in all Frankia strains, f is present in some Frankia strains and S is present in symbiotic Frankia strains (C1, C1c, C2 and C3 but not in C4). C1, C1c, C2, C3 and C4 indicate it is present in this cluster based on a threshold of 35% Id as seen on Mage [31]. Hypothetical genes have been removed from the list; they are listed in Table S1.
NCBI AccessionFRAAL #GeneFunctional DescriptionFold ChangeCOGDistribution% NSAF Nodule
WP_009741523.1FRAAL6813nifHNitrogenase iron protein NifH291.33PS0.009
WP_011607845.1FRAAL6811nifKNitrogenase molybdenum–iron protein subunit beta NifK279.00CS0.005
WP_011607832.1FRAAL6798korB2-Oxoglutarate ferredoxin oxidoreductase subunit beta KorB91.50CS0.002
WP_011606600.1FRAAL5516 ATPase89.00OC10.003
WP_011607846.1FRAAL6812nifDNitrogenase molybdenum–iron protein subunit alpha NifD86.17CS0.002
WP_041939983.1FRAAL6802nifUIron–sulfur cluster assembly accessory protein NifU84.17OS0.004
WP_041939982.1FRAAL6799korA2-Oxoglutarate ferredoxin oxidoreductase subunit alpha KorA82.00CS0.001
WP_011602860.1FRAAL1673mopBMolybdenum-binding protein MopB64.17RS0.004
WP_011605156.1FRAAL4025ddnBF420H(2)-dependent quinone nitroreductase, DdnB44.33JS0.003
WP_011607839.1FRAAL6805nifWNitrogenase-stabilizing/protective protein NifW40.33PS0.003
WP_041939985.1FRAAL6808nifXNitrogen fixation protein NifX33.83-S0.002
WP_011602101.1FRAAL0868tetRTetR family transcriptional regulator33.09KC10.002
WP_011607841.1FRAAL6807nifAssociated nitrogen fixation protein NifX23.83-S0.003
WP_011604521.1FRAAL3374merCoenzyme F420-dependent N(5),N(10)-methylenetetrahydromethanopterin reductase22.83Cf0.001
WP_011604385.1FRAAL3236ddnDeazaflavin-dependent nitroreductase18.83-C10.001
WP_011601860.1FRAAL0611ybaBNucleoid-associated protein17.67OA0.001
WP_011603220.1FRAAL2049 Glycosyl transferase15.50MC1c, C20.000
WP_011601350.1FRAAL0082arfAAmmonia release factor ArfA13.33Mf0.001
WP_011606048.1FRAAL4934fadA3-Ketoacyl-CoA thiolase12.91IA0.003
WP_011602036.1FRAAL0800thiCPhosphomethylpyrimidine synthase12.80HA0.000
WP_041939400.1FRAAL3855fdxI4Fe-4S iron–sulfur binding ferredoxin12.67CA0.001
WP_011601559.1FRAAL0301 Transcriptional regulator12.38Tc10.000
WP_011603772.1FRAAL2618 Acetyl-CoA acetyltransferase11.83IA0.000
WP_011607074.1FRAAL6022gadGamma-aminobutyraldehyde dehydrogenase11.56CA0.000
WP_011604927.1FRAAL3789 Thioesterase10.88RC1, C3, C40.001
WP_011603462.1FRAAL2298folDMethylenetetrahydrofolate dehydrogenase FolD10.86HA0.001
WP_011605870.1FRAAL4755 Membrane phosphatase10.80KA0.000
WP_011607913.1FRAAL6878rsmGRibosomal RNA small subunit methyltransferase G10.77MA0.001
WP_011607828.1FRAAL6795qorBQuinone oxydoreductase QorB10.67MC1, C3, C40.001
WP_011606846.1FRAAL5774dxr1-Deoxy-D-xylulose 5-phosphate reductoisomerase Dxr10.54IA0.001
WP_050997128.1FRAAL3342ssuASulfonate transport system substrate-binding SsuA10.50PC1, C2, C40.000
WP_041938619.1FRAAL0128purFAmidophosphoribosyltransferase PurF9.86FA0.000
WP_009742300.1FRAAL5827 DNA-binding protein HU-beta, NS1 (HU-1)9.76LA0.016
WP_041939818.1FRAAL6056 ATP-binding protein9.60-A0.004
WP_041939008.1FRAAL1862rpiBRibose 5-phosphate isomerase RplB9.07GA0.003
WP_011601953.1FRAAL0711 Transcriptional regulator8.76KC1, C1c, C3, C40.002
WP_009737569.1FRAAL4338bfrABacterioferritin BfrA8.49PA0.004
WP_011602931.1FRAAL1747 Putative carboxyvinyl–carboxyphosphonate phosphorylmutase8.43GC1, C1c, C40.000
WP_041938916.1FRAAL1429hpnDPhytoene synthase HpnD8.40IA0.003
WP_011604375.1FRAAL3226bkdCBranched-chain alpha-keto acid dehydrogenase subunit BkdC8.14CA0.000
WP_041940119.1FRAAL0708 Lipid esterase 7.91RC1, C2, C3, C40.000
WP_011602626.1FRAAL1431fppSFarnesyl diphosphate synthase7.75HA0.002
WP_041939587.1FRAAL4898 LuxR family transcriptional regulator7.67KC1, C1c, C3, C40.000
WP_011606077.1FRAAL4963adkAdenylate kinase7.50FA0.001
WP_011604508.1FRAAL3361 Short-chain fatty acid–CoA ligase7.43IA0.000
WP_011604193.1FRAAL3044 Transcriptional regulator7.36Kf0.000
WP_041939680.1FRAAL5348cspACold shock protein7.33KA0.002
WP_011602011.1FRAAL0771dgcDiguanylate cyclase7.17Tf0.000
WP_011607242.1FRAAL6192 Adenylate kinase7.17FA0.001
WP_041939231.1FRAAL2835adhAcyl-CoA dehydrogenase6.71IA0.000
WP_011606204.1FRAAL5096pksPolyketide cyclase/dehydrase and lipid transport6.56-A0.001
WP_050997247.1FRAAL5736ftsKCell division protein FtsK6.43DA0.000
WP_011604554.1FRAAL3408 Alpha/beta hydrolase6.37Rf0.001
WP_041938863.1FRAAL1156sucCSuccinyl-CoA synthetase subunit beta SucC6.14CA0.012
WP_011605559.1FRAAL4438 Sugar-phosphate dehydrogenase6.08CA0.002
WP_041939760.1FRAAL5796smcChromosome segregation protein Smc6.04DA0.000
WP_041939603.1FRAAL4991thiFThiamine biosynthesis protein ThiF6.00Hf0.000
WP_041938832.1FRAAL0997hemLGlutamate-1-semialdehyde aminotransferase HemL5.93HA0.001
WP_011603632.1FRAAL2475glbNTruncated hemoglobin GlbN5.87Rf0.001
WP_011601393.1FRAAL0129purMPhosphoribosylaminoimidazole synthetase PurM5.77FA0.001
WP_041938655.1FRAAL0282 Putative integrase/resolvase5.71Lf0.000
WP_011603454.1FRAAL2289 Putative two-component oxygen sensor kinase5.67TA0.000
WP_011604752.1FRAAL3613 Diacylglycerol O-acyltransferase5.67-f0.000
WP_011602534.1FRAAL1337pknSerine/threonine protein kinase5.50TF0.000
WP_011603231.1FRAAL2064acnAAconitate hydratase AcnA5.44CA0.006
WP_011605678.1FRAAL4561sufE(2Fe-2S)-Binding protein SufE5.40PA0.001
WP_050997107.1FRAAL2825 Putative N-fatty-acyl-amino acid synthase5.38EA0.001
WP_011606156.1FRAAL5045ahpEPeroxiredoxin AhpE5.31OA0.009
WP_041939398.1FRAAL3827braBranched-chain amino acid ABC transporter ATP-binding protein5.18EA0.001
WP_011607046.1FRAAL5991rshAAnti-sigma factor RshA5.12TA0.004
WP_011602689.1FRAAL1499cysOSulfur carrier protein CysO5.07HA0.002
WP_011602425.1FRAAL1224 Acetyl-CoA synthetase5.04IA0.001
WP_011602282.1FRAAL1055pyrEOrotate phosphoribosyltransferase pyrE5.00FA0.001
WP_041939967.1FRAAL6738purSPhosphoribosylformylglycinamidine synthase PurS4.92Ff0.001
WP_041939945.1FRAAL6664purEPhosphoribosylaminoimidazole carboxylase PurE4.90FA0.003
WP_035923438.1FRAAL6556crpCrp/Fnr family transcriptional regulator4.83TA0.005
WP_011606876.1FRAAL5804rncRibonuclease III rnc4.82KA0.000
WP_011604517.1FRAAL3370fklBPeptidyl–prolyl cis–trans isomerase FklB4.80OA0.003
WP_041939534.1FRAAL4560sufCFe–S cluster assembly ATPase SufC4.62OA0.004
WP_011603306.1FRAAL2141ruvBHolliday junction DNA helicase RuvB4.60LA0.000
WP_011605202.1FRAAL4072pksPolyketide synthase4.56Qf0.000
WP_011601665.1FRAAL0409 Acetyl-CoA acetyltransferase4.53IA0.000
WP_011606938.1FRAAL5877etfAElectron transfer flavoprotein EtfA4.48CA0.007
WP_011601417.1FRAAL0156gdhBGlutamate dehydrogenase GdhB4.47Ef0.000
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pujic, P.; Alloisio, N.; Miotello, G.; Armengaud, J.; Abrouk, D.; Fournier, P.; Normand, P. The Proteogenome of Symbiotic Frankia alni in Alnus glutinosa Nodules. Microorganisms 2022, 10, 651. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms10030651

AMA Style

Pujic P, Alloisio N, Miotello G, Armengaud J, Abrouk D, Fournier P, Normand P. The Proteogenome of Symbiotic Frankia alni in Alnus glutinosa Nodules. Microorganisms. 2022; 10(3):651. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms10030651

Chicago/Turabian Style

Pujic, Petar, Nicole Alloisio, Guylaine Miotello, Jean Armengaud, Danis Abrouk, Pascale Fournier, and Philippe Normand. 2022. "The Proteogenome of Symbiotic Frankia alni in Alnus glutinosa Nodules" Microorganisms 10, no. 3: 651. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms10030651

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop