Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-25T16:23:07.442Z Has data issue: false hasContentIssue false

Evolutionary genomic insights into cyanobacterial symbioses in plants

Published online by Cambridge University Press:  08 August 2022

Sophie de Vries*
Affiliation:
Department of Applied Bioinformatics, Institute for Microbiology and Genetics, University of Goettingen, Goettingen, Germany
Jan de Vries*
Affiliation:
Department of Applied Bioinformatics, Institute for Microbiology and Genetics, University of Goettingen, Goettingen, Germany Goettingen Center for Molecular Biosciences (GZMB), University of Goettingen, Goettingen, Germany Campus Institute Data Science (CIDAS), University of Goettingen, Goettingen, Germany
*
Authors for correspondence: Sophie de Vries E-mail: sophie.devries@uni-goettingen.deJan de Vries E-mail: devries.jan@uni-goettingen.de
Authors for correspondence: Sophie de Vries E-mail: sophie.devries@uni-goettingen.deJan de Vries E-mail: devries.jan@uni-goettingen.de

Abstract

Photosynthesis, the ability to fix atmospheric carbon dioxide, was acquired by eukaryotes through symbiosis: the plastids of plants and algae resulted from a cyanobacterial symbiosis that commenced more than 1.5 billion years ago and has chartered a unique evolutionary path. This resulted in the evolutionary origin of plants and algae. Some extant land plants have recruited additional biochemical aid from symbiotic cyanobacteria; these plants associate with filamentous cyanobacteria that fix atmospheric nitrogen. Examples of such interactions can be found in select species from across all major lineages of land plants. The recent rise in genomic and transcriptomic data has provided new insights into the molecular foundation of these interactions. Furthermore, the hornwort Anthoceros has emerged as a model system for the molecular biology of cyanobacteria–plant interactions. Here, we review these developments driven by high-throughput data and pinpoint their power to yield general patterns across these diverse symbioses.

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press in association with The John Innes Centre

1. Cyanobacterial symbioses in land plants—an overview

Chloroplasts emerged as a result of a cyanobacterial symbiosis. From a single endosymbiotic event about 1.5 bya (Bengtson et al., Reference Bengtson, Sallstedt, Belivanova and Whitehouse2017; Butterfield, Reference Butterfield2000; Eme et al., Reference Eme, Sharpe, Brown and Roger2014), almost the entire diversity of photosynthetic eukaryotes radiated (Archibald, Reference Archibald2015). The nature of that cyanobacterial plastid progenitor is still debated (de Vries & Archibald, Reference de Vries and Archibald2017; Ponce-Toledo et al., Reference Ponce-Toledo, Deschamps, López-García, Zivanovic, Benzerara and Moreira2017); while some analyses suggest that the ancestor of a recently discovered unicellular cyanobacterium was the plastid progenitor (Ponce-Toledo et al., Reference Ponce-Toledo, Deschamps, López-García, Zivanovic, Benzerara and Moreira2017) other studies suggest that this ancestral free-living, symbiotically competent cyanobacterium might have been a filamentous cyanobacterium—as these analyses recover affiliation to a section of extant filamentous cyanobacteria based on phylogenetic and sequence similarity data (Dagan et al., Reference Dagan, Roettger, Stucken, Landan, Koch, Major, Gould, Goremykin, Rippka, Tandeau de Marsac, Gugger, Lockhart, Allen, Brune, Maus, Pühler and Martin2013; Ochoa de Alda et al., Reference Ochoa de Alda, Esteban, Diago and Houmard2014). Up until today, this section of filamentous cyanobacteria brings about symbiotically competent strains that are the foundation of diverse cyanobacterial symbioses—from fungal to plant hosts—that we find on our planet (Rai et al., Reference Rai, Soderback and Bergman2000). In these interactions, the cyanobacteria provide fixed nitrogen to their hosts.

Cyanobacterial symbionts (cyanobionts) of plants are present throughout the plant kingdom, yet they have emerged multiple times independently (Bergman et al., Reference Bergman, Matveyev and Rasmussen1996; Rai et al., Reference Rai, Soderback and Bergman2000; Figure 1). As a result, the degree of intimacy and the solutions on how the symbionts are housed are as different as the plants that undergo these symbioses (Figure 1): cyanobacterial symbioses of bryophytes are found among representatives from all major lineages, that is, mosses, liverworts and hornworts (Adams et al., Reference Adams, Bergman, Nierzwicki-Bauer, Duggan, Rai and Schüßler2013; Adams & Duggan, Reference Adams and Duggan2008). Feathermosses form epiphytic associations with cyanobionts. These are not ‘accidental’ associations, but the epiphytes are recruited and they fix nitrogen for the mosses (Bay et al., Reference Bay, Nahar, Oubre, Whitehouse, Wardle, Zackrisson, Nilsson and Rasmussen2013; Stuart et al., Reference Stuart, Pederson, Weyman, Weber, Rassmussen and Dupont2020). In contrast, hornworts and some liverworts (e.g., Blasia) have special mucilage-filled cavities, auricles or canals (hornwort Leiosporoceros dussii) to host cyanobacteria (Adams et al., Reference Adams, Bergman, Nierzwicki-Bauer, Duggan, Rai and Schüßler2013; Adams & Duggan, Reference Adams and Duggan2008). Other liverworts have only epiphytic associations (Adams & Duggan, Reference Adams and Duggan2008); yet they are not as well studied as those from feathermosses (Bay et al., Reference Bay, Nahar, Oubre, Whitehouse, Wardle, Zackrisson, Nilsson and Rasmussen2013; Ininbergs et al., Reference Ininbergs, Bay, Rasmussen, Wardl and Nisson2011; Warshan et al., Reference Warshan, Espinoza, Stuart, Richter, Kim, Shapiro, Woyke, Kyrpides, Barry, Singan, Lindquist, Ansong, Purvine, Brewer, Weyman, Dupont and Rasmussen2017). Cycads (gymnosperms) induce the formation of a special organ type, the coralloid roots, that are colonised intercellularly by cyanobionts via broken tissue or lenticels (Bell-Doyon et al., Reference Bell-Doyon, Laroche, Saltonstall and Villarreal Aguilar2020; Rai et al., Reference Rai, Soderback and Bergman2000; Reinke, Reference Reinke1872; Suárez-Moo et al., Reference Suárez-Moo, Vovides, Griffith, Barona-Gómez and Cibrián-Jaramillo2019). The angiosperm Gunnera attracts its cyanobionts through mucilage-filled channels into their stems, where they enter and become intracellular (Khamar et al., Reference Khamar, Breathwaite, Prasse, Fraley, Secor, Chibane, Elhai and Chiu2010; Nilsson et al., Reference Nilsson, Rasmussen and Bergman2006; Towata, Reference Towata1985).

Figure 1. Cladogram of symbiotic interactions across land plants and diatoms. To the left, a simplified cladogram of the phylogenetic relationship among streptophytes (based on Puttick et al., Reference Puttick, Morris and Williams2018) is shown; diatoms (algae with a plastid of red algal origin) are very distantly related to streptophytes—the most recent common ancestor they share might be the last common ancestor of all eukaryotes (LECA). Among all major lineages of land plants except lycophytes (bold font), a few species that engage in symbiosis with cyanobacteria can be found (written in regular and italic font). To the right, filled dots denote key traits of the cyanobionts that dwell within those plants (and diatoms).

All aforementioned symbioses are facultative. In contrast to them stands the fern genus Azolla with their cyanobionts. Here, cyanobionts are transferred from one generation to the next (de Vries & de Vries, Reference de Vries and de Vries2018; Perkins & Peters, Reference Perkins and Peters1993; Zheng et al., Reference Zheng, Bergman, Chen, Zheng, Xiang and Rasmussen2009a). It is the only known permanent and obligate cyanobacterial symbiosis in the plant kingdom to date—that is apart from the chloroplast. Yet, in contrast to the chloroplast, the cyanobionts of Azolla are not endophytes, but live in special mucilage-filled cavities located in the dorsal leaf lobes (Figure 2).

Figure 2. The AzollaTrichormus symbiosis. At the top, the sporophyte body of Azolla filiculoides whose leaf cavities are packed with cyanobionts (red autofluorescence) is shown. At the bottom, a fluorescence micrograph of the filamentous cyanobionts (Trichormus azollae) within one leaf cavity are shown (red autofluorescence); note the heterocysts, which are larger than the vegetative cells.

Cyanobacterial symbioses do not occur in well-established model systems, although some of their hosts are currently emerging as such (Frangedakis et al., Reference Frangedakis, Shimamura, Villarreal, Li, Tomaselli, Waller, Sakakibara, Renzaglia and Szövényi2021; Li et al., Reference Li, Brouwer, Carretero-Paulet, Cheng, de Vries, Delaux, Eily, Koppers, Kuo, Li, Simenc, Small, Wafula, Angarita, Barker, Bräutigam, dePamphilis, Gould, Hosmani and Pryer2018). That said, the availability, the accessibility and the recent improvements to large-scale data have given the exploration of cyanobacterial symbioses a new boost (Li et al., Reference Li, Brouwer, Carretero-Paulet, Cheng, de Vries, Delaux, Eily, Koppers, Kuo, Li, Simenc, Small, Wafula, Angarita, Barker, Bräutigam, dePamphilis, Gould, Hosmani and Pryer2018; Reference Li, Nishiyama, Waller, Frangedakis, Keller, Li, Fernandez-Pozo, Barker, Bennett, Blázquez, Cheng, Cuming, de Vries, de Vries, Delaux, Diop, Harrison, Hauser, Hernández-García and Szövényi2020). Genomics, transcriptomics, proteomics and metagenomics provide new insights into the evolutionary biology of cyanobacterial symbioses. They establish the foundation to explore common recurrent patterns and identify the unique solutions that the evolutionary distant host plants have evolved to establish nitrogen-fixing symbioses with cyanobacteria. Comparative genomics on the cyanobacterial partners have identified patterns important in the biology of the bacteria, and those relevant for successful contributions to and gains from the symbiotic relationships. Here, we review the advances that the multiple different omics approaches have contributed to the understanding of cyanobacterial symbioses.

2. Recurrent molecular patterns in cyanobacterial symbioses: the cyanobionts

The scattered occurrences and diverse solutions in the establishment of the different plant-cyanobiont interactions may appear overwhelming. Yet, the questions that guide research on these symbioses are the same regardless of the symbiotic system. One of the key questions in plant-cyanobiont symbioses is what factors contribute to the initiation of the interaction. Here, much attention has been paid to the cyanobionts. Comparative genomics between symbiotically competent and not-competent cyanobacteria, as well as historically fingerprinting and nowadays metagenomic analyses, have been used to identify symbiotic communities and their features. Here, studies identified many recurring patterns.

2.1 Metagenomics, fingerprinting and cyanobacterial diversity

The question which cyanobacteria colonise the different hosts has led to several fingerprinting studies and isolation and Sanger sequencing of the cyanobacteria (e.g., Costa et al., Reference Costa, Paulsrud and Lindblad1999; Ininbergs et al., Reference Ininbergs, Bay, Rasmussen, Wardl and Nisson2011; Nilsson et al., Reference Nilsson, Bergman and Rasmussen2000; Zheng et al., Reference Zheng, Nilsson, Bergman and Rasmussen1999). Independent of the host, section IV/V cyanobacteria, foremost those of the order Nostocales, are consistently found in cyanobacterial symbioses (e.g., Liaimer et al., Reference Liaimer, Jensen and Dittmann2016; Nelson et al., Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019; Pratte & Thiel, Reference Pratte and Thiel2021; Ran et al., Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Robert Haselkorn and Bergman2010; Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). This suggests some inert symbiotic property that comes with the cyanobacteria of these sections. Specifically, most of the cyanobionts have been identified either as Nostoc or Anabaena (Adams et al., Reference Adams, Bergman, Nierzwicki-Bauer, Duggan, Rai and Schüßler2013; Rai et al., Reference Rai, Soderback and Bergman2000). The exception to the Nostoc/Anabaena symbionts is the cyanobiont of Azolla, which is as of late been called Trichormus azollae (see also discussions in Pereira & Vasconcelos, Reference Pereira and Vasconcelos2014). These symbiotic Nostocales are all filamentous, nitrogen-fixing and heterocyst-forming cyanobacteria. Some of them are able to colonise and associate in symbioses with various distinct hosts (Adams et al., Reference Adams, Bergman, Nierzwicki-Bauer, Duggan, Rai and Schüßler2013), suggesting that there may be limited host specificity.

Recent metagenomic data from hornworts uncovered that cyanobiont communities of the same host species from distinct locations can be quite different (Bouchard et al., Reference Bouchard, Peñaloza-Bojacá, Toupin, Guadalupe, Gudiño, Salazar Allen, Li and Villarreal Aguilar2020; Nelson et al., Reference Nelson, Hauser and Li2021). Location appears to play a major role in which (cyano)bacterial community is established in the symbiotic space. Even in the hornwort L. dussii, which has specialised canals that run through the entire lobe, geographic location of the host is the major determinant of the cyanobiont community, suggesting that the ability to colonise those canals is not uncommon for symbiotically capable cyanobacteria co-occurring with L. dussii (Bouchard et al., Reference Bouchard, Peñaloza-Bojacá, Toupin, Guadalupe, Gudiño, Salazar Allen, Li and Villarreal Aguilar2020). What cyanobiont will colonise the symbiotic space of hornworts is likely determined by the community present in the soil (Nelson et al., Reference Nelson, Hauser and Li2021). Despite that, the soil harbours certain species that never colonise the host, suggesting that while soil type dictates what species can generally become cyanobionts, the host has some measure of selectivity to allow only certain species entry into its symbiotic space (Nelson et al., Reference Nelson, Hauser and Li2021). This appears to be a common theme for the microbiome in tissues with close contact to soil. Other metabarcoding studies have highlighted that soil type is the major determined of the bacterial and eukaryotic root microbiota of plants (Lundberg et al., Reference Lundberg, Lebeis, Paredes, Yourstone, Gehring, Malfatti, Tremblay, Engelbrektson, Kunin, del Rio, Edgar, Eickhorst, Ley, Hugenholtz, Green Tringe and Dangl2012; Sapp et al., Reference Sapp, Ploch, Fiore-Donno, Bonkowski and Rose2018; Schlaeppi et al., Reference Schlaeppi, Dombrowski, Oter, Ver Loren van Themaat and Schulze-Lefert2014). In contrast to observations in cycads and hornworts, high host specificity and less influence by environmental factors were observed for cyanobacterial communities associated with feathermosses (Ininbergs et al., Reference Ininbergs, Bay, Rasmussen, Wardl and Nisson2011). Yet, in the lab, symbiotically competent Nostoc isolates from other hosts were generally capable to form a symbiotic association with these mosses (Warshan et al., Reference Warshan, Espinoza, Stuart, Richter, Kim, Shapiro, Woyke, Kyrpides, Barry, Singan, Lindquist, Ansong, Purvine, Brewer, Weyman, Dupont and Rasmussen2017). Another symbiosis that differs in terms of specificity is that of Azolla with T. azollae. This symbiosis is obligate and exists for more than 60 my (Carrapiço, Reference Carrapiço2006; Reference Carrapiço2010; Collinson, Reference Collinson2002). The obligatory nature is likely due to the symbiont’s highly reduced and eroded genome (Ran et al., Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Robert Haselkorn and Bergman2010). Concomitantly, genomic sequencing and fingerprinting studies highlighted that this cyanobiont is co-evolving with its host (Li et al., Reference Li, Brouwer, Carretero-Paulet, Cheng, de Vries, Delaux, Eily, Koppers, Kuo, Li, Simenc, Small, Wafula, Angarita, Barker, Bräutigam, dePamphilis, Gould, Hosmani and Pryer2018; Zheng et al., Reference Zheng, Nilsson, Bergman and Rasmussen1999).

2.2 Genomics of symbiotically competent cyanobacteria

Genomes of several symbiotic cyanobacteria have been sequenced. Most illuminating among these were the genomes of obligate symbionts. Data on T. azolla and the nitrogen-fixing cyanobionts of the diatoms Rhopalodia gibba, UCYN-A (a symbiont of a haptophyte) and Epithemia turgida all point towards genome erosion (Kneip et al., Reference Kneip, Voβ, Lockhart and Maier2008; Nakayama et al., Reference Nakayama, Kamikawa, Tanifuji, Kashiyama, Ohkouchi, Archibald and Inagaki2014; Ran et al., Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Robert Haselkorn and Bergman2010). All of these genomes are in the process of being streamlined while maintaining genes required for nitrogen fixation (N2-fixation). Epithemia turgida stands out by having completely eroded—or lost—genes for both photosystems.

In addition to the genome of T. azollae from Azolla filiculoides (Ran et al., Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Robert Haselkorn and Bergman2010), cyanobiont genomes from five other Azolla species have been generated (Dijkhuizen et al., Reference Dijkhuizen, Brouwer, Bolhuis, Reichart, Koppers, Huettel, Bolger, Li, Cheng, Liu, Wong, Pryer, Weber, Bräutigam and Schluepmann2018; Li et al., Reference Li, Brouwer, Carretero-Paulet, Cheng, de Vries, Delaux, Eily, Koppers, Kuo, Li, Simenc, Small, Wafula, Angarita, Barker, Bräutigam, dePamphilis, Gould, Hosmani and Pryer2018). The first genome of a cyanobacterial symbiont dwelling in plants was that of Nostoc punctiforme isolated from a cycad (Meeks et al., Reference Meeks, Elhai, Thiel, Potts, Larimer, Lamerdin, Predki and Atlas2001). In contrast to that of T. azollae (5.49 Mb), its genome is 9.06 Mb and considered large for cyanobacteria (Meeks et al. Reference Meeks, Elhai, Thiel, Potts, Larimer, Lamerdin, Predki and Atlas2001; Ran et al., Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Robert Haselkorn and Bergman2010). After that genome sequences from nine isolates from feathermosses, Blasia pusilla and one nonsymbiotic Nostoc strain CALU 996 have been reported (Warshan et al., Reference Warshan, Espinoza, Stuart, Richter, Kim, Shapiro, Woyke, Kyrpides, Barry, Singan, Lindquist, Ansong, Purvine, Brewer, Weyman, Dupont and Rasmussen2017; Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018), followed by 4 fully assembled genomes of isolates from Blasia and hornworts (Nelson et al., Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019), which are in the range of approximately 7 Mb in size, 6 isolates from a cycad (Gutiérrez-García et al., Reference Gutiérrez-García, Bustos-Díaz, Corona-Gómez, Ramos-Aboites, Sélem-Mojica, Cruz-Morales, Pérez-Farrera, Barona-Gómez and Cibrián-Jaramillo2019) and 10 isolates of facultative symbionts from Azolla (Pratte & Thiel, Reference Pratte and Thiel2021). For a comparison of genome sizes of cyanobacterial symbionts, see Figure 4a. In the next section, we will compare the findings gleaned from the different cyanobacterial genomes.

2.3 Comparative genomics: cyanobionts versus nonsymbiotic cyanobacteria

What is special about certain cyanobacteria from the order Nostocales that they are recurrently associated in symbioses? Phylogenetic analysis showed that all analysed cyanobacteria that occur in plant symbioses are included in a larger monophyletic clade that, however, also contains nonsymbiotic cyanobacteria (Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). While Nostoc isolates that can associate intracellularly with Gunnera do form a monophyletic clade, extracellular and epiphytic isolates are polyphyletic within the larger clade of symbionts (Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). A study by Nelson et al. (Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019), which sequenced genomes of additional isolates from bryophytes, corroborated this result. This further agrees with a phylogenomic survey across 650 diverse marine, freshwater and terrestrial cyanobacteria, that concluded that cyanobacteria with similar lifestyles tend to be more closely related to each other (Chen et al., Reference Chen, Teng, Zhao, Hu, Zhou, Han, Song and Shu2021).

2.3.1 Genomics of N2-fixation

The major contribution of the cyanobiont to the symbiosis with their plant hosts is fixed nitrogen. Filamentous cyanobacteria form specialised cells, called heterocysts, that have thicker cell walls, are photosynthetically inactive and create a microoxic environment inside the cell (Kumar et al., Reference Kumar, Mella-Herrera and Golden2010; Figure 3). In this oxygen-poor environment, high levels of oxygen-sensible nitrogenase occur (see Gallon, Reference Gallon1981). In plant–cyanobacterial symbioses, heterocyst formation and the rate of N2-fixation are increased; by how much depends on the respective host (Grilli Caiola et al., Reference Grilli Caiola, Canini and Moscone1989; Hill, Reference Hill1975; Wang et al., Reference Wang, Ekman and Bergman2004).

Figure 3. Syntrophic currencies exchanged by plant hosts and cyanobionts. A schematic Nostocales filament, consisting of vegetative cells and a heterocyst, is shown. The plant provides sugars (pink font), and the cyanobionts fix atmospheric nitrogen and provide it to the host via export (orange font). For an interesting discussion on putative transporters, see Roy et al. (Reference Roy, Reinders, Ward and McDonald2020).

N2-fixing cyanobacteria encode two types of nitrogenase. One nitrogenase requires molybdenum (Mo) and iron (Fe) as co-factors and is expressed only in heterocysts (Mo-nitrogenase; Thiel et al., Reference Thiel, Lyons and Erker1997). In the symbioses between T. azollae and A. filiculoides, it was hypothesised that Mo is provided and regulated by the host (Roy et al., Reference Roy, Reinders, Ward and McDonald2020). This is based on the observation that the genome of A. filiculoides encodes a Mo transporter (Li et al., Reference Li, Brouwer, Carretero-Paulet, Cheng, de Vries, Delaux, Eily, Koppers, Kuo, Li, Simenc, Small, Wafula, Angarita, Barker, Bräutigam, dePamphilis, Gould, Hosmani and Pryer2018), which, however, may likely be an importer for the plant (Roy et al., Reference Roy, Reinders, Ward and McDonald2020). Its upregulation under nitrogen deprivation hints to an involvement in symbiosis (Eily et al., Reference Eily, Pryer and Li2019). The Mo-nitrogenase is a complex that requires a range of enzymes for its synthesis and assembly (Dean et al., Reference Dean, Bolin and Zheng1993). The complex itself requires the tetraheteromeric dinitrogenase encoded by nifD and nifK and the homodimeric dinitrogenase reductase encoded by nifH. Several nitrogen-fixing cyanobacteria encode an additional vanadium-based nitrogenase, which also requires Fe (V-nitrogenase; Nelson et al., Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019; Pratte & Thiel, Reference Pratte and Thiel2021; Thiel & Pratte, Reference Thiel and Pratte2001). In Anabaena variabilis (now N. variabilis), transcription of the V-nitrogenase is repressed by both fixed nitrogen and the presence of Mo (Pratte et al., Reference Pratte, Sheridan, James and Thiel2013). In the absence of Mo, V-nitrogenase activity is induced (Thiel, Reference Thiel1996). The gene cluster encoding V-nitrogenase and its regulators (Pratte et al., Reference Pratte, Sheridan, James and Thiel2013), and the vanadate transporters (Pratte & Thiel, Reference Pratte and Thiel2006), is conserved across cyanobacteria (Nelson et al., Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019). The first two assembled genomes from plant cyanobionts (T. azollae and N. punctiforme PCC 73102), however, likely lack V-nitrogenase-encoding genes (Meeks et al., Reference Meeks, Elhai, Thiel, Potts, Larimer, Lamerdin, Predki and Atlas2001; Ran et al., Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Robert Haselkorn and Bergman2010). Yet, recent studies (Nelson et al., Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019; Pratte & Thiel, Reference Pratte and Thiel2021) report their presence in genomes of facultative, symbiotically competent cyanobacteria from Azolla, Blasia and hornworts. Nelson et al. (Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019) reported the gene cluster in two of the cyanobionts they studied located on a plasmid and suggested the option for a gain via horizontal gene transfer. If and when V-nitrogenase plays a role under Mo-limited conditions during cyanobacteria–plant symbioses needs to be further investigated. A recent study on V-nitrogenase activity in lichen-associated cyanobacteria showed that availability of Mo correlated with the rate of N2-fixation from V-nitrogenase (Darnajoux et al., Reference Darnajoux, Magain, Renaudin, Lutzoni, Bellenger and Zhang2019).

Most proteomic and transcriptomic approaches have focused on Mo-nitrogenase. Ekman et al. (Reference Ekman, Tollbäck and Bergman2008) provided a proteomic comparison between the cyanobiont of the floating fern A. filiculoides and the symbiotic Nostoc isolate N. punctiforme PCC 73102 grown in culture. These data show that in the symbiotic association of Azolla proteins related to nitrogen metabolism were enhanced while those related to photosynthesis were reduced in the cyanobiont compared to the free-living, symbiotically capable N. punctiforme isolate. Similar tendencies were observed in the proteomic comparison of N. punctiforme PCC 73102 in symbiosis with Gunnera manicata and in culture (Ekman et al., Reference Ekman, Tollbäck, Klint and Bergman2006). In contrast, the epiphytic associations appear different. In feathermosses, nitrogen is transferred from cyanobiont to host (Bay et al., Reference Bay, Nahar, Oubre, Whitehouse, Wardle, Zackrisson, Nilsson and Rasmussen2013), yet the expression of genes involved in N2-fixation or heterocyst formation is not induced in symbiotic versus free-living Nostoc (Warshan et al., Reference Warshan, Espinoza, Stuart, Richter, Kim, Shapiro, Woyke, Kyrpides, Barry, Singan, Lindquist, Ansong, Purvine, Brewer, Weyman, Dupont and Rasmussen2017). This suggests that the Nostoc isolates may have a similar rate of N2-fixation in their symbiotic and free-living state. In agreement with that, both symbiotically competent Nostoc isolates analysed in this study identified nifBHKU—that is, genes encoding structural components nifH and K, a factor involved in nitrogenase assembly (nifU; Dos Santos et al., Reference Dos Santos, Smith, Frazzon, Cash, Johnson and Dean2004; Johnson et al., Reference Johnson, Dos Santos and Dean2005) and an enzyme required for the FeMo co-factor assembly (nifB; Curatti et al., Reference Curatti, Ludden and Rubio2006)—and the regulator for heterocyst formation, hetR, in the top 10% of transcripts with the highest abundance in free-living and symbiotic condition (Warshan et al., Reference Warshan, Espinoza, Stuart, Richter, Kim, Shapiro, Woyke, Kyrpides, Barry, Singan, Lindquist, Ansong, Purvine, Brewer, Weyman, Dupont and Rasmussen2017). In summary, while in close association the symbiosis appears to lead to an induction of N2-fixation, the example of lose associations from the feathermosses suggests that here the host thrives already on the steady-state levels of fixed nitrogen provided by the cyanobionts. One could wonder whether this may be due to the specificity in symbiont recruitment described based on genomic fingerprinting of nifH sequences (Ininbergs et al., Reference Ininbergs, Bay, Rasmussen, Wardl and Nisson2011); yet the observations were likewise true for a moss-specific and the symbiotically competent N. punctiforme isolate (Warshan et al., Reference Warshan, Espinoza, Stuart, Richter, Kim, Shapiro, Woyke, Kyrpides, Barry, Singan, Lindquist, Ansong, Purvine, Brewer, Weyman, Dupont and Rasmussen2017).

What makes cyanobacteria symbiotically competent? And do they all use similar mechanisms for associating with their hosts? The latter appears likely given that most of them—with the exception of the obligate symbiont T. azollae–are able to colonise different hosts (e.g., Pratte & Thiel, Reference Pratte and Thiel2021; Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). As such the question may rather be ‘What are the molecular mechanisms of cyanobacterial symbioses?’. To answer these two questions, (comparative) genomic, transcriptomic and proteomic analyses have been carried out in different host-cyanobacteria associations.

2.3.2 Cyanobacterial symbioses from colonisation to communication during a functional symbiosis

The ability to infect plants and associate in symbiosis is found in several members of the genus Nostoc/Anabaena. The last common ancestor of the monophyletic intra/extracellular symbiotic clade may have already been primed for the ability to associate with diverse hosts (Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). Yet this particular clade includes not only symbiotically competent isolates but also a variety of isolates not found in symbiosis. Warshan et al. (Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018) provide a set of 170 gene families that are exclusively encoded in the genomes of 10 facultative Nostoc symbionts. These include, among others, genes coding for transporters involved in the import of organic sulphur compounds, phosphate and branched amino acids and ammonium (Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). Genomic data on cycad-derived symbionts found several metabolite biosynthesis pathways specific to the symbiotic cyanobacteria (Gutiérrez-García et al., Reference Gutiérrez-García, Bustos-Díaz, Corona-Gómez, Ramos-Aboites, Sélem-Mojica, Cruz-Morales, Pérez-Farrera, Barona-Gómez and Cibrián-Jaramillo2019). Pratte and Thiel (Reference Pratte and Thiel2021) confirmed the presence of the majority of the gene families identified by Warshan et al. (Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018) in the genomes of 10 Nostoc isolates that had previously been isolated from Azolla (Meeks et al., Reference Meeks1988; Zimmerman et al., Reference Zimmerman, Rosen and Lumpkin1989). These isolates had originally been obtained as the putative permanent and vertically transferred cyanobiont of Azolla, but were later identified as cyanobacteria that colonised the cavity in addition to T. azollae (Papaefthimiou et al., Reference Papaefthimiou, Van Hove, Lejeune, Rasmussen and Wilmotte2008). Infection assays with Blasia sp. confirmed that these isolates are capable of associating in symbioses and suggest a wider host range than only Azolla (Pratte & Thiel, Reference Pratte and Thiel2021). However, genomic analyses of these symbiotic isolates and strains previously described as free-living identified several ‘symbioses genes’ in the free-living cyanobacteria. This may have two reasons: (a) Warshan et al. (Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018) used one particular cyanobacterial strain as nonsymbiotic control, while Pratte and Thiel used a diverse set; and (b) symbiotic compatibility may be present even in free-living isolates that currently have not been found in metagenomic studies of symbiotic communities.

As described above, the genome of the symbiont of Azolla is smaller than that of most other symbiotically competent cyanobacteria (Meeks et al., Reference Meeks, Elhai, Thiel, Potts, Larimer, Lamerdin, Predki and Atlas2001; Nelson et al., Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019; Ran et al., Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Robert Haselkorn and Bergman2010; Figure 4a). For one, T. azollae does not belong to the genus Nostoc and has a distinct phylogenetic placement (Nelson et al., Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019; Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). Moreover, the genome shows strong evidence of erosion (Ran et al., Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Robert Haselkorn and Bergman2010). Many genes have been lost from the genome of T. azollae or subjected to pseudogenisation (Ran et al., Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Robert Haselkorn and Bergman2010; Figure 4b,c). Accordingly, 28 gene families present in facultative symbionts have been lost in T. azollae (Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). Among those were genes involved in chemotaxis and in transport and metabolism of alkane and aliphatic sulfonate, underpinning the differences between obligate and facultative symbioses.

Figure 4. Genome characteristics of genomes from symbiotic and nonsymbiotic cyanobacteria and cyanobacterial host plants. (a) Genome size of several cyanobacteria shown to be symbiotic or free-living (blue, cyano) and all sequenced plant hosts for cyanobacteria (yellow, plants). (b) Number of annotated protein-coding genes in the genomes of cyanobacteria (blue, cyano) and cyanobacterial host plants (yellow, plants). (c) Number of pseudogenes described for the here included cyanobacterial genome data (for details, see description below). (d) Relative repeat content in the assembled plant genomes from plants symbiotically associating with cyanobacteria. To have a mixed set of cyanobacteria associating with plants and free-living, we included cyanobacteria previously sequenced and/or included in the comparative analyses by Pratte & Thiel (Reference Pratte and Thiel2021) and Nelson et al. (Reference Nelson, Hauser, Gudiño, Guadalupe, Meeks, Salazar Allen, Villarreal and Li2019). NCBI accessions have been derived from these publications, and actual data have been obtained from the National Center for Biotechnology Information (NCBI). For cyanobacterial hosts, all host species that have been sequenced (Anthoceros agrestis, Anthoceros punctatus, Anthoceros angustus, Azolla filiculoides, Sphagnum fallax and Pleurozium schreberi) are included in the figure. The data are derived from Li et al. (Reference Li, Brouwer, Carretero-Paulet, Cheng, de Vries, Delaux, Eily, Koppers, Kuo, Li, Simenc, Small, Wafula, Angarita, Barker, Bräutigam, dePamphilis, Gould, Hosmani and Pryer2018, Reference Li, Nishiyama, Waller, Frangedakis, Keller, Li, Fernandez-Pozo, Barker, Bennett, Blázquez, Cheng, Cuming, de Vries, de Vries, Delaux, Diop, Harrison, Hauser, Hernández-García and Szövényi2020), Pederson et al. (Reference Pederson, Warshan and Rasmussen2019), Zhang et al. (Reference Zhang, Fu, Li, Zhao, Liu, Li, Zwaenepoel, Ma, Goffinet, Guan, Xue, Liao, Wang, Wang, Wang, Zhang, Wang, Jia, Wang and Chen2020), and Genome statistic information from Phytozome genome ID: 522. Data on repeat content were not reported for S. fallax, and the number of pseudogenes has not been reported for cyanobacteria isolated from feathermosses.

2.3.2.1 Insights into motility and chemotaxis

What all cyanobionts have in common is their motility and ability to move towards their hosts. This is likewise true for facultative and obligate cyanobionts. In the case of Azolla, the filamentous motile stage—hormogonia—is induced to infect newly developed leaf cavities and sporocarps of the fern (Zheng et al., Reference Zheng, Rang and Bergman2009b). Hosts can induce the formation of hormogonia—as is discussed later in the section on the hosts. Yet, the ability to move is not all; extracts of diverse hosts are able to attract cyanobacteria chemotactically (Nilsson et al., Reference Nilsson, Rasmussen and Bergman2006). Genes involved in chemotaxis of hormogonia (cheA, B, D, R and W) are present in the genome of the symbiotically competent N. punctiforme (Meeks et al., Reference Meeks, Elhai, Thiel, Potts, Larimer, Lamerdin, Predki and Atlas2001). The absence of functional CheR methyltransferase results in loss of motility and reduced colonisation of N. punctiforme of Blasia (Duggan et al., Reference Duggan, Thiel and Adams2013). Furthermore, symbiotically competent strains induce chemotaxis-associated genes upon chemical contact with a putative host (Warshan et al., Reference Warshan, Espinoza, Stuart, Richter, Kim, Shapiro, Woyke, Kyrpides, Barry, Singan, Lindquist, Ansong, Purvine, Brewer, Weyman, Dupont and Rasmussen2017). Accordingly, genes relevant for chemotaxis are among the 170 symbioses genes recovered by Warshan et al. (Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). In contrast, among the genes lost from the genome of T. azollae are some families involved in chemotaxis (Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018). That is interesting because one would expect some host-derived guidance to the sporocarps or a new leaf cavity. But in contrast to facultative symbioses, movement of T. azollae may not only be guided by chemical communication. Trichomes are suggested to act as physical guidance towards new leaf cavities and sporocarps (Calvert et al., Reference Calvert, Pence and Peters1985; Calvert & Peters, Reference Calvert and Peters1981; Hill, Reference Hill1989). Nonetheless, metabolic-based communication may be altered during the initiation of sporocarp formation. The transcriptional profile of transporter-encoding genes from T. azollae changed significantly during the induction of sporocarps (Dijkhuizen et al., Reference Dijkhuizen, Tabatabaei, Brouwer, Rijken, Buijs, Güngör and Schluepmann2021). Already early on in research on Azolla, it was suspected that host-derived phenolics and flavonoids in trichomes may be involved in the movement of T. azollae (Carrapiço & Tavares, Reference Carrapiço and Tavares1989; Ishikura, Reference Ishikura1982; Pereira & Carrapiço, Reference Pereira and Carrapiço2007). Once the hosts—or new host tissue in case of Azolla—are colonised, hormogonia revert back to vegetative filaments (Bay et al., Reference Bay, Nahar, Oubre, Whitehouse, Wardle, Zackrisson, Nilsson and Rasmussen2013; Meeks & Elhai, Reference Meeks and Elhai2002; Zheng, Bergman, et al., Reference Zheng, Bergman, Chen, Zheng, Xiang and Rasmussen2009a). This repression is guided by the host and its regulation is governed by proteins encoded by the hrm-locus (Adams & Duggan, Reference Adams and Duggan2012 ; Campbell et al., Reference Campbell, Wong and Meeks2003; Cohen & Meeks, Reference Cohen and Meeks1997; Meeks, Reference Meeks1998).

The broad host range of many cyanobacterial symbionts that is occasionally observed in nature and used in the lab suggests that either (a) the symbioses rely on a common mechanism despite the diverse partners, (b) symbiotically competent Nostoc evolved the ability to recognise and respond a vast variety of different signals or (c) both.

2.3.2.2 Insights into nutrient exchange and transport

An important aspect of cyanobacterial biology during symbioses is its transporter system. In symbioses, photosynthesis is reduced in cyanobionts as a trade-off for the higher rates of N2-fixation (Bergman et al., Reference Bergman, Rai, Johansson and Söderbäck1992; Ekman et al., Reference Ekman, Tollbäck and Bergman2008; Peters & Meeks, Reference Peters and Meeks1989; Rai et al., Reference Rai, Borthakur, Singh and Bergman1989; Reference Rai, Soderback and Bergman2000). Thus, cyanobionts need to be provided with carbon from their hosts (Peters & Meeks, Reference Peters and Meeks1989; Rai et al., Reference Rai, Soderback and Bergman2000; Söderbäck & Bergman, Reference Söderbäck and Bergman1993; Stewart & Rodgers, Reference Stewart and Rodgers1977; Figure 3). Indeed, proteomics of T. azollae indicated an increase in a transporter associated—by sequence homology—with a phosphotransferase system fructose-specific IIC component (Ekman et al., Reference Ekman, Tollbäck and Bergman2008). In agreement, both in symbiosis with Gunnera and Azolla, key enzymes of the oxidative pentose phosphate pathway are induced (Ekman et al., Reference Ekman, Tollbäck, Klint and Bergman2006; Reference Ekman, Tollbäck and Bergman2008). This may suggest that fructose is the sugar transported between hosts and cyanobionts. However, a mutant study on sugar transporters of symbiotically competent N. punctiforme showed that only mutants where glucose permease function was impaired were unable to infect the hornwort Anthoceros punctatus (Ekman et al., Reference Ekman, Picossi, Champbell, Meeks and Flores2013). This leaves four options: (a) cyanobionts do not receive the same type of sugar from every host system, (b) protein sequences that were by homology associated with fructose transport and metabolism may need to be reinvestigated with new additional data at hand, (c) a lack of fructose during an established interaction may be circumvented, while glucose is required as a signal for the establishment of infection or (d) glucose is the transported sugar used as a carbon source for cyanobionts.

Warshan et al. (Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018) found that the genomes of facultative symbiotic cyanobacteria are enriched in transporters for phosphate and organic sulphur and the metabolism of the latter. Both phosphate and sulphate transporters are induced upon contact between the feathermoss Pleurozium schreberi and symbiotic Nostoc isolates (Warshan et al., Reference Warshan, Espinoza, Stuart, Richter, Kim, Shapiro, Woyke, Kyrpides, Barry, Singan, Lindquist, Ansong, Purvine, Brewer, Weyman, Dupont and Rasmussen2017). These data point to organic sulphur and phosphate as important nutrients delivered by the host. And indeed, sulphur compounds are a currency in feathermoss–cyanobiont symbioses (Stuart et al., Reference Stuart, Pederson, Weyman, Weber, Rassmussen and Dupont2020). Furthermore, elimination of a functional alkane sulfonate monooxygenase by targeted mutagenesis of symbiotic N. punctiforme showed that sulphur metabolism is required for establishment of a symbiosis with feathermosses (Stuart et al., Reference Stuart, Pederson, Weyman, Weber, Rassmussen and Dupont2020).

2.3.2.3 Insights into cyanobacterial stress during symbiosis

Nitrogen starvation induces a stress response in cyanobacteria (Christman et al., Reference Christman, Campbell and Meeks2011). It is thus not surprising that early proteomic analyses on the cyanobionts of Azolla and Gunnera identified an induction of proteins involved in stress (Ekman et al., Reference Ekman, Tollbäck, Klint and Bergman2006; Reference Ekman, Tollbäck and Bergman2008). In both host systems, these included proteins in oxidative stress. This agrees with data on cyanobacteria epiphytically associating with feathermosses. While both competent as well as non-competent Nostoc isolates highly expressed proteins involved in oxidative stress responses, some of these were only induced in the competent isolates upon physical contact with the mosses—both in the secreted proteome and transcriptome (Warshan et al., Reference Warshan, Espinoza, Stuart, Richter, Kim, Shapiro, Woyke, Kyrpides, Barry, Singan, Lindquist, Ansong, Purvine, Brewer, Weyman, Dupont and Rasmussen2017). Moreover, these same proteins were reduced in the secreted protein fraction of the non-competent Nostoc isolate in this study. Additionally, in a metabolic screen, some Nostoc isolates, including some cyanobionts from Blasia, have been described to produce the cyanobacterial toxin microcystin (Liaimer et al., Reference Liaimer, Jensen and Dittmann2016). A microcystin deficient mutant showed higher sensitivity to reactive oxygen species (ROS; Zilliges et al., Reference Zilliges, Kehr, Meissner, Ishida, Mikkat, Hagemann, Kaplan, Börner and Dittmann2011), suggesting a protective role of these compounds against oxidative stress. Indeed, the microcystins were not secreted by symbiotic Nostoc isolates supporting a role inside the symbiont (Liaimer et al., Reference Liaimer, Jensen and Dittmann2016). Overall, these data suggest that a vital part of establishing and possibly maintaining N2-fixing symbioses in many host-cyanobiont interactions relies on the induction of protection against ROS. This notion is corroborated by insights into a different nitrogen-fixing symbiosis: that occurring between rhizobial bacteria and legumes. A study by Günther et al. (2007) on Lotus japonicus has shown that high respiration rates mediated by leghemoglobins (which keep free O2 levels low while acting in internal O2 trafficking; see Appleby, Reference Appleby1984) lead to elevated levels of ROS.

2.4 Metagenomics and fingerprinting point to a bacterial microcosm surrounding the cyanobionts

No organism exists alone in its natural habitat. Already decades ago, the co-occurrence of other bacteria, next to cyanobionts, was noticed (Carrapiço, Reference Carrapiço1991; Wallace & Gates, Reference Wallace and Gates1986). Metagenome sequencing has uncovered a diversity of non-cyanobacterial symbionts (e.g., Dijkhuizen et al., Reference Dijkhuizen, Brouwer, Bolhuis, Reichart, Koppers, Huettel, Bolger, Li, Cheng, Liu, Wong, Pryer, Weber, Bräutigam and Schluepmann2018; Nelson et al., Reference Nelson, Hauser and Li2021). That cyanobionts are surrounded by various bacteria was noted for many systems. Thus, the host and cyanobiont do not only interact with each other, but with an entire microbial community. One might find this unsurprising. Bacteria and archaea are omnipresent, and a diversity of them associates with roots and leaves of plants. What roles they play and how essential the bacterial composition is to their plant hosts are, in general, the topics of current research (Fitzpatrick et al., Reference Fitzpatrick, Salas-González, Conway, Finkel, Gilbert, Russ, Teixeira and Dangl2020; Trivedi et al., Reference Trivedi, Leach, Tringe, Sa and Singh2020). The question though is, how are they capable to occupy the symbiotic space from the rather ‘open’ versions such as cavities of hornworts, liverworts and Azolla to more ‘closed’ options such as coralloid roots of cycads.

Are these bacteria capable of hijacking the symbiotic crosstalk between host and cyanobiont or do symbiotic spaces fulfil a general role in attracting symbionts for the plant? In a metagenomic study on hornwort cavities, it was observed that the nonsymbiotic associates follow the same patterns of specificity as the cyanobionts (Nelson et al., Reference Nelson, Hauser and Li2021). The original soil community dictates what organisms are present and as such have a chance to be recruited; here too, the diversity of the cavity does not equal that of the soil per se. Hence, Anthoceros has some measure of control, which microbes it allows in and which it does not. In Azolla species, there are recurring members of the microbial community that are attuned to the developmental status of the symbiotic organs and the cyanobacterial physiology (Dijkhuizen et al., Reference Dijkhuizen, Brouwer, Bolhuis, Reichart, Koppers, Huettel, Bolger, Li, Cheng, Liu, Wong, Pryer, Weber, Bräutigam and Schluepmann2018). These bacteria were found to be co-transferred during vegetative and sexual transmission of the cyanobacterial population (Carrapiço, Reference Carrapiço1991). In contrast to the other symbioses, in Gunnera petaloidea and Gunnera chilensis, fungi and bacteria were observed within both the gland cells and the mucilage (Johansson & Bergman, Reference Johansson and Bergman1992; Towata, Reference Towata1985), but there are no records on the presence of such microbes in the exact same symbiotic spaces as the cyanobionts. Yet, these are only microscopic observations, and no molecular insights have been garnered.

These observations suggest a complex interaction between host, cyanobacteria and other microorganisms. The nature of these interactions is to a large degree not yet fully understood. With the omics approaches at hand, the nature of these interactions can now be further investigated. Dijkhuizen et al. (Reference Dijkhuizen, Brouwer, Bolhuis, Reichart, Koppers, Huettel, Bolger, Li, Cheng, Liu, Wong, Pryer, Weber, Bräutigam and Schluepmann2018) performed the first comprehensive survey of the microbiome that occurs in six different Azolla species. The authors found that there are recurring Rhizobiales in the Azolla leaf pockets. Interestingly, similar presence of Rhizobiales has been reported in the endospheres of the hornwort Leiosporoceros (Bouchard et al., Reference Bouchard, Peñaloza-Bojacá, Toupin, Guadalupe, Gudiño, Salazar Allen, Li and Villarreal Aguilar2020), and the cycads Dioon and Zamia (Bell-Doyon et al., Reference Bell-Doyon, Laroche, Saltonstall and Villarreal Aguilar2020; Suárez-Moo et al., Reference Suárez-Moo, Vovides, Griffith, Barona-Gómez and Cibrián-Jaramillo2019). While these recent microbiome studies suggest that Rhizobiales might have certain significances in plant–cyanobacteria symbiosis, their prevalence and functional roles remain to be tested. Additionally, Dijkhuizen et al. (Reference Dijkhuizen, Brouwer, Bolhuis, Reichart, Koppers, Huettel, Bolger, Li, Cheng, Liu, Wong, Pryer, Weber, Bräutigam and Schluepmann2018) identified putative denitrifiers in the leave pocket of Azolla, suggesting that cheaters might exist in the symbiotic spaces. Yet, with the exception of the analyses of Azolla, no functional predictions have been made for the microbial community in cyanobacterial symbioses.

Currently, no comparative metagenomic study across the cyanobacterial hosts exists. Additionally, most metagenome sequencing studies focus on marker genes instead of entire genomes; this hampers inferences on functional roles in the symbiotic community and the identification of non-cyanobacterial symbionts or cheaters. Thus, how similar the microbiome in the symbiotic space of the phylogenetically distinct hosts is compared to nonsymbiotic tissues can only be approximated. Yet, because in all cyanobacterial hosts, facultative cyanobionts have some similar, symbiosis-associated genomic content and they perform similar functions within the symbiosis (Pratte & Thiel, Reference Pratte and Thiel2021, Warshan et al., Reference Warshan, Liaimer, Pederson, Kim, Shapiro, Woyke, Altermark, Pawlowski, Weyman, Dupont and Rasmussen2018), they might create a similar milieu in different symbiotic spaces. One could, hence, expect that the functional composition of the microbiome, be it beneficial or cheating, might be similar. Indeed, with more metagenomic data accumulating, it is becoming clear that while taxonomic compositions vary strongly, there are recurrent patterns in the functional composition of a microbiome in similar environments (Burke et al., Reference Burke, Steinberg, Rusch, Kjelleberg and Thomas2011). After all, it is the biochemical properties of an environment on which selection acts (see excellent discussion in Doolittle & Inkpen, Reference Doolittle and Inkpen2018). Future studies will illuminate convergence of functions in the symbiotic microverse across these diverse interactions.

3. Recurrent molecular patterns in cyanobacterial symbioses—a host-oriented perspective

The sequencing of cyanobacteria associated with nitrogen-fixing symbioses is for a long time on-going, and as a consequence, many of the strains have been sequenced. For practical reasons regarding the much larger genomes (Figure 4a), from cost to computational power and the availability of methods to cope with a high repetitive content (Figure 4d), the sequencing of the host genomes has progressed more slowly. However, in the last few years, the genomes of several host plants have been sequenced: Azolla filiculoides (Li et al., Reference Li, Brouwer, Carretero-Paulet, Cheng, de Vries, Delaux, Eily, Koppers, Kuo, Li, Simenc, Small, Wafula, Angarita, Barker, Bräutigam, dePamphilis, Gould, Hosmani and Pryer2018), four hornwort genomes [two Anthoceros agrestis ecotypes and A. punctatus (Li et al., Reference Li, Nishiyama, Waller, Frangedakis, Keller, Li, Fernandez-Pozo, Barker, Bennett, Blázquez, Cheng, Cuming, de Vries, de Vries, Delaux, Diop, Harrison, Hauser, Hernández-García and Szövényi2020) and Anthoceros angustus (Zhang et al., Reference Zhang, Fu, Li, Zhao, Liu, Li, Zwaenepoel, Ma, Goffinet, Guan, Xue, Liao, Wang, Wang, Wang, Zhang, Wang, Jia, Wang and Chen2020)] and the feathermoss P. schreberi (Pederson et al., Reference Pederson, Warshan and Rasmussen2019). Additionally, the 1KP project (Matasci et al., Reference Matasci, Hung, Yan, Carpenter, Wickett, Mirarab, Nguyen, Warnow, Ayyampalayam, Barker, Burleigh, Gitzendanner, Wafula, Der, dePamphilis, Roure, Philippe, Ruhfel, Miles and Wong2014) covers transcriptomes of G. manicata, 4 cycads (Cycas micholoitzii, Stangeria eriopus, Dioon edule and Encephalartos barteri), 10 different hornwort species from diverse genera, the liverwort Blasia sp. and an additional Azolla species. Additional transcriptomes for A. filiculoides, two other Azolla species, Azolla pinnata and Azolla caroliniana, and the hornwort A. punctatus exist (Brouwer et al., Reference Brouwer, Bräutigam, Buijs, Tazelaar, van der Werf, Schlüter, Reichart, Bolger, Usadel, Weber and Schluepmann2017; Chatterjee et al., Reference Chatterjee, Schafran, Li and Meeks2021; de Vries et al., Reference de Vries, Fischer, Roettger, Rommel, Schluepmann, Bräutigam, Carlsbecker and Gould2016; Li et al., Reference Li, Nishiyama, Waller, Frangedakis, Keller, Li, Fernandez-Pozo, Barker, Bennett, Blázquez, Cheng, Cuming, de Vries, de Vries, Delaux, Diop, Harrison, Hauser, Hernández-García and Szövényi2020; Qi et al., Reference Qi, Kuo, Guo, Li, Li, Qi, Wang, Hu, Xiang, Zhang, Guo, Huang and Ma2018).

3.1. Establishment of interaction with cyanobacterial and other symbionts

Some of the now existing data were already available and included in a large comparative analysis that targeted the distribution and evolution of the symbiotic toolkit associated with arbuscular mycorrhizal (AM) symbioses (Radhakrishnan et al., Reference Radhakrishnan, Keller, Rich, Vernié, Mbadinga Mbadinga, Vigneron, Cottret, San Clemente, Libourel, Cheema, Linde, Eklund, Shifeng Cheng, Wong, Lagercrantz, Li, Oldroyd and Delaux2020). The authors found that symbiotic genes involved in these particular symbioses are not present in land plants undergoing cyanobacterial but no AM symbioses. This suggests that even though the symbiotic pathways associated with the establishment of AM symbioses have been recruited for nitrogen-fixing symbioses with rhizobia (Parniske, Reference Parniske2008), cyanobacterial symbioses rely on different pathways for their interaction. Several data from Gunnera, cycads, hornworts, Blasia and feathermosses show that specific attractants exist (Bay et al., Reference Bay, Nahar, Oubre, Whitehouse, Wardle, Zackrisson, Nilsson and Rasmussen2013; Campbell & Meeks, Reference Campbell and Meeks1989; Khamar et al., Reference Khamar, Breathwaite, Prasse, Fraley, Secor, Chibane, Elhai and Chiu2010; Nilsson et al., Reference Nilsson, Rasmussen and Bergman2006; Rasmussen et al., Reference Rasmussen, Johansson and Bergman1994), and that the presence of the hosts induces hormogonia formation (Bay et al., Reference Bay, Nahar, Oubre, Whitehouse, Wardle, Zackrisson, Nilsson and Rasmussen2013; Campbell & Meeks, Reference Campbell and Meeks1989; Meeks & Elhai, Reference Meeks and Elhai2002; Ow et al., Reference Ow, Gantar and Elhai1999; Rasmussen et al., Reference Rasmussen, Johansson and Bergman1994; Figure 1). Even in the interaction between Azolla and its cyanobiont, hormogonia formation and specific attraction to the sporocarps as well as the apex of the shoot (where new leaves are formed) is observed (Calvert et al., Reference Calvert, Pence and Peters1985; Hill, Reference Hill1989; Perkins & Peters, Reference Perkins and Peters1993; Peters et al., Reference Peters, Toia, Raveed and Levine1978; Peters & Meeks, Reference Peters and Meeks1989; Zheng, Bergman, et al., Reference Zheng, Bergman, Chen, Zheng, Xiang and Rasmussen2009a).

Campbell and Meeks (Reference Campbell and Meeks1989) found a hormogonia inducing factor (HIF) of a size between 12 and 14 kDa in the hornwort A. punctatus. They did not further identify the nature of the secreted signal but suggested a poly-phenol-like molecule (Campbell & Meeks, Reference Campbell and Meeks1989; Meeks & Elhai, Reference Meeks and Elhai2002). Rasmussen et al. (Reference Rasmussen, Johansson and Bergman1994) identified that mucilage of different Gunnera species induced hormogonia. Here, too, a molecule of about 12 kDa was found. In this case, proteinase K treatment suggested it to be a peptide. The accumulating transcriptomic and genomic data can now be helpful in identifying short peptides that may act as HIFs and test their function in the lab. Another HIF, a mixture of diacylglycerols, was recently identified from the cycad Cycas revoluta (Hashidoko et al., Reference Hashidoko, Nishizuka, Tanaka, Murata, Murai and Hashimoto2019). The major compound 1-palmitoyl-2-linoleoyl-sn-glycerol was the most active HIF. Additionally, sugars are in discussion as putative HIFs. Indeed, the secreted mucilage of mature glands of Gunnera and soluble sugars within it induced hormogonia formation (Khamar et al., Reference Khamar, Breathwaite, Prasse, Fraley, Secor, Chibane, Elhai and Chiu2010).

Because N2-fixation is inhibited in hormogonia, a reversion to filaments with heterocysts, the N2-fixing cells of cyanobacteria, is needed for the establishment of successful symbioses. This is triggered by all cyanobacterial symbioses (Meeks & Elhai, Reference Meeks and Elhai2002). Glucose and fructose, which are almost absent from the mucilage of Gunnera, but accumulate within the gland tissues, repressed the formation of hormogonia (Khamar et al., Reference Khamar, Breathwaite, Prasse, Fraley, Secor, Chibane, Elhai and Chiu2010). Sugars have also been highlighted as putative attractants for the cyanobionts (Nilsson et al., Reference Nilsson, Rasmussen and Bergman2006). Functional analyses of the genome of the symbiotic N. punctiforme showed that it encodes a functional fructose transporter, a glucose permease and an OprB family porin, which was associated with the uptake of both sugars (Ekman et al., Reference Ekman, Picossi, Champbell, Meeks and Flores2013). In other cyanobacteria, this porin family has recently been associated with transport of different metals (Qiu et al., Reference Qiu, Jiang, Lis, Li, Deng, Shang, Sun, Keren and Qiu2021; Schätzle et al., Reference Schätzle, Brouwer, Liebhart, Stevanovic and Schleiff2021). Additionally, some mutants of porin-like genes from Anabaena sp. PCC 7120 had altered phenotypes of the outer membrane, affected integration of other porins in the membrane or affected nitrogen demand and/or fixation (Schätzle et al., Reference Schätzle, Brouwer, Liebhart, Stevanovic and Schleiff2021). Hence, the OprB family appears to have a diversity of roles in bacterial biology; however, whether the role of OprB in sugar uptake is primary or a secondary effect cannot be distinguished because membrane integrity or integration of other proteins into the membrane was not assessed by Ekman et al. in 2013. In agreement with the data from Khamar et al. (Reference Khamar, Breathwaite, Prasse, Fraley, Secor, Chibane, Elhai and Chiu2010), all mutants for the fructose and glucose transporters were able to form hormogonia (Ekman et al., Reference Ekman, Picossi, Champbell, Meeks and Flores2013). Yet only mutants unable to produce the glucose permease resulted in the abortion of infection of the hornwort A. punctatus (Ekman et al., Reference Ekman, Picossi, Champbell, Meeks and Flores2013). Given that the sugar content, including glucose, is depleted in Gunnera glands once successfully colonised by cyanobacteria (Khamar et al., Reference Khamar, Breathwaite, Prasse, Fraley, Secor, Chibane, Elhai and Chiu2010) suggests that glucose may additionally be supplemented by the plant as food for the symbiont.

3.2. Symbiotic communication and integration into stress responses

Transcriptomic analyses of the interaction between A. filiculoides and T. azollae recently highlighted a chalcone synthase-encoding gene (CHS) that has highly induced transcript levels only in the presence of the cyanobiont (Eily et al., Reference Eily, Pryer and Li2019; Li et al., Reference Li, Brouwer, Carretero-Paulet, Cheng, de Vries, Delaux, Eily, Koppers, Kuo, Li, Simenc, Small, Wafula, Angarita, Barker, Bräutigam, dePamphilis, Gould, Hosmani and Pryer2018), suggesting a commitment to flavonoid biosynthesis in symbiosis. And indeed, deoxyanthocyanins alone can increase the induction of the hormogonia-repressing factor hrmA by 70% and even more so in combination with naringenin, but not with other phenylpropanoid(-derived) compounds that were tested by Cohen et al. (Reference Cohen, Sakihama, Takagi, Ichiba and Yamasaki2002). The inhibiting effect was also visible when extracts of leaf tissue from different Azolla species were used (Cohen et al., Reference Cohen, Sakihama, Takagi, Ichiba and Yamasaki2002). Agreeingly, several flavonoids were unable to promote hormogonia formation (Campbell & Meeks, Reference Campbell and Meeks1989; Nilsson et al., Reference Nilsson, Rasmussen and Bergman2006). Nonetheless, phenolics, including flavonoids, have been suggested to be involved in the transfer of cyanobionts to newly developed leaf cavities and sporocarps of Azolla (Pereira & Carrapiço, Reference Pereira and Carrapiço2007). Expression of homologs of genes encoding enzymes in the phenylpropanoid pathway (see also de Vries et al., Reference de Vries, Fürst-Jansen, Irisarri, Dhabalia Ashok, Ischebeck, Feussner, Abreu, Petersen, Feussner and de Vries2021b), providing the precursors for flavonoid biosynthesis, and flavonoid biosynthesis in A. filiculoides, is dependent on the circadian rhythm and the availability of nitrogen (Güngör et al., Reference Güngör, Brouwer, Dijkhuizen, Shaffar, Nierop, de Vos, Toraño, van der Meer and Schluepmann2021). Anthocyanins produced by Azolla are elevated under various stresses, including protection against light stress (Nham Tran et al., Reference Nham Tran, Miranda, Abeynayake and Mouradov2020). It is no surprise that genes coding for enzymes involved in anthocyanin biosynthesis are responsive to various triggers and stresses. One of the intriguing questions is what role do they play in the symbiosis. Given the permanent relationship of Azolla and its cyanobiont that needs to be maintained during stress, it is conceivable that stress-responsive metabolites, such as flavonoids, act as signals in the Azolla spp.–Trichormus symbioses. Proteomic analyses of salt stress on Azolla microphylla showed changes affecting the plant physiology and enzymes involved in N2-fixation of the cyanobiont (Thagela et al., Reference Thagela, Yadav, Mishra, Dahuja, Ahmad, Singh, Tiwari and Abraham2017; Yadav et al., Reference Yadav, Tripathi, Mishra, Ramteke, Singh and Abraham2019). The phytohormone salicylic acid (SA), which is involved in biotic stress in land plants (Pieterse et al., Reference Pieterse, Leon-Reyes, Van der Ent and Van Wees2009), alters cyanobiont abundance and nifE gene expression (de Vries et al., Reference de Vries, de Vries, Teschke, von Dahlen, Rose and Gould2018). Vice versa, transcriptomics and comparative genomic analyses highlight that the presence or absence of the cyanobiont alters the expression of homologous genes putatively involved in SA biosynthesis and signalling (de Vries et al., Reference de Vries, Herrfurth, Li, Feussner and de Vries2021a). More experimental and comparative data are needed to understand (a) how stress-responsive signalling integrates into Azolla’s symbiosis and (b) whether this can be transferred to non-permanent cyanobacterial symbioses.

3.3. Nutrient exchange from a genomic host perspective

A recurrent theme when analysing large-scale data from cyanobacteria and their hosts is, as already exemplified above, the transport of sugars to the cyanobionts as a counter-currency of the symbioses (Figure 3). Indeed, carbon transport from feathermosses to their epiphytic cyanobionts was demonstrated (Stuart et al., Reference Stuart, Pederson, Weyman, Weber, Rassmussen and Dupont2020). Transcriptomic analyses of the hornworts A. punctatus and A. agrestis with and without cyanobiont highlighted a sugar transporter from the SWEET1 clade that increases its transcript level in a symbiotic state (Li et al., Reference Li, Nishiyama, Waller, Frangedakis, Keller, Li, Fernandez-Pozo, Barker, Bennett, Blázquez, Cheng, Cuming, de Vries, de Vries, Delaux, Diop, Harrison, Hauser, Hernández-García and Szövényi2020). In the genome of A. filiculoides, 15 different SWEET genes have been identified, none of which can be linked to the symbioses based on transcriptomic data from cyanobiont-containing and cyanobiont-free cultures (Eily et al., Reference Eily, Pryer and Li2019). That does not mean that they play no role in symbiosis in general; furthermore, other sugar transporters could be involved. This illustrates that the picture cannot be transferred between symbiotic systems in a 1:1 manner. Additionally, the case of A. filiculoides might be special because of the perpetuity of its symbiosis. Whether facultative associations between evolutionary distinct plants and their cyanobionts use similar mechanisms is to be tested. Yet, it is not unthinkable that convergent molecular mechanisms have emerged that tip into the same pathways, given that the cyanobacteria appear capable of infecting many of these possible hosts.

4. Conclusion

Cyanobacteria can occur as symbionts in a diversity of plants that are separated by millions of years of evolution. Despite this diversity, genomic data indicate that among the facultative symbionts, similar molecular mechanisms are used to interact with their hosts. This is in agreement with the ability of these cyanobionts to infect distinct host lineages. In contrast, transcriptomic data on the cyanobacteria and their hosts hint that every host system recruits unique molecular features particular to each individual association; this is in line with the convergent evolution of these cyanobacterial symbioses and the unique characters and different degrees of intimacy observed in each system. It is conceivable that symbiotically competent cyanobacteria must have some plasticity in their response to the different hosts and an ability to recognise each individual species. Comparative genomics and transcriptomics have the power to illuminate shared patterns across these diverse systems.

Acknowledgement

We thank the anonymous reviewer for several valuable comments.

Financial support

J.d.V. thanks the European Research Council for funding under the European Union’s Horizon 2020 research and innovation programme (Grant Agreement No. 852725; ERC-StG ‘TerreStriAL’). J.d.V. is grateful for support through the German Research Foundation (DFG) within the framework of the Priority Programme ‘MAdLand—Molecular Adaptation to Land: Plant Evolution to Change’ (SPP 2237; VR 132/4-1).

Conflicts of interest

The authors declare no conflicts of interest.

Authorship contributions

S.d.V. and J.d.V. have outlined and written the manuscript together; both S.d.V. and J.d.V. created the figures.

Data availability statement

Data availability is not applicable to this article as no new data were created or analysed in this study.

References

Adams, D. G., Bergman, B., Nierzwicki-Bauer, S. A., Duggan, P. S., Rai, A. N., & Schüßler, A. (2013). Cyanobacterial-plant symbioses. In E. Rosenberg, E. F. DeLong, S. Lory, E. Stackebrandt, & F. Thompson (Eds.), The prokaryotes (pp. 360400). Springer Nature.Google Scholar
Adams, D. G., & Duggan, P. S. (2008). Cyanobacteria-bryophyte symbioses. Journal of Experimental Botany, 59, 10471058.CrossRefGoogle ScholarPubMed
Adams, D. G., & Duggan, P. S. (2012). Signalling in Cyanobacteria – Plant Symbioses. In S. Perotto, & F. Baluška (Eds), Signaling and communication in plant symbiosis (pp. 93122). Springer.CrossRefGoogle Scholar
Archibald, J. M. (2015). Genomic perspectives on the birth and spread of plastids. Proceedings of the National Academy of Sciences of the United States of America, 112, 1014710153.CrossRefGoogle ScholarPubMed
Appleby, C. A. (1984). Leghemoglobin and rhizobium respiration. Annual Review of Plant Physiology, 35, 443478.CrossRefGoogle Scholar
Bay, G., Nahar, N., Oubre, M., Whitehouse, M. J., Wardle, D. A., Zackrisson, O., Nilsson, M., & Rasmussen, U. (2013). Boreal feather mosses secrete chemical signals to gain nitrogen. New Phytologist, 200, 5460.CrossRefGoogle ScholarPubMed
Bell-Doyon, P., Laroche, J., Saltonstall, K., & Villarreal Aguilar, J. C. (2020). Specialized bacteriome uncovered in the coralloid roots of the epiphytic gymnosperm, Zamia pseudoparasitica . Environmental DNA, 2, 418428.CrossRefGoogle Scholar
Bengtson, S., Sallstedt, T., Belivanova, V., & Whitehouse, M. (2017). Three-dimensional preservation of cellular and subcellular structures suggests 1.6 billion-year-old crown-group red algae. PLoS Biology, 15, e2000735.CrossRefGoogle ScholarPubMed
Bergman, B., Matveyev, A., & Rasmussen, U. (1996). Chemical signalling in cyanobacterial-plant symbioses. Trends in Plant Science, 1, 191197.CrossRefGoogle Scholar
Bergman, B., Rai, A. N., Johansson, C., & Söderbäck, E. (1992). Cyanobacterial–plant symbioses. Symbiosis, 14, 6181.Google Scholar
Bouchard, R., Peñaloza-Bojacá, G., Toupin, S., Guadalupe, Y., Gudiño, J., Salazar Allen, N., Li, F.-W., & Villarreal Aguilar, J. C. (2020). Contrasting bacteriome of the hornwort Leiosporoceros dussii in two nearby sites with emphasis on the hornwort-cyanobacterial symbiosis, Symbiosis, 81, 3952.CrossRefGoogle Scholar
Brouwer, P., Bräutigam, A., Buijs, V. A., Tazelaar, A. O. E., van der Werf, A., Schlüter, U., Reichart, G.-J., Bolger, A., Usadel, B., Weber, A. P. M., & Schluepmann, H. (2017). Metabolic adaptation, a specialized leaf organ structure and vascular responses to diurnal N2 fixation by Nostoc azollae sustain the astonishing productivity of Azolla ferns without nitrogen fertilizer. Frontiers in Plant Science, 8, 442.CrossRefGoogle ScholarPubMed
Burke, C., Steinberg, P., Rusch, D., Kjelleberg, S., & Thomas, T. (2011). Bacterial community assembly based on functional genes rather than species. Proceedings of the National Academy of Sciences U.S.A., 108, 1428814293.CrossRefGoogle ScholarPubMed
Butterfield, N. J. (2000). Bangiomorpha pubescens n. gen., n. sp.: Implications for the evolution of sex, multicellularity, and the Mesoproterozoic/Neoproterozoic radiation of eukaryotes. Paleobiology, 26, 386404.2.0.CO;2>CrossRefGoogle Scholar
Calvert, H. E., Pence, M. K., & Peters, G. A. (1985). Ultrastructural ontogeny of leaf cavity trichomes in Azolla implies a functional role in metabolite exchange. Protoplasma, 129, 1027.CrossRefGoogle Scholar
Calvert, H. E., & Peters, G. A. (1981). The Azolla–Anabaena azollae relationship. New Phytologist, 89, 327335.CrossRefGoogle Scholar
Campbell, E. L., & Meeks, J. C. (1989). Characteristics of hormogonia formation by symbiotic Nostoc spp. in response to the presence of Anthoceros punctatus or its extracellular products. Applied and Environmental Microbiology, 55, 125131.CrossRefGoogle ScholarPubMed
Campbell, E. L., Wong, F. C. Y., & Meeks, J. C. (2003). DNA binding properties of the HrmR protein of Nostoc punctiforme responsible for transcriptional regulation of genes involved in the differentiation of hormogonia: DNA binding protein, HrmR, of Nostoc punctiforme . Molecular Microbiology, 47, 573582.CrossRefGoogle ScholarPubMed
Carrapiço, F. (1991). Are bacteria the third partner of the AzollaAnabaena symbiosis? In: M. Polsinelli, R. Materassi, & M. Vincenzini (Eds.), Nitrogen fixation (pp. 453456). Springer.CrossRefGoogle Scholar
Carrapiço, F. (2006). Is the Azolla–Anabaena symbiosis a co-evolution case? In A. Sitnykov (Ed.), “General Botany: Traditions and Perspectives”, Materials of the International Conference, dedicated to 200th anniversary of the Kazan Botanical School (January 23–27, 2006). Part I. (pp. 193195). Kazan University.Google Scholar
Carrapiço, F. (2010). Azolla as a superorganism. Its implication in symbiotic studies. In J. Seckbach & M. Grube (Eds.), Symbioses and stress: Joint ventures in biology, Cellular Origin, Life in Extreme Habitats and Astrobiology, 17 (pp. 225241). Springer.CrossRefGoogle Scholar
Carrapiço, F., & Tavares, R. (1989). New data on the Azolla–Anabaena symbiosis. II. Cytochemical and immunocytochemical aspects. In F. A. Skinner, R. M. Boddey, & I. Frederik (Eds.), Nitrogen fixation with nonlegumes (pp. 8994). Kluwer Academic.CrossRefGoogle Scholar
Chatterjee, P., Schafran, P., Li, F.-W., & Meeks, J. C. (2021). Nostoc talks back: Temporal patterns of differential gene expression during establishment of the Anthoceros–Nostoc symbiosis. BioRxiv preprint. https://doi.org/10.1101/2021.10.27.465970 CrossRefGoogle Scholar
Chen, M.-Y., Teng, W.-K., Zhao, L., Hu, C.-X., Zhou, Y.-K., Han, B.-P., Song, L.-R., & Shu, W.-S. (2021). Comparative genomics reveals insights into cyanobacterial evolution and habitat adaptation. The ISME Journal, 15, 211227.CrossRefGoogle ScholarPubMed
Christman, H. D., Campbell, E. L., & Meeks, J. C. (2011). Global transcription profiles of the nitrogen stress response resulting in heterocyst or hormogonium development in Nostoc punctiforme . Journal of Bacteriology, 193, 68746886.CrossRefGoogle ScholarPubMed
Cohen, M. F., & Meeks, J. C. (1997). A hormogonium regulating locus, hrmUA, of the cyanobacterium Nostoc punctiforme strain ATCC 29133 and its response to an extract of a symbiotic plant partner Anthoceros punctatus . Molecular Plant–Microbe Interactions, 10, 280289.CrossRefGoogle Scholar
Cohen, M. F., Sakihama, Y., Takagi, Y. C., Ichiba, T., & Yamasaki, H. (2002). Synergistic effect of deoxyanthocyanins from symbiotic fern Azolla spp. on hrmA gene induction in the cyanobacterium Nostoc punctiforme . Molecular Plant-Microbe Interactions, 15, 875882.CrossRefGoogle ScholarPubMed
Collinson, M. E. (2002). The ecology of Cainozoic ferns. Review of Palaeobotany and Palynology, 119, 5168.CrossRefGoogle Scholar
Costa, J. L., Paulsrud, P., & Lindblad, P. (1999). Cyanobiont diversity within coralloid roots of selected cycad species. FEMS Microbiology Ecology, 28, 8591.CrossRefGoogle Scholar
Curatti, L., Ludden, P. W., & Rubio, L. M. (2006). NifB-dependent in vitro synthesis of the iron–molybdenum cofactor of nitrogenase. Proceedings of the National Academy of Sciences of the United States of America, 103, 52975301.CrossRefGoogle ScholarPubMed
Dagan, T., Roettger, M., Stucken, K., Landan, G., Koch, R., Major, P., Gould, S. B., Goremykin, V. V., Rippka, R., Tandeau de Marsac, N., Gugger, M., Lockhart, P. J., Allen, J. F., Brune, I., Maus, I., Pühler, A., & Martin, W. F. (2013). Genomes of Stigonematalean cyanobacteria (subsection V) and the evolution of oxygenic photosynthesis from prokaryotes to plastids. Genome Biology and Evolution, 5, 3144.CrossRefGoogle ScholarPubMed
Darnajoux, R., Magain, N., Renaudin, M., Lutzoni, F., Bellenger, J.-F., & Zhang, X. (2019). Molybdenum threshold for ecosystem scale alternative vanadium nitrogenase activity in boreal forests. Proceedings of the National Academy of Sciences of the United States of America, 116, 2468224688.CrossRefGoogle ScholarPubMed
de Vries, J., & Archibald, J. M. (2017). Endosymbiosis: Did plastids evolve from a freshwater cyanobacterium? Current Biology, 27, R103R105.CrossRefGoogle ScholarPubMed
de Vries, J., Fischer, A. M., Roettger, M., Rommel, S., Schluepmann, H., Bräutigam, A., Carlsbecker, A., & Gould, S. B. (2016). Cytokinin-induced promotion of root meristem size in the fern Azolla supports a shoot-like origin of euphyllophyte roots. New Phytologist, 209, 705720.CrossRefGoogle ScholarPubMed
de Vries, S., & de Vries, J. (2018). Azolla: A model system for symbiotic nitrogen fixation and evolutionary developmental biology. In H. Fernández (Ed.), Current advances in fern research. (pp. 2146). Springer.CrossRefGoogle Scholar
de Vries, S., de Vries, J., Teschke, H., von Dahlen, J. K., Rose, L. E., & Gould, S. B. (2018). Jasmonic and salicylic acid response in the fern Azolla filiculoides and its cyanobiont. Plant, Cell and Environment, 41, 25302548.CrossRefGoogle ScholarPubMed
de Vries, S., Fürst-Jansen, J. M. R., Irisarri, I., Dhabalia Ashok, A., Ischebeck, T., Feussner, K., Abreu, I. N., Petersen, M., Feussner, I., & de Vries, J. (2021b). The evolution of the phenylpropanoid pathway entailed pronounced radiation and divergence of enzyme families. The Plant Journal, 107, 9751002.CrossRefGoogle Scholar
de Vries, S., Herrfurth, C., Li, F.-W., Feussner, I., & de Vries, J. (2021a). An ancient route towards salicylic acid and its implications for the perpetual TrichormusAzolla symbiosis, BioRxiv preprint. https://doi.org/10.1101/2021.03.12.435107 CrossRefGoogle Scholar
Dean, D. R., Bolin, J. T., & Zheng, L. (1993). Nitrogenase metalloclusters: Structures, organization, and synthesis. Journal of Bacteriology, 175, 67376744.CrossRefGoogle ScholarPubMed
Dijkhuizen, L. W., Brouwer, P., Bolhuis, H., Reichart, G.-J., Koppers, N., Huettel, B., Bolger, A. M., Li, F.-W., Cheng, S., Liu, X., Wong, G. K.-S., Pryer, K., Weber, A., Bräutigam, A., & Schluepmann, H. (2018). Is there foul play in the leaf pocket? The metagenome of floating fern Azolla reveals endophytes that do not fix N2 but may denitrify. New Phytologist, 217, 453466.CrossRefGoogle Scholar
Dijkhuizen, L. W., Tabatabaei, B. E. S., Brouwer, P., Rijken, N., Buijs, V. A., Güngör, E., & Schluepmann, H. (2021). Far-red light-induced Azolla filiculoides symbiosis sexual reproduction: responsive transcripts of symbiont Nostoc azollae encode transporters whilst those of the fern relate to the angiosperm floral transition. Frontiers in Plant Science, 12, 693039.CrossRefGoogle Scholar
Doolittle, W. F., & Inkpen, S. A. (2018). Processes and patterns of interaction as units of selection: an introduction to ITSNTS thinking. Proceedings of the National Academy of Sciences U.S.A., 115, 40064014.CrossRefGoogle ScholarPubMed
Dos Santos, P. C., Smith, A. D., Frazzon, J., Cash, V. L., Johnson, M. K., & Dean, D. R. (2004). Iron–sulfur cluster assembly. Journal of Biological Chemistry, 279, 1970519711.CrossRefGoogle ScholarPubMed
Duggan, P. S., Thiel, T., & Adams, D. G. (2013). Symbiosis between the cyanobacterium Nostoc and the liverwort Blasia requires a CheR-type MCP methyltransferase. Symbiosis, 59, 110120.CrossRefGoogle Scholar
Eily, A. N., Pryer, K. M., & Li, F.-W. (2019). A first glimpse at genes important to the Azolla–Nostoc symbiosis. Symbiosis, 78, 149162.CrossRefGoogle Scholar
Ekman, M., Picossi, S., Champbell, E. L., Meeks, J. C., & Flores, E. (2013). A Nostoc punctiforme sugar transporter necessary to establish a cyanobacterium–plant symbiosis. Plant Physiology, 161, 19841992.CrossRefGoogle ScholarPubMed
Ekman, M., Tollbäck, P., & Bergman, B. (2008). Proteomic analysis of the cyanobacterium of the Azolla symbiosis: Identity, adaptation, and NifH modification. Journal of Experimental Botany, 59, 10231034.CrossRefGoogle ScholarPubMed
Ekman, M., Tollbäck, P., Klint, J., & Bergman, B. (2006). Protein expression profiles in an endosymbiotic cyanobacterium revealed by a proteomic approach. Molecular Plant-Microbe Interactions, 19, 12511261.CrossRefGoogle Scholar
Eme, L., Sharpe, S. C., Brown, M W., & Roger, A. J. (2014). On the age of eukaryotes: Evaluating evidence from fossils and molecular clocks. Cold Spring Harbor Perspectives in Biology, 6, a016139.CrossRefGoogle ScholarPubMed
Fitzpatrick, C. R., Salas-González, I., Conway, J. M., Finkel, O. M., Gilbert, S., Russ, D., Teixeira, P. J. P. L., & Dangl, J. L. (2020). The plant microbiome: From ecology to reductionism and beyond. Annual Review of Microbiology, 74, 81100.CrossRefGoogle ScholarPubMed
Frangedakis, E., Shimamura, M., Villarreal, J. C., Li, F., Tomaselli, M., Waller, M., Sakakibara, K., Renzaglia, K. S., & Szövényi, P. (2021). The hornworts: Morphology, evolution and development. New Phytologist, 229, 735754.CrossRefGoogle ScholarPubMed
Gallon, J. R. (1981). The oxygen sensitivity of nitrogenase: A problem for biochemists and micro-organisms. Trends in Biochemical Sciences, 6, 1923.CrossRefGoogle Scholar
Grilli Caiola, M., Canini, A., & Moscone, D. (1989). Oxygen concentration, nitrogenase activity and heterocyst frequency in the leaf cavities of Azolla filiculoides Lam. FEMS Microbiology Letters, 59, 283288.CrossRefGoogle Scholar
Gutiérrez-García, K., Bustos-Díaz, E. D., Corona-Gómez, J. A., Ramos-Aboites, H. E., Sélem-Mojica, N., Cruz-Morales, P., Pérez-Farrera, M. A., Barona-Gómez, F., & Cibrián-Jaramillo, A. (2019). Cycad coralloid roots contain bacterial communities including cyanobacteria and Caulobacter spp. that encode niche-specific biosynthetic gene clusters. Genome Biology and Evolution, 11, 319334.CrossRefGoogle ScholarPubMed
Güngör, E., Brouwer, P., Dijkhuizen, L. W., Shaffar, D. C., Nierop, K. G. J., de Vos, R. C. H., Toraño, J. S., van der Meer, I. M., & Schluepmann, H. (2021). Azolla ferns testify: seed plants and ferns share a common ancestor for leucoanthocyanidin reductase enzymes. New Phytologist, 229, 11181132.CrossRefGoogle Scholar
Hashidoko, Y., Nishizuka, H., Tanaka, M., Murata, K., Murai, Y., & Hashimoto, M. (2019). Isolation and characterization of 1-palmitoyl-2-linoleoyl-sn-glycerol as a hormogonium-inducing factor (HIF) from the coralloid roots of Cycas revoluta (Cycadaceae). Scientific Reports, 9, 4751.CrossRefGoogle Scholar
Hill, D. J. (1975). The pattern and development of Anabaena in the Azolla – Anabaena symbiosis. Planta, 122, 179184.CrossRefGoogle ScholarPubMed
Hill, D. J. (1977). The role of Anabaena in the Azolla – Anabaena symbiosis. New Phytologist, 78, 611616.CrossRefGoogle Scholar
Hill, D. J. (1989). The control of the cell cycle in microbial symbionts. New Phytologist, 112, 175184.CrossRefGoogle Scholar
Ininbergs, K., Bay, G., Rasmussen, U., Wardl, D. A., & Nisson, M.-C. (2011). Composition and diversity of nifH genes of nitrogen-fixing cyanobacteria associated with boreal forest feather mosses. New Phytologist, 192, 507517.CrossRefGoogle ScholarPubMed
Ishikura, N. (1982). 3‐Desoxyanthocyanin and other phenolics in the water fern Azolla . The Botanical Magazine, 95, 303308.CrossRefGoogle Scholar
Johansson, C., & Bergman, B. (1992). Early events during the establishment of the Gunnera/Nostoc symbiosis. Planta, 188, 403413.CrossRefGoogle ScholarPubMed
Johnson, D. C., Dos Santos, P. C., & Dean, D. R. (2005). NifU and NifS are required for the maturation of nitrogenase and cannot replace the function of isc-gene products in Azotobacter vinelandii . Biochemical Society Transactions, 33, 9093.CrossRefGoogle ScholarPubMed
Khamar, H. J., Breathwaite, E. K., Prasse, C. E., Fraley, E. R., Secor, C. R., Chibane, F. L., Elhai, J., & Chiu, W.-L. (2010). Multiple roles of soluble sugars in the establishment of GunneraNostoc endosymbiosis. Plant Physiology, 154, 13811389.CrossRefGoogle ScholarPubMed
Kneip, C., Voβ, C., Lockhart, P. J., & Maier, U. G. (2008). The cyanobacterial endosymbiont of the unicellular algae Rhopalodia gibba shows reductive genome evolution. BMC Evolutionary Biology, 8, 30.CrossRefGoogle ScholarPubMed
Kumar, K., Mella-Herrera, R. A., & Golden, J. W. (2010). Cyanobacterial heterocysts. Cold Spring Harbor Perspectives in Biology, 2, a000315.CrossRefGoogle ScholarPubMed
Li, F.-W., Brouwer, P., Carretero-Paulet, L., Cheng, S., de Vries, J., Delaux, P.-M., Eily, A., Koppers, N, Kuo, L.-Y., Li, Z., Simenc, M., Small, I., Wafula, E., Angarita, S., Barker, M. S., Bräutigam, A., dePamphilis, C., Gould, S., Hosmani, P. S., …, & Pryer, K. M. (2018). Fern genomes elucidate land plant evolution and cyanobacterial symbioses. Nature Plants, 4, 460472.CrossRefGoogle ScholarPubMed
Li, F.-W., Nishiyama, T., Waller, M., Frangedakis, E., Keller, J., Li, Z., Fernandez-Pozo, N., Barker, M. S., Bennett, T., Blázquez, M. A., Cheng, S., Cuming, A. C., de Vries, J., de Vries, S., Delaux, P.-M., Diop, I. S., Harrison, J. C., Hauser, D., Hernández-García, J., …, & Szövényi, P. (2020). Anthoceros genomes illuminate the origin of land plants and the unique biology of hornworts. Nature Plants, 6, 259272.CrossRefGoogle ScholarPubMed
Liaimer, A., Jensen, J. B., & Dittmann, E. (2016). A genetic and chemical perspective on symbiotic recruitment of cyanobacteria of the genus Nostoc into the host plant Blasia pusilla L. Frontiers in Microbiology, 7, 1693.CrossRefGoogle ScholarPubMed
Lundberg, D. S., Lebeis, S. L., Paredes, S. H., Yourstone, S., Gehring, J., Malfatti, S., Tremblay, J., Engelbrektson, A., Kunin, V., del Rio, T. G., Edgar, R. C., Eickhorst, T., Ley, R. E., Hugenholtz, P., Green Tringe, S., & Dangl, J. L. (2012). Defining the core Arabidopsis thaliana root microbiome. Nature, 488, 8690.CrossRefGoogle ScholarPubMed
Matasci, N., Hung, L.-H., Yan, Z., Carpenter, E. J., Wickett, N. J., Mirarab, S., Nguyen, N., Warnow, T., Ayyampalayam, S., Barker, M., Burleigh, J. G., Gitzendanner, M. A., Wafula, E., Der, J. P., dePamphilis, C. W., Roure, B., Philippe, H., Ruhfel, B. R., Miles, N. W., …, & Wong, G. K.-S. (2014). Data access for the 1,000 Plants (1KP) project. GigaScience, 3, 17.CrossRefGoogle ScholarPubMed
Meeks, J. C. (1998). Symbiosis between nitrogen-fixing cyanobacteria and plants. BioScience, 48, 266276.CrossRefGoogle Scholar
Meeks, J. C., Elhai, J., Thiel, T., Potts, M., Larimer, F., Lamerdin, J., Predki, P., & Atlas, R. (2001). An overview of the genome of Nostoc punctiforme, a multicellular, symbiotic cyanobacterium. Photosynthesis Research, 70, 85106.CrossRefGoogle ScholarPubMed
Meeks, J. C., & Elhai, J. (2002). Regulation of cellular differentiation in filamentous cyanobacteria in free-living and plant-associated symbiotic growth states. Microbiology and Molecular Biology Reviews, 66, 94121.CrossRefGoogle ScholarPubMed
Meeks, J. C., Joseph, C. M., & Haselkorn, R. (1988). Organization of the nif genes in cyanobacteria in symbiotic association with Azolla and Anthoceros . Archives of Microbiology, 150, 6171.CrossRefGoogle ScholarPubMed
Nakayama, T., Kamikawa, R., Tanifuji, G., Kashiyama, Y., Ohkouchi, N., Archibald, J. M., & Inagaki, Y. (2014). Complete genome of a nonphotosynthetic cyanobacterium in a diatom reveals recent adaptations to an intracellular lifestyle. Proceedings of the National Academy of Sciences of the United States of America, 111, 1140711412.CrossRefGoogle Scholar
Nelson, J. M., Hauser, D. A., Gudiño, J. A., Guadalupe, Y. A., Meeks, J. C., Salazar Allen, N., Villarreal, J. C., & Li, F.-W. (2019). Complete genomes of symbiotic cyanobacteria clarify the evolution of vanadium-nitrogenase. Genome Biology and Evolution, 11, 19591964.CrossRefGoogle ScholarPubMed
Nelson, J. M., Hauser, D. A., & Li, F.-W. (2021). The diversity and community structure of symbiotic cyanobacteria in hornworts inferred from long-read amplicon sequencing. American Journal of Botany, 108, 17311744.CrossRefGoogle ScholarPubMed
Nham Tran, T. L., Miranda, A. F., Abeynayake, S. W., & Mouradov, A. (2020). Differential production of phenolics, lipids, carbohydrates and proteins in stressed and unstressed aquatic plants, Azolla filiculoides and Azolla pinnata . Biology, 9, 342.CrossRefGoogle Scholar
Nilsson, M., Bergman, B., & Rasmussen, U. (2000). Cyanobacterial diversity in geographically related and distant host plants of the genus Gunnera . Archives of Microbiology., 173, 97102.CrossRefGoogle ScholarPubMed
Nilsson, M., Rasmussen, U., & Bergman, B. (2006). Cyanobacterial chemotaxis to extracts of host and nonhost plants. FEMS Microbiology Ecology, 55, 382390.CrossRefGoogle ScholarPubMed
Ochoa de Alda, J. A. G., Esteban, R., Diago, M. L., & Houmard, J. (2014). The plastid ancestor originated among one of the major cyanobacterial lineages. Nature Communications, 5, 4937.CrossRefGoogle ScholarPubMed
Ow, M. C., Gantar, M., & Elhai, J. (1999). Reconstitution of a cycad-cyanobacterial association. Symbiosis, 27, 125134.Google Scholar
Papaefthimiou, D., Van Hove, C., Lejeune, A., Rasmussen, U., & Wilmotte, A. (2008). Diversity and host specificity of Azolla cyanobionts. Journal of Phycology, 44, 6070.CrossRefGoogle ScholarPubMed
Parniske, M. (2008). Arbuscular mycorrhiza: The mother of plant root endosymbioses. Nature Reviews Microbiology, 6, 763775.CrossRefGoogle ScholarPubMed
Pederson, E. R. A., Warshan, D., & Rasmussen, U. (2019). Genome sequencing of Pleurozium schreberi: the assembled and annotated draft genome of a pleurocarpous feather moss. G3 Genes|Genomes|Genetics, 9, 27912797.CrossRefGoogle ScholarPubMed
Pereira, A. L., & Carrapiço, F. (2007). Histochemistry of simple hairs from the foliar cavities of Azolla filiculoides . Plant Biosystems, 141, 323328.CrossRefGoogle Scholar
Pereira, A. L., & Vasconcelos, V. (2014). Classification and phylogeny of the cyanobiont Anabaena azollae Strasburger: An answered question? International Journal of Systematic and Evolutionary Microbiology, 64, 18301840.CrossRefGoogle ScholarPubMed
Perkins, S. K., & Peters, G. A. (1993). The Azolla–Anabaena symbiosis: Endophyte continuity in the Azolla life-cycle is facilitated by epidermal trichomes. I. Partitioning of the endophytic Anabaena into developing sporocarps. New Phyotologist, 123, 5364.CrossRefGoogle Scholar
Peters, G. A., & Meeks, J. C. (1989). The AzollaAnabaena symbiosis: Basic biology. Annual Review of Plant Biology, 40, 193210.CrossRefGoogle Scholar
Peters, G. A., Toia, R. E., Raveed, D., & Levine, N. J. (1978). The AzollaAnabaena azollae relationship VI. Morphological aspects of the association. New Phytologist, 80, 583593.CrossRefGoogle Scholar
Pieterse, C. M. J., Leon-Reyes, A., Van der Ent, S., & Van Wees, S. C. M. (2009). Networking by small-molecule hormones in plant immunity. Nature Chemical Biology, 5, 308316.CrossRefGoogle ScholarPubMed
Ponce-Toledo, R. I., Deschamps, P., López-García, P., Zivanovic, Y., Benzerara, K., & Moreira, D. (2017). An early-branching freshwater cyanobacterium at the origin of plastids. Current Biology, 27, 386391.CrossRefGoogle ScholarPubMed
Pratte, B. S., Sheridan, R., James, J. A., & Thiel, T. (2013). Regulation of V-nitrogenase genes in Anabaena variabilis by RNA processing and by dual repressors. Molecular Microbiology, 88, 413424.CrossRefGoogle ScholarPubMed
Pratte, B. S., & Thiel, T. (2006). High-affinity vanadate transport system in the cyanobacterium Anabaena variabilis ATCC 29413. Journal of Bacteriology, 188, 464468.CrossRefGoogle ScholarPubMed
Pratte, B. S., & Thiel, T. (2021). Comparative genomic insights into culturable symbiotic cyanobacteria from the water fern Azolla. Microbial Genomics, 7, 000595.CrossRefGoogle ScholarPubMed
Puttick, M. N., Morris, J. L., Williams, T. A., et al. (2018). The interrelationships of land plants and the nature of the ancestral embryophyte. Current Biology, 28, 733745.CrossRefGoogle ScholarPubMed
Qi, X., Kuo, L.-Y., Guo, C., Li, H., Li, Z., Qi, J., Wang, L., Hu, Y., Xiang, J., Zhang, C., Guo, J., Huang, C.-H., & Ma, H. (2018). A well-resolved fern nuclear phylogeny reveals the evolution history of numerous transcription factor families. Molecular Phylogenetics and Evolution, 127, 961977.CrossRefGoogle ScholarPubMed
Qiu, G.-W., Jiang, H.-B., Lis, H., Li, Z.-K., Deng, B., Shang, J.-L., Sun, C. Y., Keren, N., & Qiu, B.-S. (2021). A unique porin meditates iron-selective transport through cyanobacterial outer membranes. Environmental Microbiology, 23, 379390.CrossRefGoogle ScholarPubMed
Radhakrishnan, G. V., Keller, J., Rich, M. K., Vernié, T., Mbadinga Mbadinga, D. L., Vigneron, N., Cottret, L., San Clemente, H., Libourel, C., Cheema, J., Linde, A.-M., Eklund, , Shifeng Cheng, S., Wong, G. K. S., Lagercrantz, U., Li, F.-W., Oldroyd, G. E. D., & Delaux, P.-M. (2020) An ancestral signalling pathway is conserved in intracellular symbioses-forming plant lineages. Nature Plants, 6, 280289.CrossRefGoogle ScholarPubMed
Rai, A. N., Borthakur, M., Singh, S., & Bergman, B. (1989). AnthocerosNostoc symbiosis: Immunoelectronmicroscopic localization of nitrogenase, glutamine synthetase, phycoerythrin and ribulose-1,5-bisphosphate carboxylase/oxygenase in the cyanobiont and the cultured (free-living) isolate Nostoc 7801. Journal of General Microbiology, 135, 385395.Google Scholar
Rai, A. N., Soderback, E., & Bergman, B. (2000). Cyanobacterium-plant symbioses. New Phytologist, 147, 449481.CrossRefGoogle ScholarPubMed
Ran, L., Larsson, J., Vigil-Stenman, T., Nylander, J. A. A., Ininbergs, K., Zheng, W.-W., Lapidus, A., Lowry, S., Robert Haselkorn, R., & Bergman, B. (2010). Genome erosion in a nitrogen-fixing vertically transmitted endosymbiotic multicellular cyanobacterium. PLoS One, 5, e11486.CrossRefGoogle Scholar
Rasmussen, U., Johansson, C., & Bergman, B. (1994). Early communication in the Gunnera-Nostoc symbiosis: plant-induced cell differentiation and protein synthesis in the cyanobacterium. Molecular Plant-Microbe Interactions, 7, 696702.CrossRefGoogle Scholar
Reinke, J. (1872). Über die anatomischen Verhältnisse einiger Arten von Gunnera L. Nachrichten von der Königlichen Gesellschaft der Wissenschaften und der Georg-August-Universität, 57, 100108.Google Scholar
Roy, R., Reinders, A., Ward, J. M., & McDonald, T. R. (2020). Understanding transport processes in lichen, Azolla–cyanobacteria, ectomycorrhiza, endomycorrhiza, and rhizobia–legume symbiotic interactions. F1000Research, 9, 39.CrossRefGoogle ScholarPubMed
Sapp, M., Ploch, S., Fiore-Donno, A. M., Bonkowski, M., & Rose, L. E. (2018). Protists are an integral part of the Arabidopsis thaliana microbiome: Protists associated with Arabidopsis thaliana . Environmental Microbiology, 20, 3043.CrossRefGoogle Scholar
Schätzle, H., Brouwer, E.-M., Liebhart, E., Stevanovic, M., & Schleiff, E. (2021). Comparative phenotypic analysis of Anabaena sp. PCC 7120 mutants of porin-like genes. Journal of Microbiology and Biotechnology, 31, 645658.CrossRefGoogle Scholar
Schlaeppi, K., Dombrowski, N., Oter, R. G., Ver Loren van Themaat, E., & Schulze-Lefert, P. (2014). Quantitative divergence of the bacterial root microbiota in Arabidopsis thaliana relatives. Proceedings of the National Academy of Sciences of the United States of America, 111, 585592.CrossRefGoogle ScholarPubMed
Stewart, W. D. P., & Rodgers, G. A. (1977). The cyanophyte-hepatic symbiosis II: nitrogen fixation and the interchange of nitrogen and carbon. New Phytologist, 78, 459471.CrossRefGoogle Scholar
Stuart, R. K., Pederson, E. R. A., Weyman, P. D., Weber, P. K., Rassmussen, U., & Dupont, C. L. (2020). Bidirectional C and N transfer and a potential role for sulfur in an epiphytic diazotrophic mutualism. The ISME Journal, 14, 30683078.CrossRefGoogle Scholar
Söderbäck, E., & Bergman, B. (1993). The Nostoc-Gunnera symbiosis: carbon fixation and translocation. Physiologia Plantarum, 89, 125132.CrossRefGoogle Scholar
Suárez-Moo, P. d. J., Vovides, A. P., Griffith, M. P., Barona-Gómez, F., Cibrián-Jaramillo, A. (2019). Unlocking a high bacterial diversity in the coralloid root microbiome from the cycad genus Dioon . PLoS ONE, 14, e0211271.CrossRefGoogle ScholarPubMed
Thagela, P., Yadav, R. K., Mishra, V., Dahuja, A., Ahmad, A., Singh, P. K., Tiwari, B. S., & Abraham, G. (2017). Salinity-induced inhibition of growth in the aquatic pteridophyte Azolla microphylla primarily involves inhibition of photosynthetic components and signaling molecules as revealed by proteome analysis. Protoplasma, 254, 303313.CrossRefGoogle ScholarPubMed
Thiel, T. (1996). Isolation and characterization of the vnfEN genes of the cyanobacterium Anabaena variabilis . Journal of Bacteriology, 178, 44934499.CrossRefGoogle ScholarPubMed
Thiel, T., Lyons, E. M., & Erker, J. C. (1997). Characterization of genes for a second Mo-dependent nitrogenase in the cyanobacterium Anabaena variabilis . Journal of Bacteriology, 179, 52225225.CrossRefGoogle ScholarPubMed
Thiel, T., & Pratte, B. (2001). Effect on heterocyst differentiation of nitrogen fixation in vegetative cells of the cyanobacterium Anabaena variabilis ATCC 29413. Journal of Bacteriology, 183, 280286.CrossRefGoogle ScholarPubMed
Towata, E. (1985). Mucilage glands and cyanobacterial colonization in Gunnera kaalensis (Haloragaceae). Botanical Gazette, 146, 5662.CrossRefGoogle Scholar
Trivedi, P., Leach, J. E., Tringe, S. G., Sa, T., & Singh, B. K. (2020). Plant–microbiome interactions: From community assembly to plant health. Nature Reviews Microbiology, 18, 607621.CrossRefGoogle ScholarPubMed
Wallace, W. H., & Gates, J. E. (1986). Identification of Eubacteria isolated from leaf cavities of four species of the N-Fixing Azolla fern as Arthrobacter Conn and Dimmick. Applied and Environmental Microbiology, 52, 425429.CrossRefGoogle ScholarPubMed
Wang, C.-M., Ekman, M., & Bergman, B. (2004). Expression of cyanobacterial genes involved in heterocyst differentiation and dinitrogen fixation along a plant symbiosis development profile. Molecular Plant-Microbe Interactions, 17, 436443.CrossRefGoogle ScholarPubMed
Warshan, D., Espinoza, J. L., Stuart, R. K., Richter, R. A., Kim, S.-Y., Shapiro, N., Woyke, T., Kyrpides, N. C., Barry, K., Singan, V., Lindquist, E., Ansong, C., Purvine, S. O., Brewer, H. M., Weyman, P. D., Dupont, C. L., & Rasmussen, U. (2017). Feathermoss and epiphytic Nostoc cooperate differently: expanding the spectrum of plant–cyanobacteria symbiosis. The ISME Journal, 11, 28212833.CrossRefGoogle ScholarPubMed
Warshan, D., Liaimer, A., Pederson, E., Kim, S.-Y., Shapiro, N., Woyke, T., Altermark, B., Pawlowski, K., Weyman, P. D., Dupont, C. L., & Rasmussen, U. (2018). Genomic changes associated with the evolutionary transitions of Nostoc to a plant symbiont. Molecular Biology and Evolution, 35, 11601175.CrossRefGoogle ScholarPubMed
Yadav, R. K., Tripathi, K., Mishra, V., Ramteke, P. W., Singh, P. K., & Abraham, G. (2019). Proteomic evaluation of the freshly isolated cyanobionts from Azolla microphylla exposed to salinity stress. Symbiosis, 77, 249256.CrossRefGoogle Scholar
Zhang, J., Fu, X.-X., Li, R.-Q., Zhao, X., Liu, Y., Li, M.-H., Zwaenepoel, A., Ma, H., Goffinet, B., Guan, Y.-L., Xue, J.-Y., Liao, Y.-Y., Wang, Q.-F., Wang, Q.-H., Wang, J.-Y., Zhang, G.-Q., Wang, Z.-W., Jia, Y., Wang, M.-Z., …, & Chen, Z.-D. (2020). The hornwort genome and early land plant evolution. Nature Plants, 6, 107118.CrossRefGoogle ScholarPubMed
Zheng, W., Bergman, B., Chen, B., Zheng, S., Xiang, G., & Rasmussen, U. (2009a). Cellular responses in the cyanobacterial symbiont during its vertical transfer between plant generations in the Azolla microphylla symbiosis. New Phytologist, 181, 5361.CrossRefGoogle Scholar
Zheng, W., Rang, L., & Bergman, B. (2009b). Structural characteristics of the cyanobacterium–Azolla symbioses. In K. Pawlowski (Ed.), Prokaryotic symbionts in plants. Microbiology monographs (Vol. 8, pp. 235263). Springer.CrossRefGoogle Scholar
Zheng, W. W., Nilsson, M., Bergman, B., & Rasmussen, U. (1999). Genetic diversity and classification of cyanobacteria in different Azolla species by the use of PCR fingerprinting. Theoretical and Applied Genetics, 99, 11871193.CrossRefGoogle Scholar
Zilliges, Y., Kehr, J.-C., Meissner, S., Ishida, K., Mikkat, S., Hagemann, M., Kaplan, A., Börner, T., & Dittmann, E. (2011). The cyanobacterial hepatotoxin microcystin binds to proteins and increases the fitness of Microcystis under oxidative stress conditions. PLoS One, 6, e17615.CrossRefGoogle ScholarPubMed
Zimmerman, W. J., Rosen, B., & Lumpkin, T. (1989). Enzymatic, lectin, and morphological characterization and classification of presumptive cyanobionts from Azolla lam. New Phytologist, 113, 497503.CrossRefGoogle Scholar
Figure 0

Figure 1. Cladogram of symbiotic interactions across land plants and diatoms. To the left, a simplified cladogram of the phylogenetic relationship among streptophytes (based on Puttick et al., 2018) is shown; diatoms (algae with a plastid of red algal origin) are very distantly related to streptophytes—the most recent common ancestor they share might be the last common ancestor of all eukaryotes (LECA). Among all major lineages of land plants except lycophytes (bold font), a few species that engage in symbiosis with cyanobacteria can be found (written in regular and italic font). To the right, filled dots denote key traits of the cyanobionts that dwell within those plants (and diatoms).

Figure 1

Figure 2. The AzollaTrichormus symbiosis. At the top, the sporophyte body of Azolla filiculoides whose leaf cavities are packed with cyanobionts (red autofluorescence) is shown. At the bottom, a fluorescence micrograph of the filamentous cyanobionts (Trichormus azollae) within one leaf cavity are shown (red autofluorescence); note the heterocysts, which are larger than the vegetative cells.

Figure 2

Figure 3. Syntrophic currencies exchanged by plant hosts and cyanobionts. A schematic Nostocales filament, consisting of vegetative cells and a heterocyst, is shown. The plant provides sugars (pink font), and the cyanobionts fix atmospheric nitrogen and provide it to the host via export (orange font). For an interesting discussion on putative transporters, see Roy et al. (2020).

Figure 3

Figure 4. Genome characteristics of genomes from symbiotic and nonsymbiotic cyanobacteria and cyanobacterial host plants. (a) Genome size of several cyanobacteria shown to be symbiotic or free-living (blue, cyano) and all sequenced plant hosts for cyanobacteria (yellow, plants). (b) Number of annotated protein-coding genes in the genomes of cyanobacteria (blue, cyano) and cyanobacterial host plants (yellow, plants). (c) Number of pseudogenes described for the here included cyanobacterial genome data (for details, see description below). (d) Relative repeat content in the assembled plant genomes from plants symbiotically associating with cyanobacteria. To have a mixed set of cyanobacteria associating with plants and free-living, we included cyanobacteria previously sequenced and/or included in the comparative analyses by Pratte & Thiel (2021) and Nelson et al. (2019). NCBI accessions have been derived from these publications, and actual data have been obtained from the National Center for Biotechnology Information (NCBI). For cyanobacterial hosts, all host species that have been sequenced (Anthoceros agrestis, Anthoceros punctatus, Anthoceros angustus, Azolla filiculoides, Sphagnum fallax and Pleurozium schreberi) are included in the figure. The data are derived from Li et al. (2018, 2020), Pederson et al. (2019), Zhang et al. (2020), and Genome statistic information from Phytozome genome ID: 522. Data on repeat content were not reported for S. fallax, and the number of pseudogenes has not been reported for cyanobacteria isolated from feathermosses.

Author comment: Evolutionary genomic insights into cyanobacterial symbioses in plants — R0/PR1

Comments

Dear Dr. Hamant,

Thank you for the invitation to submit our Review article "Evolutionary genomic insights into cyanobacterial symbioses in plants" to Quantitative Plant Biology.

As previously outlined, we discuss recent developments in the field of plant–cyanobacteria symbiosis, which were driven by the generation of high-throughput-data. We pinpoint their power to yield general patterns across these diverse symbioses occurring in representative plant species from across all major lineages.

Yours sincerely

Sophie de Vries and Jan de Vries

Review: Evolutionary genomic insights into cyanobacterial symbioses in plants — R0/PR2

Conflict of interest statement

Reviewer declares none.

Comments

Comments to Author: This is a thorough review about symbioses of terrestrial plants with cyanobacteria. I have mostly minor comments, except for one serious complaint:

Lines 329/330 and Lines 504-510: you refer to Ekman et al. 2008 and Ekman et al. 2013 (which, by the way, are missing in your References section) about the function of proteins encoded by genes induced during symbiosis. At the time, the function of these proteins was assessed based on sequence similarity. Meanwhile, more protein functions have been analysed, see specifically Qiu et al. 2020 https://pubmed.ncbi.nlm.nih.gov/33196124/. Please check whether the definitions of protein function can be upheld, and adapt your discussion accordingly.

Line 58: ”(Reinke 1872).” – I like the historical depth, but one more reference that is (more easily) accessible would be appreciated.

Line 60: “(Towata 1985).” – and here, ditto.

Line 71: replace “emerge” with “emerging”

Line 75: “It sets the seed…” – should be “They”, not “It”, and “sets the seed” should be replaced by a more common phrase… lies the foundations, maybe?

Line 92: “the symbiotic communities and its features” – should be “their features”

Line 96: “The question on what cyanobacteria colonize…” – should be “The question which cyanobacteria colonize…”

Line 99: “…section IV/V cyanobacteria, foremost that of the order of Nostocales,…” – should be “foremost those of the order Nostocales”…

Line 114: Nelson et al. 2020 (and Nelson et al. 2019) is missing in the References section

Lines 117/118: “suggesting that no particular specialization is needed to colonize those canals” – This is a bit vague. The trait enabling the colonization of the canals may not be uncommon in some habitats, but it’s still a specialization.

Line 123: “(Nelson et al. 20210)” – missing in the References section (even after correction of the year)

Line 134: “the genome of T. azollae is obligate “ – replace "genome" with "symbiosis"

Lines 144/145: “…Nakayama et al., 2014)” – missing in the References section

Lines 152/153: “The first genome of a cyanobacterial symbiont of plants was sequenced by JGI and was an isolate of N. punctiforme from a cycad” – change of subject; please rephrase

Line 153, line 155: “(Meeks et akl. 2001)” – please correct

Lines 160/161: “we will set the findings from the different cyanobacterial genomes in comparison.” – “we will compare…” would be preferable

Lines 168/169: “While those Nostoc isolates that can associate ntracellular with Gunnera do form a monophyletic clade, extracellular and epiphytically” – intracellularly, extracellularly

Lines 187/188: “One nitrogenase requires molybdenum (Mo) as a co-factor and is expressed only in heterocysts” – It also requires iron.

Lines 197/198: “an additional vanadium-based nitrogenase (V-nitrogenase; Thiel and Pratte 2001, Nelson et al. 2019, Pratte and Thiel 2021).” – It also needs iron.

Lines 210/211: “As to if and when V-nitrogenase can play a role under Mo-limited conditions during symbioses needs to be further investigated.” – I recommend Darnajoux et al. 2019 https://www.pnas.org/content/116/49/24682

Lines 216/217: “These data show that proteins related to nitrogen metabolism were enhanced while those related to photosynthesis were reduced.” – you mean, in symbiosis?

Lines 221/222: “yet the expression of genes involved in nitrogen fixation or heterocysts formation are not induced…” – should be “…or heterocyst formation is not induced…”

Lines 223/224: ”This suggests a continuous rate of nitrogen fixation in contrast to an increased one in the tighter associations.” – better explain “continuous”, e.g., “a rate of nitrogen fixation similar to that observed in the free-living state”

Lines 227-229: “and an enzyme required for the iron-Mo co-factor assembly (nifB; Curatti et al. 2006—and the regulator for heterocysts formation, hetR,...” – It’s the FeMo cofactor. Close the parentheses. Heterocyst formation, not heterocysts formation

Line 235: “nifH sequences” – nifH in italics, please

Line 256: “These include, among other genes a large set of transporters…” – are you talking about genes of proteins?

Line 264: “as the putative permanent and inherited cyanobiont of Azolla” – why not use “vertically transferred” instead of “inherited”?

Lines 270/271: “(i) Warshan et al. (2018) used one particular cyanobacteria as non-symbiotic control,” – should be “cyanobacterium”. The singular is “cyanobacterium”. Or write “cyanobacterial strain”.

Line 277: “For one as its current name indicates T. azollae that it does not belong to the genus” – comma after “For one” and “indicates”, delete “that is”

Lines 292/293: “Genes involved in chemotaxis of hormogonia (CheA, B, D, R and W)” – do not capitalize cheA

Lines 294/295: “The absence of the functional CheR Methyltransferase” – do not capitalize methyltransferase

Lines 301/302: “to the sporocarps or a new leave cavity.” – leaf

Lines 311/312: “in case of Azolla—are colonized hormogonia revert back to” – comma after ”colonized”

Lines 329/330: “proteomics of the cyanobiont of Azolla indicated an increase of a fructose transporter (Ekman et al. 2008).” – The authors did not determine the protein function; they used sequence homology. I checked the protein sequence and found that nowadays, it shows the strongest homology to “iron uptake porin”.

Besides, the Ekman et al. papers (2008, 2013) are missing in the References section

Lines 336/337: “(ii) that a lack of fructose during an established interaction may be circumvented,” – delete ”that” for syntax reasons, and delete the entire point because the fructose transport function seems to be dead. Besides, the specificity of a sugar transporter for a _single_ monosaccharide should always be confirmed experimentally; sequence homology is not sufficient here.

Lines 347-350: “The gene families involved in transport and metabolism of organic sulfur, which are retained by the facultative symbionts, are lost from the reduced genome of T. azollae, the cyanobiont of Azolla (Ran et al. 2010; Warshan et al. 2018).” – you mentioned that already.

Lines 373-376: “Overall, these data suggest that a vital part of establishing and possibly maintaining nitrogen fixing symbioses—or the constitutive nitrogen fixing state in such symbioses—in many host-cyanobiont interactions relies on the induction of protection against reactive oxygen species.” – you could make a reference to legume nodules here – high respiration rates under microaerobic conditions, enabled by leghemoglobin, lead to increased ROS production during respiration, Günter et al. 2007 https://pubmed.ncbi.nlm.nih.gov/17990967/

Lines 384/385: ”That symbiotic cyanobacteria are surrounded by various bacteria has been noted for every system.” – Seriously? What about Gunnera symbioses?

Lines 392/393: “…to the more ‘closed’ options such as coralloid roots of cycads or the glands of Gunnera species.” – I still do not see any reference for non-cyanobacteria being found in infected cells of Gunnera glands. It’s not that I do not believe the possibility – there are non-rhizobia in infected cells of legume nodules, after all – but I want proof.

Lines 409-412: “such microbes have been observed in the same symbiotic spaces as the cyanobiont; i.e. in those specific gland cells that harbour the symbiotic Nostoc filaments. Yet, these are only microscopic observations-based data and no molecular insights have been garnered.” – I did not see any bacteria or fungi in the infected cells in Towata 1985. They were in the mucilage or cell walls - based on the figures and the figure legends.

Lines 477-479: “that is the formation of motile filaments, which are further specifically attracted to their hosts (Nilsson et al. 2006).” – The only new information about hormogonia here is that they are filaments. I think it would be better to have the full introduction at first occurrence in the text.

Lines 504-507: “Functional analyses of the genome of the symbiotic N. punctiforme has shown that it encodes a functional fructose transporter, a glucose permease as well as a carbohydrate porin which can transport both sugars (Ekman et al. 2013).” – the carbohydrate porin is now annotated as an iron porin. Please check the function of the proteins.

Lines 509-510: “Yet only mutants unable to produce the glucose permease resulted in the abortion of infection of the hornwort A. punctatus (Ekman et al. 2013).” – please check whether the glucose permease definition can be upheld.

Lines 557/558: “…That does not mean that they play no role in symbiosis in general. It however shows...” – There could be other sugar transporters involved.

Lines 558/559: “…shows that the picture cannot be transferred between the symbiotic systems in a 1:1 manner.” - you would only expect orthologous genes to be involved if the symbioses had a common evolutionary origin. Is there any reason to assume this?

Lines 559/560: ”... This may be for an instance because the relationship between A. filiculoides and its permament and obligate cyanobiont” – I do not know what you mean with “for an instance” here.

Lines 573-576: “In contrast, transcriptomic data on the cyanobacterial and host hint on the fact that every host system— in line with the convergent evolution of these cyanobacterial symbioses and the unique characters and different degrees of intimacy observed in each system—highlight unique molecular features particular to each individual association.” – sentence does not compute; please rephrase

Recommendation: Evolutionary genomic insights into cyanobacterial symbioses in plants — R0/PR3

Comments

Comments to Author: Dear Dr. de Vries

Thank you for your submission. I am sorry for the serious delay in our review process.

We have contacted 16 reviewers and only 2 agreed to review. The first reviewer returned the review on time. But the second reviewer withdrew his/her consent of review after 3 weeks overdue.

Since I do not want to keep you waiting, as it might take a very long time to look for another reviewer, after reading your manuscript and the first review, I think the review of the first reviewer is sufficient to further polish the article.

Hence, please kindly amend your manuscript according to the recommendations of this reviewer.

Sorry again and thank you very much for your patience.

Yours sincerely

Boon Leong Lim

Associate Editor

Decision: Evolutionary genomic insights into cyanobacterial symbioses in plants — R0/PR4

Comments

No accompanying comment.

Author comment: Evolutionary genomic insights into cyanobacterial symbioses in plants — R1/PR5

Comments

Dear Professor Hamant, dear Professor Lim

Thank you handling our manuscript. We are pleased that the reviewer found our work interesting and thorough.

In this revised version of our manuscript, we have made sure to tackle all of the constructive comments made by the reviewer. Please see our responses in the response letter, where we have labelled our responses with “>>>>AU:”.

Furthermore, we have added a completely new figure (Figure 4), that better captions quantitative aspects (coding content, genome sizes, pseudogenization etc.) of the sequenced diversity discussed here.

17. December: Additionally, we have now introduced the requested shortening, additional layout changes, and further integration of the figures into the text.

20. December: We have re-ordered the sections.

We remain

Yours sincerely

Sophie de Vries and Jan de Vries

Review: Evolutionary genomic insights into cyanobacterial symbioses in plants — R1/PR6

Comments

Comments to Author: All points of criticism have been answered satisfactorily.

Recommendation: Evolutionary genomic insights into cyanobacterial symbioses in plants — R1/PR7

Comments

Comments to Author: We are sorry for the delay. We invited a second reviewer and today, one month after he accepted the invitation, he just recommended acceptance without submitting a detailed report . Thank you very much for revising this manuscript to its current

Decision: Evolutionary genomic insights into cyanobacterial symbioses in plants — R1/PR8

Comments

No accompanying comment.

Author comment: Evolutionary genomic insights into cyanobacterial symbioses in plants — R2/PR9

Comments

No accompanying comment.

Recommendation: Evolutionary genomic insights into cyanobacterial symbioses in plants — R2/PR10

Comments

No accompanying comment.

Decision: Evolutionary genomic insights into cyanobacterial symbioses in plants — R2/PR11

Comments

No accompanying comment.