Skip to main content

A tale of two lineages: how the strains of the earliest divergent symbiotic Frankia clade spread over the world

Abstract

It is currently assumed that around 100 million years ago, the common ancestor to the Fabales, Fagales, Rosales and Cucurbitales in Gondwana, developed a root nodule symbiosis with a nitrogen-fixing bacterium. The symbiotic trait evolved first in Frankia cluster-2; thus, strains belonging to this cluster are the best extant representatives of this original symbiont. Most cluster-2 strains could not be cultured to date, except for Frankia coriariae, and therefore many aspects of the symbiosis are still elusive. Based on phylogenetics of cluster-2 metagenome-assembled genomes (MAGs), it has been shown that the genomes of strains originating in Eurasia are highly conserved. These MAGs are more closely related to Frankia cluster-2 in North America than to the single genome available thus far from the southern hemisphere, i.e., from Papua New Guinea.

To unravel more biodiversity within Frankia cluster-2 and predict routes of dispersal from Gondwana, we sequenced and analysed the MAGs of Frankia cluster-2 from Coriaria japonica and Coriaria intermedia growing in Japan, Taiwan and the Philippines. Phylogenetic analyses indicate there is a clear split within Frankia cluster-2, separating a continental from an island lineage. Presumably, these lineages already diverged in Gondwana.

Based on fossil data on the host plants, we propose that these two lineages dispersed via at least two routes. While the continental lineage reached Eurasia together with their host plants via the Indian subcontinent, the island lineage spread towards Japan with an unknown host plant.

Peer Review reports

Background

The nitrogen-fixing clade encompasses all symbiotic plants able to form root nodules which host diazotrophic bacteria. These plants can be traced back to the common ancestor of the Fabales, Fagales, Cucurbitales, and Rosales. While legumes (Fabaceae, Fabales) and Paraspona (Cannabaceae, Rosales) engage with Gram-negative rhizobia, the remaining root nodule-forming plants all engage with Gram-positive Frankia. The analysis of the evolution of this symbiosis involved several different scenarios [1,2,3], but the most recent phylogenomic studies have shown that in the most parsimonious hypothesis the common ancestor was symbiotic and the symbiotic capability was subsequently lost in the majority of lineages [4, 5].

A detailed study on the distribution of the genus Nothofagus (Fagales) concluded that the order evolved in the supercontinent Gondwana [6]. The same origin was found for the genus Coriaria (Cucurbitales) [7]. Therefore, the common symbiotic ancestor of the nitrogen-fixing plant clade must have evolved in Gondwana. This ancestor should have had either a rhizobial or a Frankia microsymbiont. Van Velzen et al. [8] argued persuasively that, given the polyphyletic origin of the oxygen protection system for bacterial nitrogenase in nodules, which is required for rhizobial but not Frankia nitrogen fixation, the original microsymbiont cannot have been a rhizobial strain, i.e. it must have been a Frankia strain.

Actinobacteria from the genus Frankia are the microsymbionts of actinorhizal plants. Phylogenetically, Frankia strains can be grouped into four phylogenetic clades called clusters, three of which encompass symbiotic strains and roughly represent host specificity groups [9]. Phylogenetic analysis shows that after the split of symbiotic cluster-2 and non-symbiotic cluster-4, the precursor of the symbiotic clusters 1 and 3 split off from cluster-4 [10,11,12,13,14]. Frankia cluster-2, where the symbiotic trait evolved for the first time, is therefore of particular interest for understanding the evolution of actinorhizal symbiosis as its strains seem to represent the closest approximation of the original symbiont. The evolution of cluster-2 should yield insight into the evolution of root nodule symbiosis in general.

Frankia strains grow as a mycelium ex planta. Cluster-2 strains show low saprotrophic potential [12, 15] and were, in spite numerous unsuccessful isolation efforts, considered uncultivable until Gtari et al. [11] and Gueddou et al. [16] published the isolation of two closely related alkaliphilic strains of Frankia coriariae, from nodules of the Mediterranean species Coriaria myrtifolia that fulfilled Koch’s postulate. Generally, cluster-2 inocula represent assemblages of different strains, and not all of these strains can enter an effective symbiosis with the host plant the inoculum came from [14]. This has been illustrated by the fact that metagenome-assembled genomes (MAGs) [17] differ when sequencing whole nodules, as compared to when isolating Frankia vesicles – i.e. nitrogen-fixing symbiotic structures – before sequencing [14]. Some strains can enter the nodule but cannot fix nitrogen there.

Frankia cluster-2 strains have a wide host range and the geographic distribution of their host plants is disjunct [18]. Actinorhizal members of the order Rosales nodulated by cluster-2 strains, namely Dryadoideae (Rosaceae) and Ceanothus sp. (Rhamnaceae), are restricted to North America. The family Datiscaceae (Cucurbitales) consists of two species: Datisca cannabina which is mostly restricted to northern India/Pakistan/Nepal with some occurrences in Turkey [19], and Datisca glomerata restricted to California and Northern Mexico [20]. Members of the genus Coriaria (Coriariaceae, Cucurbitales) have the broadest geographic range. They can be found in New Zealand, Papua New Guinea, the Philippines, Taiwan, Japan, China, Nepal, Pakistan, northern India, the Mediterranean, and the west coast of South America [7, 21]. New Zealand is the center of diversity for Coriaria, with eight endemic species, while in any of the other regions only one or two different species can be found.

Interestingly, the biodiversity of Frankia cluster-2 strains in Eurasia is very low. There is 96–99.8% mean average nucleotide identity (ANI) between MAGs sequenced from nodules of inocula originating in France, Pakistan and Japan [14]. However, a MAG from an inoculum originating in Papua New Guinea (Candidatus Frankia meridionalis Cppng1) shows only ca. 85% mean ANI with the Eurasian Frankia cluster-2 MAGs [14]. Given that the symbiosis had evolved in Gondwana, it would seem plausible that Frankia cluster-2 split already into different lineages in Gondwana or the southern hemisphere. The question arises how did the Frankia cluster-2 and their host plants dispersed from Gondwana to Eurasia? We hypothesize that several lineages could have dispersed in the southern hemisphere, while only a few –or one– of them might have reached the northern hemisphere.

Therefore, in the hope to find representatives of a second Frankia cluster-2 lineage in the northern hemisphere, we obtained nodules from Coriaria intermedia from Taiwan and from the Philippines, as well as from Coriaria japonica in Japan, to analyse the corresponding Frankia cluster-2 MAGs.

Results and discussion

Sequencing of Frankia MAGs from Coriaria intermedia growing in Taiwan and the Philippines, and from Coriaria japonica growing in Japan

The Frankia metagenome-assembled genomes (MAGs, 17) obtained in this study were named according to the nomenclature established by Nguyen et al. [14]: [name of inoculum]_[initials of host plant from which DNA was isolated]_[“nod” for direct isolation of mixed plant and bacterial DNA from nodules, “vc” for isolation of DNA from vesicle clusters isolated from nodules].

Coriaria intermedia nodules collected at Taiping Mountain (Taiwan) were used for DNA isolation, sequencing and the assembly of a Frankia cluster-2 MAG termed CiT1_Ci_nod (Tables 1 and 2). Six sets of C. intermedia nodules from Poblacion, Atok (Philippines) were available; the first set, CiP1, was used as inoculum. The five other nodule samples from Poblacion were used for DNA isolation and sequencing, yielding MAGs CiP2_Ci_nod to CiP6_Ci_nod (Tables 1 and 2). CiP2_Ci_nod consisted of two different strains too similar to each other to be separated by bioinformatics means; therefore, the MAG was not of high quality. CiP3_Ci_nod and CiP4_Ci_nod represented one strain each and showed good BUSCO values (91.9 and 87.2% BUSCO, respectively; Table 2; Supplementary Table S2). CiP5_Ci_nod and CiP6_Ci_nod were not of high quality since the metagenomes contained a large contribution of non-Frankia bacterial DNA. The inoculum CiP1 was used for cross-inoculation studies. DNA was isolated from nodules induced on the Mediterranean species Coriaria myrtifolia (CiP1_Cm_nod1, CiP1_Cm_nod2). These two Frankia MAGs were quite dissimilar from all other genomes going back to nodules from Poblacion (Table 2; Fig. 1).

Table 1 Inocula used and MAGs sequenced in this study. The Philippine inoculum CiP1 was used for cross-inoculation studies and was not sequenced on its original host
Table 2 Features of MAGs sequenced in this study. The complete results of the BUSCO analysis are presented in Supplementary Table S2
Fig. 1
figure 1

Phylogenetic tree based on whole-genome sequences inferred using the neighbor-joining algorithm as implemented in the PHYLIP package [22]. The tree was built for 34 Frankia genomes out of a core of 214 genes per genome, 7276 in total, by EDGAR 2.0 [23, 24]. The core has 92,816 amino acid residues per genome, 3,155,744 in total. Genomes used were Candidatus Frankia datisca Dg1 (NC_015656.1), Frankia coriariae BMG5.1 (NZ_JWIO00000000.1), Candidatus Frankia californiensis Dg2 (FLUV00000000.1), Candidatus Frankia meridionalis Cppng1 (CADDZT010000001-CADDZT010000101), Frankia asymbiotica M16386 (NZ_MOMC00000000.1), Frankia saprophytica CN3 (NZ_AGJN00000000.2), Frankia inefficax EuI1c (NC_014666.1), Frankia irregularis DSM45899 (NZ_FAOZ00000000.1), Frankia elaeagni BMG5.12 (NZ_ARFH00000000.1), Frankia discariae BCU110501 (NZ_ARDT00000000.1), Frankia soli NRRL B-16219 (MAXA00000000.1), Frankia casuarinae CcI3 (NC_007777.1), Frankia canadensis ARgP5 (GCF_900241035.1), Candidatus Frankia nodosporulans AgTrS (NZ_CADCWS000000000.1), Candidatus Frankia alpina AiOr (GCA_902806485), Frankia alni ACN14a (NC_008278.1), Frankia torreyi CpI1 (NZ_JYFN00000000.1) and Frankia sp. QA3 (AJWA00000000). The genomes of Cryptosporangium arvum DSM44712 (JFBT01000000) and Jatrophihabitans endophyticus DSM45617 (NZ_FQVU00000000.1) were added for rooting. Bootstrap values are 100 for every branch; they were calculated in R [25] using the packages APE [26] and phangorn [27]. The genomes sequenced from nodules of Coriaria myrtifolia induced by the inoculum from the Philippines, CiP1, are labeled by a green outline. The size bar denotes 0.01 changes

DNA from four sets of nodules from C. japonica from Japan was isolated (Cj2_Cj_nod to Cj5_Cj_nod). The MAGs of these nodules were near-complete (84.5–91.2% BUSCO, Table 2). Unlike the Cj1_Dg_nod Frankia MAG which also originated from an inoculum from Japan [14], the MAGs of these strains did not contain any canonical nod genes which previously had been identified [12,13,14]. While neither the MAG CiT1_Ci_nod, CiP4_Ci_nod, CiP1_Cm_nod1 or CiP1_Cm_nod2 contained any representatives of the canonical nod genes, CiP3_Ci_nod and CiP5_Ci_nod contained a truncated version of the nod2 region identified in the genome of Candidatus Frankia meridionalis Cppng1, nodC-nltIJ-nodU (Supplementary Fig. S1 [14]). The nod gene status of CiP2_Ci_nod was unclear; it contained a nodU copy but not the rest of the operon, which might be related to the fact that it has only 48% BUSCO.

The CiT1 and CiP1 inocula could nodulate Coriaria species from both hemispheres, but not Datisca glomerata

Frankia cluster-2 inocula usually have a broad host range, which may be related to the fact that they contain more than one strain [14]. While CiT1 could nodulate the New Zealand species Coriaria arborea, and CiP1 could nodulate the Mediterranean species C. myrtifolia, neither CiT1 nor CiP1 could nodulate Datisca glomerata (Table 3). The Frankia MAGs of nodules induced by CiP1 on C. myrtifolia were sequenced (Tables 1 and 2) and shown to differ significantly from the MAGs isolated from the samples CiT1, CiP2-CiP5, and Cj2-Cj5 (Fig. 1).

Table 3 Host specificity analysis for different inocula. Data on Cppng1 and Dg1 were presented in Nguyen et al. [14]. PNG: Papua New Guinea. (nodulation)* - the MAG that was sequenced in the induced nodules belonged to a different lineage than the MAGs present in nodules induced on the original host

Cluster-2 Frankia strains from Papua New Guinea, the Philippines, Taiwan and Japan form a common lineage

To analyse the phylogeny of Frankia cluster-2, a core genome tree was constructed using the genomes of the type strains of all Frankia species available thus far as well as the MAGs obtained in this study. The results show that within Frankia cluster-2, two different lineages can be identified. On the one hand, Candidatus Frankia meridionalis Cppng1, the strains from Taiwan and the Philippines, and the novel strains from Japan form a common lineage. This lineage is distinct from the strains from the Eurasian continent represented by Candidatus Frankia datiscae Dg1 and Frankia coriariae BMG5.1, and the North American strains represented by Candidatus Frankia californiensis Dg2 (Fig. 1). Thus, we are calling the lineage represented by the MAGs Cppng1_Ca_nod, CiT1_Ci_nod, CiP2_Ci_nod to CiP5_Cj_nod, and Cj2_Cj_nod to Cj5_Cj_nod the ‘island lineage’ of Frankia cluster-2, in contrast with the ‘continental lineage’ represented by Candidatus F. datiscae, F. coriariae, Candidatus F. californiensis, and CiP1_Cm_nod1 and CiP1_Cm_nod2. Within the island lineage, the strains from the Philippines, Taiwan and Japan form a separate clade from Candidatus F. meridionalis.

CiP1_Cm_nod1 and CiP1_Cm_nod2, however, the Frankia MAGs from nodules induced by the CiP1 inoculum on the Mediterranean Coriaria species C. myrtifolia, clearly represented members of the continental lineage of Frankia cluster-2. Like members of F. coriariae [11, 16], they did not contain the canonical nod genes (Table 2). To ensure that no samples had been mixed up, we amplified and sequenced the plant matK phylogenetic markers from the raw data of the MAGs and confirmed host plant identity (Supplementary Table S3). In this context, we also wanted to confirm that the nodules induced by the inoculum from Papua New Guinea, Cppng1, on the Chinese Coriaria species C. terminalis [14] contained a MAG representing the island lineage. These nodules had been previously examined for nod gene expression using primers designed based on the Cppng1_Ca_nod sequence, indicating that indeed, the same strain was present in nodules of C. arborea and C. terminalis [14]. To settle any doubts, we sequenced the MAG from C. terminalis nodules (Cppng1_Ct_nod; Tables 1 and 2) which indeed showed 99.5% Average Nucleotide Identity (ANI) and 99.75% Average Amino Acid Identity (AAI) with Cppng1_Ca_nod (Supplementary Fig. S2; Supplementary Table S4A, B).

Inocula from the Philippines and Japan can contain Frankia species from both the island lineage and the continental lineage

Average Nucleotide Identity (ANI) comparisons were performed for the 10 novel MAGs of cluster-2 included in the phylogenetic tree (Supplementary Fig. S2, Supplementary Table S4). Based on the usually applied ANI threshold range of 95–96% for species demarcation [28, 29], the mean ANI values presented in Supplementary Fig. S2 and Supplementary Table S4A show that the genomes from Japan, Taiwan and the Philippines represent a novel species. Because the mean ANI values for two MAGs with low BUSCO values, CiP2_Ci_nod and CiP5_Ci_nod, were below 95%, the analysis was repeated using AAI values (Supplementary Table S4B). Based on the results, all novel Frankia MAGs from Japan, Taiwan and the Philippines analysed in this study belong to the same species, different from Candidatus Frankia meridionalis. The only exceptions were the MAGs of CiP1_Cm_nod1 and CiP1_Cm_nod2, which belonged to the continental lineage (Fig. 1, Supplementary Fig. S2). They showed less than 90.1% ANI with those of Frankia coriariae BMG5.1 [30] or with Candidatus Frankia datiscae Dg1 [12] (Supplementary Fig. S2), indicating they form a novel species. Since nodulation of C. terminalis by the CiP1 inoculum failed (Table 3), these Eurasian lineage strains seemed to have a very narrow host specificity, or, more likely, they represented a very minor contribution to the CiP1 strain assemblage. Nodulation of D. glomerata by CiP1 also failed (Table 3), but this was expected since like F. coriariae BMG5.1, CiP1_Cm_nod1 and CiP1_Cm_nod2 do not contain the canonical nod genes and in our hands, no inoculum without the canonical nod genes could ever nodulate D. glomerata [14].

Altogether, the strain assemblage that made up the inoculum CiP1 contained both representatives of a novel species of the island lineage and a novel species of the continental lineage. The island lineage was present in five different sets of C. intermedia nodules, while no sequences of a member of the continental lineage were obtained from nodules of this species (Supplementary Fig. S2; Fig. 2). It thus seems that while the strain assemblages contain both lineages, the members of the island lineage routinely outcompete the ones of the continental one when it comes to nodule induction on C. intermedia. It is likely that the members of the continental lineage are present on the outside of nodules. This has previously been shown for the cluster-3 species Frankia irregularis [31], members of which had been isolated from nodules of Casuarina species from different continents, but which cannot nodulate the Casuarina genus [32,33,34].

Fig. 2
figure 2

Map illustrating the location of Frankia samples and their host specificity. Black symbol: nodulation is possible. White symbol: nodulation was tried and was unsuccessful. Square: Coriaria myrtifolia or Coriaria nepalensis; diamond: Coriaria terminalis; down-facing triangle: Coriaria japonica or Coriaria intermedia; up-facing triangle: Coriaria sp. from the Southern hemisphere lineage; hexagon: Datisca glomerata. Colored dots of sampling represent species of Frankia as followed: magenta: Candidatus Frankia californiensis; green: Candidatus Frankia datiscae; blue: novel island lineage species presented in this study; yellow: Candidatus Frankia meridionalis, also part of the island lineage

The Frankia MAGs from nodules of D. glomerata induced by an inoculum of C. japonica nodules (Cj1_Dg_vc and Cj1_Dg_nod) had been found to represent members of the species Ca. F. datiscae of the continental lineage; Cj1_Dg_nod also contained another genome that could not be separated bioinformatically [14]. Furthermore, the type strain of the species F. coriariae, BMG5.1, was isolated from nodules of C. myrtifolia which were induced by crushed nodules of C. japonica collected in Japan [11, 30]. Thus, inocula from C. japonica can contain strains of the continental lineage. Here we show that direct sequencing of Frankia-enriched metagenomes from C. japonica nodules yields representatives of the island lineage of Frankia cluster-2 (Table 2, Fig. 1, Fig. 2). In summary, C. japonica inocula contain strains of the continental as well as of the island lineage of Frankia cluster-2, but the latter outcompetes the former on C. japonica.

What distinguishes the genomes of the island lineage from those of the continental lineage?

The MAGs of the island lineage all have a lower GC content than those of the continental lineage of Frankia cluster-2 (Table 2). A search for genes appearing in all MAGs available of members of the island lineage, but not in those of the continental lineage of cluster-2, revealed 230 genes (Supplementary Table S5A). The most interesting result was that all MAGs of the island lineage sequenced thus far contain the gene for 2-oxoglutarate dioxygenase (ethylene-forming; efe), first reported for Pseudomonas syringae strains [35, 36], later also for pathogenic fungi [37]. Ethylene production has been directly implicated in the pathogenicity of P. syringae pv. glycinea, though not of P. syringae pv. phaseolicola [38]. It is surprising that this enzyme is found in a plant symbiont. Based on Johansson et al. [37], the enzymes of strains of the island lineage of Frankia cluster-2 have all amino acid residues relevant for enzyme function (Supplementary Fig. S3). Upon investigation, we found that the efe gene was expressed in Frankia, in field nodules collected from C. japonica (Supplementary Fig. S4).

For legumes, it has been shown that ethylene inhibits symbiotic signalling [39] and also generally reduces nodulation via the root hair infection pathway [40]. This is confirmed by the fact that providing an enzyme that degrades the plant precursor of ethylene, 1-aminocyclopropane-1-carboxylate deaminase (AcdS), via the rhizobial microsymbionts themselves [41] or via rhizosphere bacteria [42], can improve nodulation. However, ethylene positively affects nodulation of peanut, which follows an intercellular infection pathway [43] and is required for nodulation of Sesbania rostrata via crack entry [44]. Thus, the negative effect of ethylene seems to be specific to nodulation via root hairs. Altogether, the presence and expression of efe genes in the MAGs of the strains of the island lineage suggest that nodulation by cluster-2 strains, at least in the case of Coriaria spp., requires ethylene and presumably follows an intercellular pathway.

A search for genes appearing in all MAGs available of members of the continental lineage, but not in MAGs of the island lineage of cluster-2, revealed 218 genes (Supplementary Table S5B). Only the MAGs from the continental lineage contain copies of the mammalian cell entry (mce) gene cluster from other actinobacteria like Mycobacterium tuberculosis (AEH09479 – AEH09484 in Candidatus Frankia datiscae Dg1) which in Nocardia farcinia was implicated in the interaction with, and invasion of, mammalian cells, and in Streptomyces coelicolor is required for plant root colonization [45, 46]. Furthermore, genes encoding the sodium/proton antiporter NhaA, which are present in all representatives of the continental lineage (AEH09573 and AEH09214 in Candidatus F. datiscae Dg1), were not found in the island lineage (Supplementary Table S5B), which could indicate lower levels of salt tolerance of the island lineage strains.

How did Frankia cluster-2 strains spread from Gondwana across the world?

As mentioned above, the common ancestor of all root nodule-forming plants including the actinorhizal Cucurbitales evolved in Gondwana. Based on a fossil-dated phylogeny [7], a northern hemisphere clade of Coriariaceae (NH) diverged from the southern hemisphere clade of Coriariaceae (SH) in the Paleocene (ca. 57 Mya). The NH clade is split in Coriaria nepalensis, Coriaria myrtifolia and Coriaria terminalis on the one hand, and Coriaria japonica and Coriaria intermedia on the other hand. The SH clade of Coriariaceae encompasses the South American species Coriaria ruscifolia, the eight Coriaria species from New Zealand, and Coriaria papuana from Papua New Guinea. Why do the two parts of the NH clade of Coriaria have microsymbionts from different lineages of Frankia cluster-2, i.e. the continental or the island lineage?

The other family of actinorhizal Cucurbitales, Datiscaceae, evolved in India. This is based on a fossil wood of a precursor of the closely related Tetramelaceae, Tetramelioxylon prenudiflora, which has been found in the Deccan Intertrappean beds of Mohgaonkalan, Madhya Pradesh, India [47]. Thus, some Cucurbitales spread from Gondwana to Eurasia via India. The route via India would also be consistent with what is known about the phylogeny of the NH Coriariaceae [7].

Three scenarios are possible to explain how the two parts of the NH clade of Coriaria ended up with microsymbionts from different lineages of Frankia cluster-2. In the first one, Datisca dispersed to Eurasia via India together with its inoculum which represented the precursor of the continental lineage of Frankia cluster-2. The fact that Frankia cluster-2 strains from Papua-New Guinea, the Philippines, Taiwan and Japan belong to a common lineage, separate from the continental lineage, might suggest that Frankia cluster-2 spread from the SH to the NH via Papua New Guinea and Indonesia to the Philippines. The only extant host plant that fits this distribution would be Coriaria. Coriaria would then have dispersed to the mainland and would have eventually spread to areas where Datisca was growing. The strains of the continental lineage could have outcompeted the strains of the island lineage. This is unlikely based on the timing: given the changing distances between the islands involved, the proposed spread of Coriaria from the SH to the NH could have happened at the earliest near the end of the Oligocene 23 Mya. However, there is a Coriaria leaf fossil found in the Armissan bed in France which dates back to 23.2 to 33.9 Mya [7, 48]. Furthermore, this scenario does not account for the presence of the continental lineage of Frankia cluster-2 in Japan and the Philippines. In summary, this scenario can be excluded.

In the second scenario, Datisca and Coriaria came to Eurasia via India with Coriaria hosting both lineages of Frankia cluster-2. Eventually, the continental lineage outcompeted the island lineage because it was better adapted to the local soil or climate. However, the island lineage strains remained part of the strain assemblages. When Coriaria had spread from the mainland to the East Asian islands, the island lineage began to outcompete the continental lineage. This scenario is unlikely because it would have required uniform conditions on the Eurasian continent.

In the third scenario, Datisca and Coriaria came to Eurasia via India, both with the continental lineage of Frankia cluster-2. In the Miocene, the island lineage spread northward from New Zealand or Papua New Guinea with another host plant. When Coriaria spread from the mainland to Japan/Taiwan/Philippines, the plants encountered the island lineage of Frankia cluster-2 which mostly outcompeted the continental lineage. Nevertheless, strains of the continental lineage remained part of some strain assemblages. The other host plant subsequently lost the symbiosis or became extinct. This is the most likely scenario and is depicted in Fig. 3.

Fig. 3
figure 3

How cluster-2 Frankia strains spread from Gondwana across the world. (A) Geography from 100 mya, (B) current geography. Geography is according to the plate reconstruction of Zahirovic et al. [49] and paleo-environments from Cao et al. [50]. Land is given in green, deep sea in dark blue, and shallow sea as light blue. For reference, present-day coastlines and geological terrane boundaries are reconstructed using brown lines. Black lines indicate tectonic plate boundaries. Spread of Coriaria with cluster-2 Frankia is indicated by red arrows, hypothetical alternatives are indicated by dashed arrows. Coriaria nodule sampling points referred to in this study are indicated by white triangles. (A) Coriaria spread with the future continental cluster-2 lineage from Gondwana to India (1; continental lineage of Frankia cluster-2), and from Gondwana to South America (2; South American lineage). Distribution from Gondwana to New Zealand and Papua New Guinea (PNG; 3; island lineage of Frankia cluster-2) could have taken place via Australia (3a) or via New Zealand (3b) or in both directions simultaneously. (B) Distribution between PNG and New Zealand is ascertained by the fact that Coriaria is indigenous in all islands of the area [7], but the direction cannot be determined at this point (dashed line). When India had collided with Asia, Coriaria spread in Northern India-Pakistan-Nepal and from there westward to the Mediterranean and eastward into China. The precursor of C. nepalensis spread to the North (dashed arrow, since Coriaria is not found in Northern China today) towards Japan (C. japonica), and from there southward to Taiwan and the Philippines (C. intermedia); the separation of C. japonica and C. intermedia was recent (ca. 10 mya) [7]. The island lineage of Frankia cluster-2 spread from PNG to the Philippines and further to Taiwan (blue arrow), but the original host plants lost the symbiosis or went extinct. Strains of the island lineage outcompeted those of the continental lineage of Frankia when Coriaria spp. spread to Japan and from there to Taiwan and the Philippines. All cluster-2 Frankia strains sampled in continental Eurasia belong to the continental lineage. MD, Madagascar; PH, Philippines; PNG, Papua New Guinea; TW, Taiwan

How could the SH Coriaria clade and its microsymbiont come to Papua New Guinea? Based on the phylogeny of Renner et al. [7], Coriaria seems to have spread from Gondwana directly to New Zealand (see arrow 3a in Fig. 3A), and later, based on the position of C. papuana in the phylogeny, via Papua-New Guinea. Due to the disjunct distribution of the genus, the plant phylogeny cannot show how Coriaria spread from Gondwana to Papua-New Guinea; two interpretations are possible. Either Coriaria spread early to Papua-New Guinea via Australia where the genus would then have become extinct (see arrow 3b in Fig. 3A), or it spread from New Zealand to Papua-New Guinea, and later plants from Papua-New Guinea dispersed back to New Zealand.

Conclusions

Strains of the earliest divergent Frankia clade, cluster-2, split into two lineages in Gondwana. One lineage came to the northern hemisphere via India with Coriaria sp. and the precursors of Datiscaceae. The other lineage spread in the southern hemisphere but eventually spread to the northern hemisphere when the distance between Papua New Guinea/New Guinea and the northern hemisphere South East Asian islands had reduced in the Miocene.

Loss of the symbiosis, or extinction of a symbiotic plant species, should have taken place for a non-Coriaria Frankia cluster-2 host after the split of the northern hemisphere Coriaria species in C. myrtifolia, C. nepalensis, C. terminalis on the one hand and C. japonica, C. intermedia on the other hand, i.e., at the earliest in the Miocene. This might be consistent with the hypothesis [8] that the decrease in atmospheric CO2 in the Oligocene and Miocene accounted for the loss of nodulation.

Analysis of inocula from Coriaria spp. nodules from New Zealand will have to show how cluster-2 Frankia strains reached this island.

Methods

Plant and bacterial material

Nodules of Coriaria intermedia Matsum. were collected at Taiping Mountain, Taiwan (24°29′51.5″ N 121°32′07.9″ E) in January 2017. Leaves were collected from the sample plant on July 8, 2018. A voucher was deposited in the herbarium of the Swedish Museum of Natural History, leg. K. Pawlowski s.n. (S; Reg. No. S18–40315). Nodules of C. intermedia were collected in April 2018 in Poblacion, Atok, Benguet/Sto. Tomas, Tuba, Benguet-CAR, Philippines. Vouchers were deposited in the Herbarium at the Department of Botany, University of the Philippines, leg. C.M. Bandong s.n. (Reg. No. 21414) and in the herbarium of the Swedish Museum of Natural History, leg. K. Pawlowski s.n. (S; Reg. No. S19–5452). Nodules of Coriaria japonica A. Gray were collected in RNAlater (Sigma-Aldrich, Japan) on the riverbanks of Nikko City (36°44′ N, 139°37′ E; 520 m a.s.l.) in central Japan by Sae Katayama and Masaki Tateno. Nodules were also collected in Oshu city (39°12′01.2“ N 141°23’48.1” E) in Iwate province in the northern part of Honshu, Japan, by Takashi Yamanaka in absolute ethanol. All nodules were immediately frozen and kept at − 20 °C upon arrival in Stockholm, Sweden. Vouchers were deposited in the herbarium of the Swedish Museum of Natural History, leg. F. Berckx s.n. (Reg. No. S22–86 and S22–87).

Datisca glomerata (C. Presl) Baill., Coriaria myrtifolia L., Coriaria terminalis Hemsl., Coriaria arborea Linds., and C. intermedia Matsum. were germinated from seeds in a greenhouse at 13 h light / 11 h dark, 25 °C in the light phase and 19 °C in the dark phase. D. glomerata seeds were obtained from plants in the greenhouse, going back to seeds collected in Vaca Hills, California. Seeds of C. myrtifolia were kindly provided by Paul Goetghebeur from the Botanical Garden at Gent University (Gent, Belgium). Seeds of C. terminalis var. xanthocarpa were purchased from www.plant-world-seeds.com. Seeds of C. arborea were kindly provided by Warwick Silvester. Seeds of C. intermedia were collected in Poblacion, Atok (Benguet-CAR, Philippines). After eight (D. glomerata) or twelve (Coriaria spp.) weeks, respectively, plantlets were inoculated with crushed nodules collected at Taiping Mountain (Taiwan) or in Poblacion, Atok (Philippines) and plants were maintained as described by Nguyen et al. [14]. Three to five months after inoculation, plants were inspected for nodulation status, and nodules were harvested into liquid nitrogen and kept at -80 °C.

DNA isolation and sequencing

Total nodule DNA was isolated as described by Nguyen et al. [14] with small modifications. The nodules of C. japonica in the field were collected in RNAlater (Sigma-Aldrich, Japan) or ethanol, which was removed by gently washing with sterile milliQ water, followed by patting the nodules dry. DNA was isolated from C. japonica nodules using the NucleoSpin Plant II kit (Macherey-Nagel, Sweden). For isolation of DNA from nodules from C. intermedia, C. terminals, and C. myrtfolia, the GenElute™ Plant Genomic DNA Miniprep Kit (Sigma-Aldrich, Sweden) was used.

The genomic sequencing library of CiT1_Ci_nod was constructed from 1 ng of gDNA with the Nextera XT DNA Sample Preparation Kit (Illumina, Germany) according to the manufacturer’s protocol. The library was quality controlled by analysis on an Agilent 2000 Bioanalyzer with the Agilent High Sensitivity DNA Kit (Agilent Technologies, Germany) for fragment sizes of ca. 500–800 bp. Sequencing on a MiSeq sequencer (Illumina; 2 × 250 bp paired-end sequencing, v3 chemistry) was performed in the Genomics Service Unit (LMU Biocenter, Martinsried, Germany).

Libraries for genome sequencing of CiP2_Ci_nod, CiP3_Ci_nod, CiP4_Ci_nod, CiP5_Ci_nod, CiP6_Ci_nod, CiP1_Cm_nod, CiP1_Cm_nod2, Cj2_Cj_nod, Cj3_Cj_nod, Cj4_Cj_nod, Cj5_Cj_nod, and Cppng1_Ct_nod were constructed from 1 ng of DNA, sheared on Covaris M220 with Covaris MicroCaps 50 μl to approx. 600 bp, using the NEBNext Ultra II DNA Library Kit Kit (New England Biolabs, Germany) and the sparQ DNA Library Prep Kit (QuantaBio, MA, USA) following the manufacturer’s protocols. Libraries were quality controlled by analysis on an Agilent 2000 Bioanalyzer with the Agilent High Sensitivity DNA Kit (Agilent Technologies) for fragment sizes of ca. 500–800 bp. Sequencing was performed on a MiSeq sequencer as described above.

Genome assembly and bioinformatics analyses

Assembly, binning and annotation were done as previously described [13, 14] with small modifications. In brief, de novo assembly was performed by applying the gsAssember 2.8. (Roche) with default settings. Raw reads were aligned to the corresponding assembled (meta-)genome contigs using Bowtie 2 (v2.4.1 [51];). The resulting SAM files were processed through SAMtools (v1.0 [52];). For binning, MetaBAT2 (v2.12.1 [53];) was applied with default settings. Raw reads were exported by mapping to the resulting Frankia bins and reassembled using again the gsAssembler 2.8 (Roche) with default settings. Completeness, contamination, and strain heterogeneity were estimated with BUSCO (v3.0.2 [54];), using the bacterial-specific single-copy marker genes database (odb9). For the annotation of the genomes, Prokka [55] and GenDB [56] were applied at default settings. Draft genome sequences were deposited at the EMBL/GenBank/DDBJ databases in BioProjects PRJEB47857 (CiT1), PRJEB47848 (CiP2), PRJEB47853 (CiP3), PRJEB47854 (CiP4), PRJEB47855 (CiP5), PRJEB47858 (Cj2). PRJEB47859 (Cj3), PRJEB47860 (Cj4), PRJEB47861 (Cj5), PRJEB48851 (CiP1_Cm_nod1), PRJEB48852 (CiP1_Cm_nod2). Core genome trees were inferred, average nucleotide identity was calculated, and genome comparisons were performed using the EDGAR 2.0 platform [23, 24]. Pairwise genome-to-genome distance calculations [digital DNA–DNA hybridization (dDDH)] were performed as recently described [57].

Phylogenetic analysis of host plants

The phylogeny of the host plants was analysed based on the combination of the nuclear internal transcribed spacer (ITS) region, the large subunit of the ribulose-bisphosphate carboxylase gene (rbcL), maturase K (matK), or the chloroplast trnL gene. Primers were used as described in previous studies [7, 21]. All primers used in this study are listed in Supplementary Table S1.

RNA isolation and quantitative reverse transcription polymerase chain reaction (RT-qPCR)

Total RNA was isolated from C. japonica nodules as described previously [14, 58] with some small modifications. In brief, nodules were washed in sterile MilliQ water to remove excess RNAlater (Sigma-Aldrich, Japan), and patted dry, and ground in liquid nitrogen with the addition of Polyclar AT. In the lysis buffer, samples were subjected to ultrasonication for three rounds at 30% pulsing, 25 s each round (ultrasonic homogenizer Sonoplus HD 2070, Bandelin Electronic, Berlin, Germany). RNA was extracted using the Spectrum Plant Total RNA kit from Sigma–Aldrich (Stockholm, Sweden) with on-column gDNA digestion (Sigma-Aldrich) according to the manufacturer’s instructions. Total removal of gDNA was verified using PCR with primers targeting the housekeeping gene infC [59], using cDNA of Cppng_Ca_nod as a positive control [14]. cDNA was then synthesized using TATAA GrandScript cDNA Synthesis Kit (TATAA, Sweden). Primers for the housekeeping gene infC, nitrogenase gene nifD, and 2-oxoglutarate dioxygenase (ethylene-forming) gene efe were designed using the primer design tool Primer3, available through NCBI. Primer efficiency between 90 and 110% was verified for each pair, and the best out of two designed pairs was chosen. Primer sequences are given in Supplementary Table S1. Each qPCR reaction contained 1x Maxima SYBR Green/ROX qPCR Master Mix (ThermoFisher Scientific), 300 nM of each primer, and 4 ng of cDNA in a reaction volume of 10 μl. The conditions of qPCR were as follows: 10 min at 95 °C, followed by 40 cycles of 15 s at 95 °C, 30 s at 60 °C, followed by a melt curve program of 15 s at 95 °C, 15 s at 60 °C, and 15 s at 95 °C. Gene expression values were normalized against infC, the gene encoding translation initiation factor IF3. The statistical analysis and data visualisation were performed in Rstudio [25].

Availability of data and materials

Draft genome sequences were deposited at the EMBL/GenBank/DDBJ databases in BioProjects PRJEB47857 (CiT1), PRJEB47848 (CiP2), PRJEB47853 (CiP3), PRJEB47854 (CiP4), PRJEB47855 (CiP5), PRJEB47858 (Cj2), PRJEB47859 (Cj3), PRJEB47860 (Cj4), PRJEB47861 (Cj5), PRJEB48851 (CiP1_Cm_nod1), PRJEB48852 (CiP1_Cm_nod2).

References

  1. Soltis DE, Soltis PS, Morgan DR, Swensen SM, Mullin BC, Dowd JM, et al. Chloroplast gene sequence data suggest a single origin of the predisposition for symbiotic nitrogen fixation in angiosperms. Proc Natl Acad Sci U S A. 1995;92(7):2647–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Doyle JJ. Phylogenetic perspectives on the origins of nodulation. Mol Plant-Microbe Interactions. 2011;24(11):1289–95.

    Article  CAS  Google Scholar 

  3. Werner GDA, Cornwell WK, Sprent JI, Kattge J, Kiers ET. A single evolutionary innovation drives the deep evolution of symbiotic N2-fixation in angiosperms. Nat Commun. 2014;5(1):4087.

    Article  CAS  PubMed  Google Scholar 

  4. Griesmann M, Chang Y, Liu X, Song Y, Haberer G, Crook MB, et al. Phylogenomics reveals multiple losses of nitrogen-fixing root nodule symbiosis. Science. 2018;361(6398):eaat1743.

  5. van Velzen R, Holmer R, Bu F, Rutten L, van Zeijl A, Liu W, et al. Comparative genomics of the nonlegume Parasponia reveals insights into evolution of nitrogen-fixing rhizobium symbioses. Proc Natl Acad Sci U S A. 2018;115(20):E4700–9.

    PubMed  PubMed Central  Google Scholar 

  6. Cook LG, Crisp MD. Not so ancient: the extant crown group of Nothofagus represents a post-Gondwanan radiation. Proc R Soc B Biol Sci. 2005;272(1580):2535–44.

    Article  Google Scholar 

  7. Renner SS, Barreda VD, Tellería MC, Palazzesi L, Schuster TM. Early evolution of Coriariaceae (Cucurbitales) in light of a new early Campanian (ca. 82 Mya) pollen record from Antarctica. Taxon. 2020;69(1):87–99.

    Article  Google Scholar 

  8. van Velzen R, Doyle JJ, Geurts R. A resurrected scenario: single gain and massive loss of nitrogen-fixing nodulation. Trends Plant Sci. 2019;24(1):49–57.

    Article  PubMed  CAS  Google Scholar 

  9. Pawlowski K, Demchenko KN. The diversity of actinorhizal symbiosis. Protoplasma. 2012;249(4):967–79.

    Article  PubMed  Google Scholar 

  10. Sen A, Daubin V, Abrouk D, Gifford I, Berry AM, Normand P. Phylogeny of the class Actinobacteria revisited in the light of complete genomes. The orders “Frankiales” and Micrococcales should be split into coherent entities: proposal of Frankiales ord. nov., Geodermatophilales ord. nov., Acidothermales ord. nov. and Nakamurellales ord. nov. Int J Syst Evol Microbiol. 2014;64(Pt 11):3821–32.

  11. Gtari M, Ghodhbane-Gtari F, Nouioui I, Ktari A, Hezbri K, Mimouni W, et al. Cultivating the uncultured: growing the recalcitrant cluster-2 Frankia strains. Sci Rep. 2015;5(1):13112.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Persson T, Battenberg K, Demina IV, Vigil-Stenman T, Heuvel BV, Pujic 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(5):e0127630.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  13. Nguyen TV, Wibberg D, Battenberg K, Blom J, Vanden Heuvel B, Berry AM, et al. An assemblage of Frankia cluster II strains from California contains the canonical nod genes and also the sulfotransferase gene nodH. BMC Genomics. 2016;17(1):796.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Nguyen TV, Wibberg D, Vigil-Stenman T, Berckx F, Battenberg K, Demchenko KN, et al. Frankia-enriched metagenomes from the earliest diverging symbiotic Frankia cluster: they come in teams. Genome Biol Evol. 2019;19:2273–91.

    Article  CAS  Google Scholar 

  15. Battenberg K, Wren JA, Hillman J, Edwards J, Huang L, Berry AM. The influence of the host plant is the major ecological determinant of the presence of nitrogen-fixing root nodule symbiont cluster II Frankia species in soil. Appl Environ Microbiol. 2017;83(1):e02661–16.

    Article  CAS  PubMed  Google Scholar 

  16. Gueddou A, Swanson E, Hezbri K, Nouioui I, Ktari A, Simpson S, et al. Draft genome sequence of the symbiotic Frankia sp. strain BMG5.30 isolated from root nodules of Coriaria myrtifolia in Tunisia. Antonie Van Leeuwenhoek 2019;112(1):67–74.

  17. Hugerth LW, Larsson J, Alneberg J, Lindh MV, Legrand C, Pinhassi J, et al. Metagenome-assembled genomes uncover a global brackish microbiome. Genome Biol. 2015;16(1):279.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Benson DR, Dawson JO. Recent advances in the biogeography and genecology of symbiotic Frankia and its host plants. Physiol Plant. 2007;130(3):318–30.

    Article  CAS  Google Scholar 

  19. Clement WL, Tebbitt MC, Forrest LL, Blair JE, Brouillet L, Eriksson T, et al. Phylogenetic position and biogeography of Hillebrandia sandwicensis (Begoniaceae): a rare Hawaiian relict. Am J Bot. 2004;91(6):905–17.

    Article  CAS  PubMed  Google Scholar 

  20. Davidson C. An anatomical and morphological study of Datiscaceae. Aliso 8(1):49–110.

  21. Yokoyama J, Suzuki M, Iwatsuki K, Hasebe M. Molecular phylogeny of Coriaria, with special emphasis on the disjunct distribution. Mol Phylogenet Evol. 2000;14(1):11–9.

    Article  CAS  PubMed  Google Scholar 

  22. Felsenstein J. PHYLIP (phylogeny inference package), version 3.5 c. Joseph Felsenstein; 1993.

  23. Blom J, Albaum SP, Doppmeier D, Pühler A, Vorhölter FJ, Zakrzewski M, et al. EDGAR: a software framework for the comparative analysis of prokaryotic genomes. BMC Bioinformatics. 2009;10:154.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Blom J, Kreis J, Spänig S, Juhre T, Bertelli C, Ernst C, et al. EDGAR 2.0: an enhanced software platform for comparative gene content analyses. Nucleic Acids Res. 2016;44(W1):W22–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. RStudio Team. RStudio: integrated development environment for R [internet]. Boston: RStudio, PBC; 2022. Available from: http://www.rstudio.com/

  26. Paradis E, Claude J, Strimmer K. APE: analyses of phylogenetics and evolution in R language. Bioinforma Oxf Engl. 2004;20(2):289–90.

    Article  CAS  Google Scholar 

  27. Schliep KP. Phangorn: phylogenetic analysis in R. Bioinforma Oxf Engl. 2011;27(4):592–3.

    Article  CAS  Google Scholar 

  28. Goris J, Konstantinidis KT, Klappenbach JA, Coenye T, Vandamme P, Tiedje JM. DNA-DNA hybridization values and their relationship to whole-genome sequence similarities. Int J Syst Evol Microbiol. 2007;57(Pt 1):81–91.

    Article  CAS  PubMed  Google Scholar 

  29. Richter M, Rosselló-Móra R. Shifting the genomic gold standard for the prokaryotic species definition. Proc Natl Acad Sci U S A. 2009;106(45):19126–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Nouioui I, Ghodhbane-Gtari F, Rohde M, Klenk HP, Gtari M. Frankia coriariae sp. nov., an infective and effective microsymbiont isolated from Coriaria japonica. Int J Syst Evol Microbiol. 2017;67(5):1266–70.

    Article  CAS  PubMed  Google Scholar 

  31. Vemulapally S, Guerra T, Hahn D. Localization of typical and atypical Frankia isolates from Casuarina sp. in nodules formed on Casuarina equisetifolia. Plant Soil. 2019;435(1):385–93.

    Article  CAS  Google Scholar 

  32. Pujic P, Bolotin A, Fournier P, Sorokin A, Lapidus A, al KHR et. Genome sequence of the atypical symbiotic Frankia R43 strain, a nitrogen-fixing and hydrogen-producing actinobacterium. Genome Announc. 2015;3:e01387–15.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Mansour S, Swanson E, McNutt Z, Pesce C, Harrington K, Abebe-Alele F, et al. Permanent draft genome sequence for Frankia sp. strain CcI49, a nitrogen-fixing bacterium isolated from Casuarina cunninghamiana that infects Elaeagnaceae. J Genomics. 2017;5:119–23.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Nouioui I, Ghodhbane-Gtari F, Rhode M, Sangal V, Klenk HP, Gtari M. Frankia irregularis sp. nov., an actinobacterium unable to nodulate its original host, Casuarina equisetifolia, but effectively nodulates members of the actinorhizal Rhamnales. Int J Syst Evol Microbiol. 2018;68(9):2883–914.

    Article  CAS  PubMed  Google Scholar 

  35. Zhang Z, Smart TJ, Choi H, Hardy F, Lohans CT, Abboud MI, et al. Structural and stereoelectronic insights into oxygenase-catalyzed formation of ethylene from 2-oxoglutarate. Proc Natl Acad Sci U S A. 2017;114(18):4667–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Fukuda H, Ogawa T, Tazaki M, Nagahama K, Fujii T, Tanase S, et al. Two reactions are simultaneously catalyzed by a single enzyme: the arginine-dependent simultaneous formation of two products, ethylene and succinate, from 2-oxoglutarate by an enzyme from Pseudomonas syringae. Biochem Biophys Res Commun. 1992;188(2):483–9.

    Article  CAS  PubMed  Google Scholar 

  37. Johansson N, Persson KO, Larsson C, Norbeck J. Comparative sequence analysis and mutagenesis of ethylene forming enzyme (EFE) 2-oxoglutarate/Fe(II)-dependent dioxygenase homologs. BMC Biochem. 2014;15:22.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Weingart H, Ullrich H, Geider K, Völksch B. The role of ethylene production in virulence of Pseudomonas syringae pvs. Glycinea and phaseolicola. Phytopathology. 2001;91(5):511–8.

    Article  CAS  PubMed  Google Scholar 

  39. Oldroyd GE, Engstrom EM, Long SR. Ethylene inhibits the nod factor signal transduction pathway of Medicago truncatula. Plant Cell. 2001 Aug;13(8):1835–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Zhu F, Deng J, Chen H, Liu P, Zheng L, Ye Q, et al. A CEP peptide receptor-like kinase regulates auxin biosynthesis and ethylene signaling to coordinate root growth and symbiotic nodulation in Medicago truncatula. Plant Cell. 2020 Sep;32(9):2855–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Ma W, Charles TC, Glick BR. Expression of an exogenous 1-aminocyclopropane-1-carboxylate deaminase gene in Sinorhizobium meliloti increases its ability to nodulate alfalfa. Appl Environ Microbiol. 2004;70(10):5891–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Nascimento FX, Tavares MJ, Franck J, Ali S, Glick BR, Rossi MJ. ACC deaminase plays a major role in Pseudomonas fluorescens YsS6 ability to promote the nodulation of alpha- and Betaproteobacteria rhizobial strains. Arch Microbiol. 2019;201(6):817–22.

    Article  CAS  PubMed  Google Scholar 

  43. Muñoz VL, Figueredo MS, Reinoso H, Fabra A. Role of ethylene in effective establishment of the peanut-bradyrhizobia symbiotic interaction. Plant Biol. 2021;23(6):1141–8.

    Article  PubMed  CAS  Google Scholar 

  44. Goormachtig S, Capoen W, James EK, Holsters M. Switch from intracellular to intercellular invasion during water stress-tolerant legume nodulation. Proc Natl Acad Sci U S A. 2004;101(16):6303–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Clark LC, Seipke RF, Prieto P, Willemse J, van Wezel GP, Hutchings MI, et al. Mammalian cell entry genes in Streptomyces may provide clues to the evolution of bacterial virulence. Sci Rep. 2013;3:1109.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Ji X, Tan X, Hou X, Si C, Xu S, Tang L, et al. Cloning, expression, invasion, and immunological reactivity of a mammalian cell entry protein encoded by the mce1 operon of Nocardia farcinica. Front Microbiol. 2017;8:281.

    PubMed  PubMed Central  Google Scholar 

  47. Lakhanpal R, Verma J. Fossil wood of Tetrameles from the Deccan Intertrappean beds of Mohgaonkalan, Madhya Pradesh. Palaeobotanist 1962;14(1,2,3):209–13.

  48. De Saporta MG. Remarques sur les genres de végétaux actuels dont l’existence a été constatée a l’état fossile, leur ancienneté relative, leur distribution, leur marche et leur développement successifs. Bull Société Bot Fr. 1866;13(3):189–213.

    Article  Google Scholar 

  49. Zahirovic S, Matthews KJ, Flament N, Müller RD, Hill KC, Seton M, et al. Tectonic evolution and deep mantle structure of the eastern Tethys since the latest Jurassic. Earth-Sci Rev. 2016;162:293–337.

    Article  CAS  Google Scholar 

  50. Cao W, Zahirovic S, Flament N, Williams S, Jan G, Müller D. Improving global paleogeography since the late Paleozoic using paleobiology. Biogeosciences. 2017;14:5425–39.

    Article  Google Scholar 

  51. Langmead B, Salzberg SL. Fast gapped-read alignment with bowtie 2. Nat Methods. 2012;9(4):357–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, et al. The sequence alignment/map format and SAMtools. Bioinforma Oxf Engl. 2009;25(16):2078–9.

    Article  CAS  Google Scholar 

  53. Kang DD, Li F, Kirton E, Thomas A, Egan R, An H, et al. MetaBAT 2: an adaptive binning algorithm for robust and efficient genome reconstruction from metagenome assemblies. PeerJ. 2019;7:e7359.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Simão FA, Waterhouse RM, Ioannidis P, Kriventseva EV, Zdobnov EM. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinforma Oxf Engl. 2015;31(19):3210–2.

    Article  CAS  Google Scholar 

  55. Seemann T. Prokka: rapid prokaryotic genome annotation. Bioinforma Oxf Engl. 2014;30(14):2068–9.

    Article  CAS  Google Scholar 

  56. Meyer F, Goesmann A, McHardy AC, Bartels D, Bekel T, Clausen J, et al. GenDB--an open source genome annotation system for prokaryote genomes. Nucleic Acids Res 2003;31(8):2187–2195.

  57. Lick S, Wibberg D, Winkler A, Blom J, Grimmler C, Goesmann A, Kalinowski J, Kröckel L. Pseudomonas paraversuta sp. nov. isolated from refrigerated dry-aged beef. Int J Syst Evol Microbiol 2021;71:004822.

  58. Berckx F, Wibberg D, Kalinowski J, Pawlowski K. The peptidoglycan biosynthesis gene murC in Frankia: actinorhizal vs. plant type. Genes. 2020;11(4):E432.

  59. Alloisio N, Queiroux C, Fournier P, Pujic P, Normand P, Vallenet D, et al. The Frankia alni symbiotic transcriptome. Mol Plant-Microbe Interact. 2010;23(5):593–607.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are indebted to Susanne Renner (LMU Munich, Germany) for helpful discussions about the evolution and systematics of Cucurbitales, to Solveig Pospiech (Helmholtz-Zentrum Dresden-Rossendorf, Germany) for helpful discussions about palaeogeography, to Sabin Zahirovic (University of Sydney, Australia) for generous help with palaeogeographic maps and to Anna Zdyb (TU Dresden, Germany) for generous help with figure design. We thank Anna Pettersson and Johanna Grahn (Stockholm University) for taking care of the plants in Stockholm. We would like to thank Warwick Silvester and Paul Goetghebeur for providing seeds of Coriaria spp. and Ciara Morrison (Stockholm University) for help with the RT-qPCR analysis.

Funding

Open access funding provided by Stockholm University. This project was supported by two grants from the Swedish Research Council Vetenskapsrådet (VR 2012–03061 and 2019–05540) to KP. The bioinformatics support of the BMBF-funded project “Bielefeld-Gießen Center for Microbial Bioinformatics” (BiGi) and the BMBF grant FKZ 031A533 within the German Network for Bioinformatics Infrastructure (de.NBI) are gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Contributions

CMB, HHL, TY, CTL, MT and SK collected the field material. FB, TVN, and CMB performed the experiments. AB performed the sequencing. Genome assembly was performed by DW. Genome analysis was performed by DW, JB, FB, KP, CMB. The initial manuscript was written by FB and KP. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Katharina Pawlowski.

Ethics declarations

Ethics approval and consent to participate

Does not apply.

Consent for publication

Does not apply.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Berckx, F., Nguyen, T.V., Bandong, C.M. et al. A tale of two lineages: how the strains of the earliest divergent symbiotic Frankia clade spread over the world. BMC Genomics 23, 602 (2022). https://doi.org/10.1186/s12864-022-08838-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-022-08838-5

Keywords