Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-27T08:40:35.002Z Has data issue: false hasContentIssue false

Lichen algae: the photosynthetic partners in lichen symbioses

Published online by Cambridge University Press:  13 October 2021

William B. Sanders*
Affiliation:
Department of Biological Sciences, Florida Gulf Coast University, Ft. Myers, FL33965-6565, USA
Hiroshi Masumoto
Affiliation:
Laboratory of Terrestrial Microbiology and Systematics, Graduate School of Global Environmental Studies, Kyoto University, Yoshida-Honmachi, Sakyo-ku, Kyoto606-8501, Japan
*
Author for correspondence: William B. Sanders. E-mail: wsanders@fgcu.edu

Abstract

A review of algal (including cyanobacterial) symbionts associated with lichen-forming fungi is presented. General aspects of their biology relevant to lichen symbioses are summarized. The genera of algae currently believed to include lichen symbionts are outlined; approximately 50 can be recognized at present. References reporting algal taxa in lichen symbiosis are tabulated, with emphasis on those published since the 1988 review by Tschermak-Woess, and particularly those providing molecular evidence for their identifications. This review is dedicated in honour of Austrian phycologist Elisabeth Tschermak-Woess (1917–2001), for her numerous and significant contributions to our knowledge of lichen algae (some published under the names Elisabeth Tschermak and Liesl Tschermak).

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
Copyright © The Author(s), 2021. Published by Cambridge University Press on behalf of the British Lichen Society

Introduction

The principal components of the lichen symbiosis are fungus and alga. Their intimate trophic relationship remains central to the lichen concept, despite our growing appreciation that other micro-organisms harboured within the thallus might also play significant roles (Lakatos et al. Reference Lakatos, Lange-Bertalot and Büdel2004; Grube & Berg Reference Grube and Berg2009; Bates et al. Reference Bates, Cropsey, Caporaso, Knight and Fierer2011; Grube et al. Reference Grube, Cernava, Soh, Fuchs, Aschenbrenner, Lassek, Wegner, Becher, Riedel and Sensen2015; Spribille et al. Reference Spribille, Tuovinen, Resl, Vanderpool, Wolinski, Aime, Schneider, Stabentheiner, Toome-Heller and Thor2016; Muggia & Grube Reference Muggia and Grube2018; Mark et al. Reference Mark, Laanisto, Bueno, Niinemets, Keller and Scheidegger2020; Smith et al. Reference Smith, Dal Grande, Muggia, Keuler, Divakar, Grewe, Schmitt, Lumbsch and Leavitt2020; Tzovaras et al. Reference Tzovaras, Segers, Bicker, Dal Grande, Otte, Anvar, Hankeln, Schmitt and Ebersberger2020). The lichen-forming fungi typically build distinctive vegetative tissues and characteristic sexual structures, providing numerous biological features for study and significant clues about phylogenetic relationships, which are now relatively well delimited at broader taxonomic levels (Jaklitsch et al. Reference Jaklitsch, Baral, Lücking and Lumbsch2016; Lücking et al. Reference Lücking, Hodkinson and Leavitt2017a). Lichen algae, by contrast, have proved much more elusive. Most are unicells or simple filaments, with sexual structures unknown or seldom reported. The paucity of phenotypic characters is often aggravated by their plasticity. Lichen algae may look and behave quite differently in symbiosis with different lichen-forming fungi, in the free-living condition in nature and in aposymbiotic laboratory culture (Fig. 1; Ahmadjian Reference Ahmadjian1967; Bubrick Reference Bubrick and Galun1988). All this has hindered progress in clarifying their identities, phylogenies and life histories. Schwendener (Reference Schwendener1869) was the first to survey lichen ‘gonidia’ in a phycological context, recognizing them as organisms distinct from the surrounding fungus that correspond to known taxa of free-living algae. In the last half-century, the diversity of lichen-forming algae has been reviewed by various authors (Ahmadjian Reference Ahmadjian1967; Létrouit-Galinou Reference Letrouit-Galinou1968; Henssen & Jahns Reference Henssen and Jahns1974; Friedl & Büdel Reference Friedl, Büdel and Nash2008), with a particularly thorough literature summary compiled and annotated by Tschermak-Woess (Reference Tschermak-Woess and Galun1988a).

In recent decades, our understanding of algal diversity and biosystematics has advanced substantially with the accumulation, analysis and integration of DNA sequence data. Systematic schemes for the eukaryotic algae have changed considerably, as the broad contours of consensus emerge concerning phylogenies and their reconstruction. Recent works have reviewed the current status of some principal algal groups with lichen-forming taxa, such as the genus Trebouxia (Muggia et al. Reference Muggia, Candotto-Carniel, Grube, Grube, Seckbach and Muggia2017), the class Trebouxiophyceae (Muggia et al. Reference Muggia, Leavitt and Barreno2018), the Coccomyxa-Elliptochloris clade (Gustavs et al. Reference Gustavs, Schiefelbein, Darienko, Grube, Seckbach and Muggia2017), the Trentepohliaceae (Grube et al. Reference Grube, Muggia, Baloch, Hametner, Stocker-Wörgötter, Grube, Seckbach and Muggia2017a), and the cyanobacteria (Rikkinen Reference Rikkinen, Grube, Seckbach and Muggia2017). Yet most taxa remain insufficiently understood. Even the most intensively studied genera, such as Trebouxia, are still unresolved with respect to species delimitation, and much new diversity continues to be uncovered (Muggia et al. Reference Muggia, Nelsen, Kirika, Barreno, Beck, Lindgren, Lumbsch and Leavitt2020). A great many algal symbionts, identified phenotypically (often without isolation into culture) or recorded merely as ‘trebouxioid’ or ‘chlorococcalean’, have yet to be revisited with DNA sequence analyses. Identities and relationships remain especially problematic among the cyanobacteria (blue-green algae), where sexual reproduction is absent, diversification is ancient (Garcia-Pichel Reference Garcia-Pichel and Schaechter2009) and horizontal gene transfer events may obscure the vertical components of phylogenies (Zhaxybayeva et al. Reference Zhaxybayeva, Gogrten, Charlebois, Doolittle and Papke2006). The aposymbiotic lives of lichen algae also remain largely unknown, despite their potential importance in active genetic mixing. Here an attempt is made to focus more attention on the algal side of the lichen partnership, still relatively neglected compared to that of the fungus. We include a synopsis of the relevant genera and list citations of algal taxa in lichen symbiosis (Table 1), emphasizing those published since Tschermak-Woess's (Reference Tschermak-Woess and Galun1988a) landmark review, and particularly those accompanied by genetic sequence data.

Fig. 1. Three filamentous lichen photobiont genera in aposymbiotic and symbiotic states. A–C, Trentepohlia. A, branching filament free-living on bark. B, lichenized by Coenogonium hyphae (arrows) growing over morphologically unchanged algal filament and its new branches (horizontal arrow). C, lichenized by Arthonia rubrocincta; the alga is largely broken up into individual cells or short segments. D–F, Rhizonema. D, cultured isolate from Dictyonema; note false branching (arrowhead). E, trichome ensheathed by cells of mycobiont Dictyonema. F, contorted or broken filaments (arrow) within thallus of Coccocarpia palmicola. G–J. Nostoc. G, free-living thallus-like macrocolony on soil. H, cultured strain. I, more or less intact filaments (arrows) within thallus of Collema furfuraceum. J, contorted or broken up into cell groups (arrows) within cyanomorph of Sticta canariensis. Scales: A–F, H–J = 10 μm; G = 1 cm.

Table 1. Taxonomically grouped list of photobiont genera and mycobionts reported in association with them. The family names of the mycobionts are included in places where emphasis might be useful. id = procedures used in the study to identify the photobiont. LM = light microscopy, TEM = transmission electron microscopy. See table 1 in Tschermak-Woess (Reference Tschermak-Woess and Galun1988a) for a comprehensive list of photobiont reports prior to 1988. Taxon names follow those used in the original articles.

The Major Algal Groups Involved

Lichen algae are diverse. This may contribute to the distinct distributions and climatic preferences of the symbiotic thalli that enclose them (Marini et al. Reference Marini, Nascimbene and Nimis2011). Most are green algae, a paraphyletic grouping of two major clades: the charophytes (Streptophyta), from which embryophytes descend, and the Chlorophyta s. str. (Leliaert et al. Reference Leliaert, Smith, Moreau, Herron, Verbruggen, Delwiche and De Clerck2012). The latter includes nearly all green algae reported as lichen symbionts. Within the Chlorophyta, lichen symbionts are found principally in the classes Trebouxiophyceae and Ulvophyceae. A third class, the Chlorophyceae, is known or suspected to include the partners of several lichens. The prokaryotic blue-green algae (cyanobacteria) encompass most of the remainder, occurring in c. 10% of the nearly 20 000 known lichen associations (Rikkinen Reference Rikkinen, Grube, Seckbach and Muggia2017). Additionally, two stramenopile algae (a xanthophyte and a phaeophyte) are known to enter into lichen symbioses. The full range of phylogenetic disparity among lichen-forming algae is therefore much wider than that found among the lichen-forming fungi, which all fall within the kingdom's Dikarya crown group (mostly Ascomycota, with several genera of Basidiomycota). Just what common features might permit those disparate algal lineages to form comparable symbioses with lichen-forming fungi remain enigmatic. As colonizers of exposed, subaerial substrata, potentially suitable algae may be pre-adapted to coping with hydric stresses and high radiation loads (Lange et al. Reference Lange, Pfanz, Kilian and Meyer1990; Gustavs et al. Reference Gustavs, Eggert, Michalik and Karsten2010; Candotto Carniel et al. Reference Candotto, Zanelli, Bertuzzi and Tretiach2015). It is striking that most lineages of basidiomycete fungi that independently adopted the lichen lifestyle did not domesticate novel algal genera; instead they chose taxa that associate with ascolichens, such as Coccomyxa, Elliptochloris and Rhizonema (Oberwinkler Reference Oberwinkler and Hock2012; Dal Forno et al. Reference Dal Forno, Lawrey, Sikaroodi, Gillevet, Schuettpelz and Lücking2020; Masumoto Reference Masumoto2020; but see Hodkinson et al. (Reference Hodkinson, Moncada and Lücking2014) concerning Lepidostromatales). It is also noteworthy that quite a number of lichen algae belong to genera (e.g. Chlorella s. str., Coccomyxa, Elliptochloris and Nostoc) that include species occurring in symbiosis (often endosymbioses) with diverse protists, plants and animals (Adams et al. Reference Adams, Duggan, Jackson and Whitton2012; Grube et al. Reference Grube, Seckbach and Muggia2017b).

Algal partners in lichen symbioses were termed phycobionts by Scott (Reference Scott1957). Subsequently, Ahmadjian (Reference Ahmadjian1982) proposed that photobiont replace phycobiont where cyanobacteria are meant to be included, because they ‘are not algae per se but actually bacteria’. No further argumentation was provided; it was presumed self-evident that algae and bacteria must denote mutually exclusive concepts. Some contemporary treatments distinguish cyanobacteria from algae (e.g. Friedl & Büdel Reference Friedl, Büdel and Nash2008; Grube et al. Reference Grube, Seckbach and Muggia2017b), while others consider them as algae (e.g. Graham et al. Reference Graham, Graham and Wilcox2009; Büdel & Kauff Reference Büdel, Kauff and Frey2012; Lee Reference Lee2018). Clearly, there are significant differences between prokaryotes and eukaryotes. At issue, however, is whether those differences are relevant to the concept of algae. This term has no biosystematic status and cannot attain any by exclusion of the blue-greens. The emblematic algal trait, oxygen-generating photosynthesis, is ultimately derived from cyanobacteria. It was subsequently acquired in multiple events involving primary, secondary and tertiary endosymbioses (Keeling Reference Keeling2004, Reference Keeling2013), and now characterizes diverse lineages included within most of the major eukaryote clades (Archaeplastida, Alveolata, Excavata, Rhizaria, Stramenopila, Cryptista and Haptista). The one and only unifying thread in this polyphyletic algal tapestry (Delwiche Reference Delwiche1999) is the common photosynthetic apparatus, originating in cyanobacteria and passed on vertically as well as horizontally. The present work therefore uses the term algae to encompass all non-embryophyte lineages that inherited oxygenic photosynthesis. Phycobiont and photobiont are considered synonymous terms.

The Algal Role in Lichen Symbiosis

The algal partner is the primary producer, sustaining the lichen association by supplying the fungal partner with carbohydrate products of photosynthesis (Smith Reference Smith1974). Those with pyrenoids (Fig. 2) possess CO2-concentrating mechanisms that improve the efficiency of carbon fixation (Smith & Griffiths Reference Smith and Griffiths1996). Green algal symbionts (chlorobionts) transfer their photosynthate as polyol sugar alcohols such as ribitol (Richardson et al. Reference Richardson, Hill and Smith1968). Significantly, these compounds also confer desiccation tolerance by providing osmolarity and protecting cell membranes from damage as water is lost (Smith Reference Smith2019). Polyols are likewise produced by non-symbiotic, aeroterrestrial green algae, particularly under osmotic stress conditions (Darienko et al. Reference Darienko, Gustavs, Mudimu, Menendez, Schumann, Karsten, Friedl and Pröschold2010; Gustavs et al. Reference Gustavs, Eggert, Michalik and Karsten2010, Reference Gustavs, Görs and Karsten2011). Blue-green symbionts (cyanobionts) transfer glucose, or glucan, which their fungal partners take up and immediately convert into the sugar alcohol mannitol (Smith & Drew Reference Smith and Drew1965; Hill Reference Hill1972). When lichenized, the algal symbionts are somehow induced to leak large amounts of carbohydrate to the surrounding fungal cells, a process that quickly ceases when the algae are isolated into culture (Drew & Smith Reference Drew and Smith1967). Fungal penetration of photobionts may occur to varying degrees (Geitler Reference Geitler1934; Tschermak Reference Tschermak1941a; Plessl Reference Plessl1963; Galun et al. Reference Galun, Paran and Ben-Shaul1970, Reference Gallun [sic], Ben-Shaul and Paran1971; Honegger Reference Honegger1986; Matthews et al. Reference Matthews, Tucker and Chapman1989), but these so-called haustoria do not appear to be principal conduits of carbohydrate transfer in ascolichens (Jacobs & Ahmadjian Reference Jacobs and Ahmadjian1971; Collins & Farrar Reference Collins and Farrar1978; Hessler & Peveling Reference Hessler and Peveling1978). The intrusive hyphae of certain basidiolichens that deeply penetrate longitudinally through the centre of their cyanobiont trichomes (Roskin Reference Roskin1970; Oberwinkler Reference Oberwinkler, Schwemmler and Schenk1980, Reference Oberwinkler and Hock2012) have not yet been examined with respect to substance transfer. In most foliose and fruticose lichens examined, haustorial penetrations are either absent altogether or do not fully traverse the algal cell wall. To facilitate transfer, the mycobiont secretes a hydrophobic sealant that envelops the cell surfaces of both symbionts at their contact zones, thereby funnelling carbohydrate released by the alga to the fungus (Honegger Reference Honegger1991; Trembley et al. Reference Trembley, Ringli and Honegger2002a). At least that is the case in the selection of taxa examined so far. Where cyanobacterial symbionts are involved, they provide the lichen fungus with fixed nitrogen as well as carbon (Millbank & Kershaw Reference Millbank, Kershaw, Ahmadjian and Hale1974). In those lichens (chiefly Peltigerales) where a chlorobiont constitutes the main algal layer and cyanobionts are localized within nodules known as cephalodia, the cyanobacteria become highly specialized for nitrogen fixation, with an elevated percentage of cells differentiating as heterocytes (Hitch & Millbank Reference Hitch and Millbank1975). In lichens with only cyanobacterial photobionts, heterocyte frequency can be much lower at the growing margins of the thallus (Bergman & Hällbom Reference Bergman and Hällbom1981), where photosynthate may be in higher demand.

Fig. 2. TEM micrographs of some photobiont pyrenoids, with plastoglobuli (round black dots) and penetrating membranes in various positions and orientations. A, Trebouxia, within thallus of Lasallia pustulata. Note pyrenoid structure here more closely resembles that of distantly related Heveochlorella (B) than that of another species (C) of Trebouxia. B, Heveochlorella, within thallus of Calopadia. C, Trebouxia, within thallus of Ramalina usnea. D, bulging exserted pyrenoid of Petroderma maculiforme. E, Diplosphaera, within thallus of Endocarpon pusillum. S = starch grain or plates. Scales: A = 1 μm; B = 200 nm; C–E = 500 nm.

Whether any substance is transferred from fungus to alga in exchange has yet to be demonstrated. At least some genes relevant to such metabolic transfers appear to be differentially expressed in symbiosis (Kono et al. Reference Kono, Kon, Ohmura, Satta and Terai2020). Certainly, there has been speculation that the fungal partner might apportion carbohydrate, nitrogen, or other substances back to the algal symbiont to regulate its growth (Ahmadjian Reference Ahmadjian1995) in coordination with that of the mycobiont (Greenhalgh & Anglesea Reference Greenhalgh and Anglesea1979; Hill Reference Hill and Brown1985, Reference Hill1989; Honegger Reference Honegger1987). The heterotrophic tendencies shown by many lichen algae (Trebouxia, Asterochloris, Elliptochloris, Coccomyxa, Apatococcus) when cultured in the laboratory (Ahmadjian Reference Ahmadjian1993; Gustavs et al. Reference Gustavs, Schumann and Karsten2016, Reference Gustavs, Schiefelbein, Darienko, Grube, Seckbach and Muggia2017) suggest the possibility that they could be susceptible to such control. Indeed, Ahmadjian (Reference Ahmadjian and Seckbach2001) proposed that Trebouxia is fully dependent upon its mycobiont for nutrition and is therefore unable to survive in the free-living state (Ahmadjian Reference Ahmadjian1988). However, he also promoted the seemingly contradictory viewpoint that Trebouxia is a victim of fungal parasitism rather than a mutualist partner (Ahmadjian Reference Ahmadjian1993, Reference Ahmadjian1995, Reference Ahmadjian2002). This would make Trebouxia a host that cannot survive without its parasite.

In any event, proof of fungus-to-alga nutrient transfer is not required to make the case that lichen symbiosis offers advantages to the algal partner. There is considerable evidence that the surrounding fungal tissues and their secondary metabolites may help protect the lichenized alga from desiccation, photoinhibition, temperature extremes and herbivory (e.g. Solhaug & Gauslaa Reference Solhaug and Gauslaa1996; Kranner et al. Reference Kranner, Beckett, Hochman and Nash2008; Kosugi et al. Reference Kosugi, Arita, Shizuma, Moryama, Kashino, Koike and Satoh2009; Asplund & Wardle Reference Asplund and Wardle2013; Gauslaa et al. Reference Gauslaa, Alam, Lucas, Chowdhury and Solhaug2017; Míguez et al. Reference Míguez, Schiefelbein, Karsten, García-Plazaola and Gustavs2017; Sadowsky & Ott Reference Sadowsky and Ott2016; Beckett et al. Reference Beckett, Solhaug, Gauslaa and Minibayeva2019; Fernández-Marín et al. Reference Fernández-Marín, López-Pozo, Perera-Castro, Irati Arzac, Sáenz-Ceniceros, Colesie, de los Ríos, Sancho, Pintado and Laza2019). Symbiosis may significantly improve the alga's ability to avoid cellular damage caused by highly reactive forms of oxygen (ROS) generated under stress conditions (Kranner et al. Reference Kranner, Cram, Zorn, Wornik, Yoshimura, Stabentheiner and Pfeifhofer2005). With these protections, and the facilitated display for light capture afforded by a supportive mycobiont structure, lichen algae may greatly expand their ecological range and abundance via symbiosis (Honegger Reference Honegger and Hock2012). On the other hand, lichen symbioses are diverse and it is likely that the parameters of the relationship vary among taxa, along environmental gradients, and perhaps also during the course of a single lichen's development. The long history of attempts to maintain or resynthesize lichens in the laboratory has provided a key insight into the nature of this seemingly well-integrated association: it is very much a relationship of contingency. That the partners can often be cultured separately on appropriate media in the laboratory (Ahmadjian Reference Ahmadjian1993; Crittenden et al. Reference Crittenden, David, Hawksworth and Campbell1995; Stocker-Wörgötter & Hager Reference Stocker-Wörgötter, Hager and Nash2008) shows there is no strict physiological impediment to growth without symbiosis. To initiate and support lichen formation, a fluctuating balance of conditions suboptimal for separate fungal or algal growth appears to be necessary. Any combination of culture conditions (light, moisture, nutrient availability) that continuously favours either fungal or algal growth results in the breakdown of symbiotic structures, and the dissociated proliferation of the micro-organisms separately (Thomas Reference Thomas1939; Scott Reference Scott1960; Ahmadjian Reference Ahmadjian1962; Stocker-Wörgötter Reference Stocker-Wörgötter2001; but see Marton & Galun Reference Marton and Galun1976). It therefore seems logical to view the lichen symbiosis as a more or less mutualistic response to conditions that permit neither partner to thrive independently.

Although both partners may derive benefits, the lichen symbiosis is clearly not symmetrical (Hill Reference Hill2009). The heterotrophic mycobionts, with their elaborate structural adaptations for algal cultivation, are more fully committed to symbiosis than their trophically autonomous photobionts. The mycobiont frees itself of symbiosis only in spore dispersal, seeking algal partners again immediately upon germination. To carry out sexual reproduction, it must be in symbiosis, whereas its photobiont appears to need aposymbiotic freedom to do so. From the alga's point of view, whenever unfavourable conditions reduce its possibilities of aposymbiotic success, the benefits of lichenization may begin to outweigh any disadvantages. Photobionts may rely on lichen symbioses for long-term persistence in habitats periodically subject to adverse conditions, while needing intervals of independence under favourable conditions to complete their life cycles. Thus, mycobiont and photobiont life histories do not fully coincide, but produce a lichen where they intersect compatibly. To varying degrees, natural selection has optimized the mycobiont principally for symbiosis, the photobiont for autonomy as well as symbiosis. The trade-off is that greater adaptation to symbiotic compatibility is likely to constrain the possibilities for competitive success in the aposymbiotic state. However, the lingering notion that certain photobionts may not ever occur free-living is probably attributable to insufficient sampling, and the conflation of invisibility with absence. Unsurprisingly, those photobionts that are macroscopically visible (Nostoc, Cephaleuros, Phycopeltis, Trentepohlia, Prasiola, Petroderma) have not had their aposymbiotic occurrence disputed.

Both fungus and alga must adapt, at least to some extent, to be compatible symbionts. For some authors, such mutual adaptation constitutes coevolution (Ahmadjian Reference Ahmadjian1987; Saini et al. Reference Saini, Nayaka, Bast, Satyanarayana, Johri and Das2019); for others, coevolution supposes parallel cladogenesis in partners’ phylogenies, a criterion not generally met by lichen symbioses analyzed in this regard (Piercey-Normore & DePriest Reference Piercey-Normore and DePriest2001; Stenroos et al. Reference Stenroos, Högnabba, Myllys, Hyvönen and Thell2006). However, it has been argued that focusing exclusively on this fine scale ignores broader patterns of co-adaptation, whereby ‘guilds’ of different mycobionts share common pools of photobionts to mutual advantage (Rikkinen Reference Rikkinen2003, Reference Rikkinen2013). According to Hill (Reference Hill2009), photobionts cannot coevolve with their mycobionts because they lack sexual reproduction in the thallus, are not subject to natural selection from one lichen to the next, and are not perpetuated when a lichen thallus dies. Yet photobionts are continually escaping from lichen thalli by means of soredia, isidia, thallus fragments, co-dispersed hymenial, epithecial or conidiomal algae (Fig. 3), and the excreta of lichenivorous invertebrates (Fröberg et al. Reference Fröberg, Björn, Baur and Baur2001; Meier et al. Reference Meier, Scherrer and Honegger2002; Boch et al. Reference Boch, Prati, Werth, Rüetschi and Fischer2011). Such diaspores afford many chances of finding microconditions where independent algal growth is favoured; aposymbiotic, potentially sexual populations may then develop, be they brief or enduring. Selection among genotypes for compatibility (or resistance) will occur when the opportunity for relichenization next presents itself. Compatible genotypes incorporated into a developing lichen may then be subject to further winnowing selection in the course of thallus growth.

Fig. 3. Liberation and potential co-dispersal of photobionts from the spore-producing structures of certain mycobionts. A, Diplosphaera photobiont (arrows) within perithecium of Endocarpon pusillum; note much smaller size compared to photobiont cells within thalline tissue (t); s = ascospore. B, apothecial surface of foliicolous lichen colonizing plastic cover slip; note epithecial algal cells (arrows) among emerging ascospores (s). C, Heveochlorella photobionts (vertical arrow) within conidiogenous tissue of campylidia and intermixed among filiform macroconidia (oblique arrow). D, hyphophore of Gyalectidium paolae showing diahyphal propagules (bundles of conidial chains dispersed as a unit) with adhering or intermixed Heveochlorella photobionts (arrows). E, campylidial macroconidia, with co-dispersed Heveochlorella photobionts loosely encircled, germinating (arrowheads) on a plastic cover slip. F, diahyphal propagules of Gyalectidium germinating (arrowheads) on a plastic cover slip, with co-dispersed Heveochlorella photobionts. Scales: A, C & D = 20 μm; B = 50 μm; E & F = 10 μm.

Patterns of Symbiont Pairing

The asymmetrical needs of the lichen symbionts are reflected in the non-reciprocal patterns of pairing that have evolved between mycobionts and photobionts. Photobiont genera frequently associate with multiple, phylogenetically disparate lineages of lichen-forming fungi. The converse, however, is much less common; mycobiont genera, and often families and even orders, generally tend to lichenize a single algal genus (Rambold et al. Reference Rambold, Friedl and Beck1998; Peršoh et al. Reference Peršoh, Beck and Rambold2004). There are a number of notable exceptions. Lichen-forming fungi of the family Verrucariaceae partner with an extremely diverse array of eukaryotic algae, including the only reported cases of stramenopile phycobionts (Thüs et al. Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011). The pin-lichen genus Chaenotheca (Coniocybomycetes) includes species associating with Trebouxia, Trentepohlia, Symbiochloris or Tritostichococcus (Tibell Reference Tibell2001; Škaloud et al. Reference Škaloud, Friedl, Hallmann, Beck and Dal Grande2016; Pröschold & Darienko Reference Pröschold and Darienko2020). The fruticose lichen genus Stereocaulon may harbour thallus photobionts of either Asterochloris, Vulcanochloris or Chloroidium (Vančurová et al. Reference Vančurová, Muggia, Peksa, Řídká and Škaloud2018). Species of Sticta may partner with chlorobionts of Symbiochloris, Coccomyxa, Elliptochloris, Heveochlorella or Chloroidium (Lindgren et al. Reference Lindgren, Moncada, Lücking, Magain, Simon, Goffinet, Sérusiaux, Nelsen, Mercado-Díaz and Widhelm2020). Squamulose Psora decipiens is reported to partner with either Asterochloris, Trebouxia, Chloroidium (Ruprecht et al. Reference Ruprecht, Brunauer and Türk2014) or Myrmecia photobionts (Williams et al. Reference Williams, Colesie, Ullmann, Westberg, Wedin and Büdel2017; Moya et al. Reference Moya, Chiva, Molins, Jadrná, Škaloud, Peksa and Barreno2018). In addition, it is well known that many individual mycobionts, particularly in the Peltigerales, may associate with both green and blue-green algae simultaneously, giving rise to cyanobacterial cephalodia within or upon a chlorophyte-containing thallus, or distinct cyanomorph and chloromorph thalli separately or conjoined (Fig. 4) via a common fungal individual (e.g. James & Henssen Reference James, Henssen, Brown, Hawksworth and Bailey1976). Association with both a chlorobiont and a cyanobiont in separate thallus components has also been reported for certain basidiolichen species in Cyphellostereum (Oberwinkler Reference Oberwinkler and Hock2012) and Lichenomphalia (Gasulla et al. Reference Gasulla, Barrasa, Casano and del Campo2020). In a small number of lichens, green and blue-green photobionts are known to occur intermixed within the same thallus structure (Büdel & Henssen Reference Büdel and Henssen1987; Henskens et al. Reference Henskens, Green and Wilkins2012). There are distinct physiological advantages to each of these two kinds of photobionts. Cyanobionts can fix nitrogen as well as carbon but require liquid water to rehydrate and resume physiological activity, whereas chlorobionts can rehydrate from vapour, although their CO2 fixation rates may be more adversely affected by high thallus water contents (Lange et al. Reference Lange, Kilian and Ziegler1986, Reference Lange, Büdel, Meyer and Kilian1993; Green et al. Reference Green, Büdel, Heber, Meyer, Zellner and Lange1993, Reference Green, Schlensog, Sancho, Winkler, Broom and Schroeter2002). Less obvious are the implications of choosing Trentepohlia (Ulvophyceae) versus Trebouxia (Trebouxiophyceae) photobionts; neither fix nitrogen, although they may differ in their tolerance of freezing temperatures (Nash et al. Reference Nash, Kappen, Lösch, Larson and Matthes-Sears1987). Interestingly, mycobiont genera Ionaspis and Hymenelia (Lecanoromycetes) include trentepohliophilic and trebouxiophilic taxa, and the single species H. epulotica can apparently associate with photobionts of either of these two very different genera (Lutzoni & Brodo Reference Lutzoni and Brodo1995; McCune et al. Reference McCune, Arup, Breuss, Di Meglio, Di Meglio, Esslinger, Magain, Miadlikowska, Miller and Muggia2018). Recently, Ertz et al. (Reference Ertz, Guzow-Krzemińska, Thor, Łubek and Kukwa2018) demonstrated that the lichen fungus Lecanographa amylacea can form morphologically distinct sexual and asexual thalli with Trentepohlia and Trebouxia photobionts, respectively. While the above examples show that significant divergences in photobiont selection have arisen in a number of mycobiont lineages, far more conservative tendencies appear to predominate in the majority of lichen-forming fungal groups.

Fig. 4. Dichotomously lobed chloromorphs of Sticta canariensis emerging from lower surfaces of cyanomorph thalli (arrows). Scale = 5 mm.

Photobiont choice and the range of compatible pairings for a given mycobiont were first explored experimentally in classic laboratory resynthesis studies using Cladonia cristatella and Lecanora chrysoleuca (Ahmadjian et al. Reference Ahmadjian, Russell and Hildreth1980; Ahmadjian & Jacobs Reference Ahmadjian and Jacobs1981). Varying degrees of compatibility were observed, with thallus formation reaching different developmental stages depending on the photobiont strain introduced. Nonetheless, overall results generally reflected patterns observable in natural lichens: Cladonia successfully lichenized strains of Asterochloris but not those of Trebouxia (as currently defined), while Lecanora did just the opposite. In the last two decades, genetic markers have been used to characterize the range of photobiont diversity chosen by individual lichen-forming fungal species in nature, and to assess the parameters that might affect their choices. This complex topic has attracted much attention and merits a review of its own, but some general findings can be summarized here. Most mycobiont species appear to be fairly selective; they tend to partner with a limited range of strains or species within a single photobiont genus, but to differing degrees. Some mycobionts accept a substantially broader range of taxa within the photobiont partner genus; this relative liberality is often characteristic of lichen-forming fungi that have attained wider, more cosmopolitan distributions (Blaha et al. Reference Blaha, Baloch and Grube2006; Guzow-Krzemińska Reference Guzow-Krzemińska2006; Leavitt et al. Reference Leavitt, Nelsen, Lumbsch, Johnson and St Clair2013; Muggia et al. Reference Muggia, Pérez-Ortega, Kopun, Zellnig and Grube2014; Magain et al. Reference Magain, Miadlikowska, Goffinet, Sérusiaux and Lutzoni2017; Vančurová et al. Reference Vančurová, Muggia, Peksa, Řídká and Škaloud2018), or those capable of colonizing extreme environments with probably fewer photobiont options available (Romeike et al. Reference Romeike, Friedl, Helm and Ott2002; Wirtz et al. Reference Wirtz, Lumbsch, Green, Türk, Pintado, Sancho and Schroeter2003; Engelen et al. Reference Engelen, Convey and Ott2010; Pérez-Ortega et al. Reference Pérez-Ortega, Ortiz-Álvarez, Green and de los Ríos2012; Osyczka et al. Reference Osyczka, Lenart-Borón, Borón and Rola2021; Rola et al. Reference Rola, Lenart-Borón, Borón and Osyczka2021). Such mycobionts may be closely related to species that accept a much narrower range of photobiont partners (Piercey-Normore Reference Piercey-Normore2004; Yahr et al. Reference Yahr, Vilgalys and DePriest2004; Otálora et al. Reference Otálora, Martínez, O'Brien, Molina, Aragón and Lutzoni2010; Onuţ-Brännström et al. Reference Onuţ-Brännström, Tibell and Johannesson2017). Some studies have correlated symbiont selection patterns with environmental parameters, such as latitude (Singh et al. Reference Singh, Dal Grande, Divakar, Otte, Crespo and Schmitt2017), climate (Řidká et al. Reference Řidká, Peksa, Rai, Upreti, Škaloud, Rai and Upreti2014) and ecological conditions that influence the distribution and availability of photobionts (Yahr et al. Reference Yahr, Vilgalys and DePriest2006; Fernández-Mendoza et al. Reference Fernández-Mendoza, Domaschke, García, Jordan, Martín and Printzen2011; Peksa & Škaloud Reference Peksa and Škaloud2011; Vargas Castillo & Beck Reference Vargas Castillo and Beck2012; Werth & Sork Reference Werth and Sork2014). Photobiont tolerance of heavy metals appears to influence their selection by mycobionts in some lichen communities colonizing metal-rich substrata (Vančurová et al. Reference Vančurová, Muggia, Peksa, Řídká and Škaloud2018; Rola et al. Reference Rola, Lenart-Borón, Borón and Osyczka2021) but not others (Beck Reference Beck2002; Hauck et al. Reference Hauck, Helms and Friedl2007; Bačkor et al. Reference Bačkor, Peksa, Škaloud and Bačkorová2010). Many studies stress the intrinsic compatibility requirements of individual fungal taxa as primary determinants of pairing patterns (Yahr et al. Reference Yahr, Vilgalys and DePriest2004; Stenroos et al. Reference Stenroos, Högnabba, Myllys, Hyvönen and Thell2006; Myllys et al. Reference Myllys, Stenroos, Thell and Kuusinen2007; Leavitt et al. Reference Leavitt, Kraichak, Vondrak, Nelsen, Altermann, Divakar, Alors, Esslinger, Crespo and Lumbsch2015; Joneson & O'Brien Reference Joneson and O'Brien2017), often in conjunction with ecological factors (Elvebakk et al. Reference Elvebakk, Papaefthimiou, Robertsen and Liaimer2008; O'Brien et al. Reference O'Brien, Miadlikowsa and Lutzoni2013; Dal Grande et al. Reference Dal Grande, Rolshausen, Divakar, Crespo, Otte, Schleuning and Schmitt2018; Jüriado et al. Reference Jüriado, Kaasalainen, Jylhä and Rikkinen2019; Pino-Bodas & Stenroos Reference Pino-Bodas and Stenroos2020). In some communities, mycobionts may have adapted to utilize a common pool or pools of photobionts, whose local availability might thereby be sustained for all users (Beck et al. Reference Beck2002; Rikkinen et al. Reference Rikkinen, Oksanen and Lohtander2002; Rikkinen Reference Rikkinen2003; Sanders et al. Reference Sanders, Pérez-Ortega, Nelsen, Lücking and de los Ríos2016; Onuţ-Brännström et al. Reference Onuţ-Brännström, Benjamin, Scofield, Heiđmarsson, Andersson, Lindström and Johannesson2018; Cardós et al. Reference Cardós, Prieto, Jylhä, Aragón, Molina, Martínez and Rikkinen2019). Thallus growth form may also affect photobiont selection patterns. Some authors have suggested that crustose lichens may associate with a broader range of photobionts than do related foliose and fruticose taxa (Helms et al. Reference Helms, Friedl, Rambold and Mayrhofer2001), perhaps because their more extensive and intimate contact with the substratum offers more opportunity to take up additional algae in the course of development. Lichen reproductive mode can also be superimposed upon these factors. Some studies have found that lichens reproducing primarily by vegetative propagules, such as soredia or isidia, associate with a narrower range of photobiont genotypes, presumably due to chiefly vertical transmission of both symbionts together (Dal Grande et al. Reference Dal Grande, Widmer, Wagner and Scheidegger2012; Werth & Scheidegger Reference Werth and Scheidegger2012; Otálora et al. Reference Otálora, Salvador, Martínez and Aragón2013; Cao et al. Reference Cao, Zhang, Liu, Hao, Tian, Zhu and Zhou2015; Hestmark et al. Reference Hestmark, Lutzoni and Miadlikowska2016; Steinová et al. Reference Steinová, Škaloud, Yahr, Bestová and Muggia2019). However, other vegetatively reproducing lichens accept a much broader range of photobionts, suggesting that the fungus does not necessarily maintain partnership with its co-dispersed photobiont throughout development (Ohmura et al. Reference Ohmura, Kawachi, Kasai, Watanabe and Takeshita2006, Reference Ohmura, Takeshita and Kawachi2019; Nelsen & Gargas Reference Nelsen and Gargas2008, Reference Nelsen and Gargas2009; Wornik & Grube Reference Wornik and Grube2010).

Acquisition of New Algal Symbionts

Acquisition of new and different photobionts, ‘photobiont switching’, has clearly been significant in the evolution of lichen relationships. However, this phrase may refer variably to events occurring at different levels of organization. A single mycobiont individual might acquire new photobionts at different times in the course of its development (Friedl Reference Friedl1987; Wedin et al. Reference Wedin, Maier, Fernández-Brime, Cronholm, Westberg and Grube2016), or at separate places along its somatic extension (Létrouit-Galinou & Asta Reference Letrouit-Galinou and Asta1994). The degree to which the newly lichenized alga may differ genetically from algal strain(s) already in possession will be limited by the innate compatibility range of that mycobiont individual. In contrast, a new fungal individual developing from a meiospore may encounter and select a photobiont strain different from the one its parental genotypes associated with. In this case, a generational change in photobiont partner could be enabled by a generational change in mycobiont genotype. At a phylogenetic level, a cladogram may provide evidence that a fungal lineage has changed its association from one photobiont to another in the course of evolution. But at a finer scale, a great many photobiont switches, perhaps back and forth, might have taken place over many generations; comparing taxa will indicate only the overall result.

New photobionts may be acquired in multiple ways. Contact and capture of free-living photobionts in nature by hyphae emerging from germinated spores (Fig. 5), once thought to be unlikely (Lamb Reference Lamb1959), has been documented in a number of studies (Ward Reference Ward1884; Werner Reference Werner1931; Bubrick et al. Reference Bubrick, Galun and Frensdorff1984; Garty & Delarea Reference Garty and Delarea1988; Scheidegger Reference Scheidegger1995; Sanders & Lücking Reference Sanders and Lücking2002; Sanders Reference Sanders2014). In theory, a single compatible algal individual might be sufficient to generate the entire population within a developing thallus. However, there appear to be many opportunities for additional photobionts to be incorporated from exterior sources. Particularly in early developmental stages, prothallic hyphae extending outward along the substratum from the lichenized portions of the organizing thallus can incorporate additional algal cells (Sanders & Lücking Reference Sanders and Lücking2002; Sanders Reference Sanders2014). Vegetative propagules, such as soredia or isidia, also begin development with the emergence and proliferation of such hyphae (Jahns et al. Reference Jahns, Mollenhauer, Jenniger and Schönborn1979; Schuster et al. Reference Schuster, Ott and Jahns1985), anchoring the structure and greatly expanding the available surfaces for potential contact with other compatible photobionts as the thallus is organized. In many crustose lichens, a prothallus remains active at the growing margins of the lichen and may continue to incorporate compatible photobionts falling upon it or encountered on the substratum (Fig. 6; see also Galløe Reference Galløe1927: p. 40, Reference Galløe1932: p. 78; Letrouit-Galinou & Asta Reference Letrouit-Galinou and Asta1994). The multitude of discrete, lichenized units that comprise the thallus of squamulose lichens probably also arise from repeated algal capture by a network of prothallic hyphae interconnecting the squamules. Certain soil- and rock-colonizing squamulose lichens produce hyphal aggregates (cords or rhizomorphs) of indeterminate growth that penetrate the substratum extensively (Poelt & Baumgärtner Reference Poelt and Baumgärtner1964; Sanders et al. Reference Sanders, Wierzchos and Ascaso1994), giving rise to new thallus squamules where compatible algal symbionts are encountered and lichenized (Wagner & Letrouit-Galinou Reference Wagner and Létrouit-Galinou1988; Sanders & Rico Reference Sanders and Rico1992; Sanders Reference Sanders1994). The structurally similar rhizinomorphs of certain umbilicate lichens also appear to have this capability (Schuster Reference Schuster1992). In some foliose and fruticose lichens, organized thallus surfaces may themselves be capable of incorporating compatible algal cells that make external contact (Bitter Reference Bitter1904). Lichens that form cephalodia and/or joined chloromorph and cyanomorph thalli clearly retain this ability (see discussion under Nostoc below). Additionally, certain lichen-forming fungi appear capable of obtaining photobionts from other lichens, upon which their spores may germinate (Hawksworth et al. Reference Hawksworth, Coppins and James1979). The host thallus is eventually destroyed as its photobionts are taken over by the invading hyphae of the aggressor, giving rise to a new lichen (Poelt Reference Poelt1958; Friedl Reference Friedl1987; Feige et al. Reference Feige, Lumbsch and Mies1993; Lücking & Grube Reference Lücking and Grube2002; Wedin et al. Reference Wedin, Maier, Fernández-Brime, Cronholm, Westberg and Grube2016). Thus, capture of free-living algae by spore germlings is clearly not the only opportunity for a mycobiont to acquire new photobionts. On the other hand, some interesting transplant experiments with Psora decipiens suggest that lichens may not always be able to switch to more favourable photobionts when needed (Williams et al. Reference Williams, Colesie, Ullmann, Westberg, Wedin and Büdel2017).

Fig. 5. Muriform ascospore (a), probably of Calopadia, germinating on a plastic cover slip placed in a south-west Florida oak hammock, and lichenizing a group of algal cells (arrow), most likely Heveochlorella. Scale = 20 μm.

Fig. 6. Phycopeltis free-living and in stages of lichenization. A, free-living. B, edge of developed Phycopeltis thallus (left) lichenized by a network of hyphae (probably foliicolous Porina sp.) that extend over substratum and capture additional young Phycopeltis germlings (arrows). Scales: A = 20 μm; B = 10 μm.

If acquisition of additional photobionts is indeed a common occurrence in the course of lichen development, lichen thalli may be expected to contain a heterogeneous photobiont population, at least at certain stages. Some authors have observed and illustrated quite different chlorobionts occurring together within single thalli (Voytsekhovich et al. Reference Voytsekhovich, Dymytrova and Nadyeina2011). Data from molecular markers have also addressed this question. Some authors found no evidence of multiple photobiont genotypes in single thalli examined (Paulsrud & Lindblad Reference Paulsrud and Lindblad1998; Beck & Koop Reference Beck and Koop2001; Singh et al. Reference Singh, Dal Grande, Divakar, Otte, Crespo and Schmitt2017; Škaloud et al. Reference Škaloud, Moya, Molins, Peksa, Santos-Guerra and Barreno2018); others found occasional occurrences (Guzow-Krzemińska Reference Guzow-Krzemińska2006; Bačkor et al. Reference Bačkor, Peksa, Škaloud and Bačkorová2010; Muggia et al. Reference Muggia, Vancurova, Škaloud, Peksa, Wedin and Grube2013; Nyati et al. Reference Nyati, Bhattacharya, Werth and Honegger2013; Řidka et al. Reference Nyati, Scherrer, Werth and Honegger2014; Onuţ-Brännström et al. Reference Onuţ-Brännström, Benjamin, Scofield, Heiđmarsson, Andersson, Lindström and Johannesson2018; Vančurová et al. Reference Vančurová, Muggia, Peksa, Řídká and Škaloud2018; Molins et al. Reference Molins, Chiva, Calatayud, Marco, García-Breijo, Reig-Armiñana, Carrasco and Moya2020), or frequent presence (Piercey-Normore Reference Piercey-Normore2006; Muggia et al. Reference Muggia, Pérez-Ortega, Kopun, Zellnig and Grube2014; Park et al. Reference Park, Kim, Elvebakk, Kim, Jeong and Hong2015; Dal Grande et al. Reference Dal Grande, Rolshausen, Divakar, Crespo, Otte, Schleuning and Schmitt2018; Osyczka et al. Reference Osyczka, Lenart-Borón, Borón and Rola2021). Intrathalline populations of Trebouxia can also vary in simple sequence DNA regions, which may result from clonal replication errors (Mansournia et al. Reference Mansournia, Wu, Matsushita and Hogetsu2012; Dal Grande et al. Reference Dal Grande, Alors, Divakar, Bálint, Crespo and Schmitt2014a). Individual thalli of Parmotrema pseudotinctorum from the Canary Islands were reported to encompass distinct lineages of Trebouxia as well as Asterochloris (Molins et al. Reference Molins, García-Breijo, Reig-Armiñana, del Campo, Casano and Barreno2013). According to Casano et al. (Reference Casano, del Campo, García-Breijo, Reig-Armiñana, Gasulla, del Hoyo, Guéra and Barreno2011), two genetically distinct strains of Trebouxia are always present together in thalli of Ramalina farinacea, and high-throughput sequencing results suggest that a number of other, minority algae might also be present in this lichen (Moya et al. Reference Moya, Molins, Martínez-Alberola, Muggia and Barreno2017). One constant challenge in assessing photobiont identities is that lichen thallus surfaces are colonized by epibiontic algae (including possible photobionts of other lichens) that are not intimate symbionts of the lichen in question, yet may figure prominently in cultures established or samples obtained from thallus fragments (Warén Reference Warén1920; Muggia et al. Reference Muggia, Vancurova, Škaloud, Peksa, Wedin and Grube2013). Confidence that sampled algae are indeed the thallus photobionts can be improved by establishing cultures from single algal cells extracted from within the thallus using a micromanipulator (Beck & Koop Reference Beck and Koop2001), although the procedure is time-consuming. Additional evidence may be sought in TEM micrographs of photobionts within the same thallus (e.g. Catalá et al. Reference Catalá, del Campo, Barreno, García-Breijo, Reig-Armiñana and Casano2016; Molins et al. Reference Molins, Moya, García-Breijo, Reig-Armiñana and Barreno2018), particularly where more than one pyrenoid type (Friedl Reference Friedl1989) is present. However, variability should first be assessed among individuals of the same genetic strain because chloroplast structure may vary from cell to cell and often looks substantially different according to the plane of ultrathin section examined. In sequencing, conventional dideoxy chain termination (Sanger) technology will reliably identify a predominant photobiont and ignore any others present in low abundance, while the procedure fails if there are secondary photobionts in sufficient abundance (c. 30%; Paul et al. Reference Paul, Otte, Schmitt and Dal Grande2018). High-throughput sequencing will detect minority photobionts but will also be more sensitive to epibiontic algae. A recent comparison of the two sequencing approaches concluded that in most lichens there is a single dominant photobiont genotype, representative of most of the thallus population (Paul et al. Reference Paul, Otte, Schmitt and Dal Grande2018).

The Genera of Lichen Algae

Approximately 50 algal genera are currently said to include lichen photobionts. Some may represent identifications that are erroneous or based on outdated circumscriptions of taxa. Others may spin off new genera as their cryptic genetic diversity is further elucidated. It is evident that a small number of very prominent photobiont genera (Asterochloris, Nostoc, Rhizonema, Trebouxia, Trentepohlia) each partner with many hundreds or thousands of lichen-forming fungal species; a number of others (e.g. Coccomyxa, Elliptochloris, Heveochlorella, Symbiochloris) are lichenized by many dozens or hundreds of different mycobiont species, while much of the remainder participate in only a small number of known lichen associations. It seems probable that further surveys will uncover more photobiont genera in the latter category. While it is widely agreed that the diversity of lichen-forming algae remains considerably less well known than that of lichen-forming fungi, this fact alone is unlikely to account for the enormous disparity between the currently recognized number of photobiont genera (c. 50) and that of mycobiont genera (c. 1000; Lücking et al. Reference Lücking, Hodkinson and Leavitt2017a). The number of photobiont species described, estimated at c. 100 not long ago (Škaloud & Peksa Reference Škaloud and Peksa2010), shows a similar disparity with the number of lichen-forming fungal species (20 000). Indeed, both the generic and species estimates differ between mycobiont and phycobiont by the same factor of 20. Thus, the imbalance is not likely due to differences in genus/species concepts between algae and fungi. Of course, much of the genetic diversity discovered within photobiont genera in the last few years has been reported as clades that still lack taxonomic recognition; species numbers will surely increase substantially in the near future as such diversity becomes formalized biosystematically. However, this still seems unlikely to close the enormous gap with mycobiont species numbers. Rather, the disparities probably indicate a real ecological asymmetry: the large number of lichen-forming fungal taxa may be partnering with a substantially smaller pool of photobiont taxa, many of which are shared among mycobionts. Such was the conclusion reached recently by Dal Forno et al. (Reference Dal Forno, Lawrey, Sikaroodi, Gillevet, Schuettpelz and Lücking2020) in their detailed comparison of genetic diversity in Dictyonema and its Rhizonema photobionts.

A synopsis of algal genera to which lichen photobionts are currently attributed is given below.

Cyanobacteria

Anabaena Bory ex É. Bornet & C. Flahault — See Nostoc. Strains of Anabaena versus Nostoc are resolved in some analyses (Henson et al. Reference Henson, Watson and Barnum2002; Rajaniemi et al. Reference Rajaniemi, Hrouzek, Kaštovská, Willame, Rantala, Hoffmann, Komárek and Sivonen2005; Liu et al. Reference Liu, Zhu, Lu and Song2013; Elshobary et al. Reference Elshobary, Osman, Abushady and Piercey-Normore2015) but formal distinction of the two genera remains controversial (Makra et al. Reference Makra, Gell, Juhász, Soós, Kiss, Molnár, Ördög, Vörös and Balázs2019). Tschermak-Woess (Reference Tschermak-Woess and Galun1988a) recommended re-examination of earlier reports that Anabaena occurs as cephalodial photobiont of Stereocaulon.

Anacystis Meneghini — According to Bold & Wynne (Reference Bold and Wynne1985), this generic name has been applied to ellipsoid to cylindrical cyanobacteria that often accumulate in a common gelatinous matrix, with some authors also including spheriodal-celled taxa such as Gloeocapsa and Chroococcus. The much-studied ‘Anacystis nidulans’ is usually treated now under Synecococcus; other taxa are currently placed in Microcystis. Photobionts attributed to Anacystis in the past include the partners of a small number of Peltula species and the cephalodial symbionts of a Stereocaulon (see Tschermak-Woess Reference Tschermak-Woess and Galun1988a); determining their identities with confidence will require further study.

Brasilonema Fiore et al. — This cyanobacterial genus, forming a distinct clade in molecular analyses (Fiore et al. Reference Fiore, Sant'Anna, Azevedo, Komárek, Kaštovský, Sulek and Lorenzi2007), has aggregated filaments morphologically similar to Scytonema but only rarely showing false branching. A recent paper reported new species of both Brasilonema and Chroococcidiopsis as co-occurring photobionts of an unidentified lichen growing on gravestones in a northern Florida cemetery (Villanueva et al. Reference Villanueva, Hašler, Dvořák, Poulíčková and Casamatta2018). However, as no description or evidence of this association has yet been published, the status of Brasilonema as lichen photobiont awaits corroboration.

Calothrix C. Agardh ex É. Bornet & C. Flahault and Dichothrix G. Zanardini ex É. Bornet & C. Flahault — These filamentous cyanobacteria are members of the Rivulariaceae; their trichomes have a basal heterocyte and gradually narrow towards the apex. The two genera are morphologically similar and both have been reported as lichen photobionts, particularly in association with certain species of Lichina (see Tschermak-Woess Reference Tschermak-Woess and Galun1988a). However, DNA sequences obtained from two such examples instead placed the algae in question in the genus Rivularia (Ortiz-Álvarez et al. Reference Ortiz-Álvarez, de los Ríos, Fernández-Mendoza, Torralba-Burrial and Pérez-Ortega2015). The photobiont of Placynthium nigrum isolated into culture also shows the distinctive Rivulariaceae morphology (apically tapering filaments with basal heterocytes) while the lichenized filaments rather resemble those now placed in Rhizonema (see Geitler Reference Geitler1934). The circumscription of Calothrix and Dichothrix with respect to lichen photobionts currently remains unresolved.

Chroococcidiopsis Geitler (and Myxosarcina H. Printz) — These unicellular cyanobacteria are found in a great diversity of habitats and include extremophiles. Cells divide in sequence by binary fission, often in alternating planes to produce more or less cubical packages of cells. Cells can also undergo multiple fission to produce four or more autospore-like products known as baeocytes, initially contained within the sheath-like, fibrous outer wall layer of the mother cell (Waterbury & Stanier Reference Waterbury and Stanier1978). The baeocytes of Myxosarcina, unlike those of Chroococcidiopsis, have a brief stage of gliding motility; the genera are said to be otherwise indistinguishable morphologically. The baeocyte-forming cyanobacteria were formerly grouped together in the order Pleurocapsales (Waterbury & Stanier Reference Waterbury and Stanier1978), but SSU sequence analysis has shown this trait to be a convergence shared by a number of lineages of quite different origin (Fewer et al. Reference Fewer, Friedl and Büdel2002). In that study, several photobionts isolated from Lichinaceae appear within the same clade as Chroococcidiopsis thermalis, sister to the heterocyte-forming Stigonematales and Nostocales, and distant from Myxosarcina as well as other morphologically similar taxa formerly attributed to Chroococcidiopsis (Fewer et al. Reference Fewer, Friedl and Büdel2002). Sequences obtained from photobionts of several Peltula species collected in Vietnam also suggested affinities within a broad ‘Chroococcidiopsidales’ clade (Võ Reference Võ2016). Other algal partners of Lichinaceae have been attributed to Chroococcidiopsis based on morphology and the production of baeocytes observed in cultured isolates (Büdel & Henssen Reference Büdel and Henssen1983). Tschermak-Woess (Reference Tschermak-Woess and Galun1988a) suggested that some taxa identified as Chroococcidiopsis might actually belong to Gloeocapsa and require study in culture. Most photobiont isolates attributed to Chroococcidiopsis and Myxosarcina await more detailed molecular scrutiny.

Chroococcus Nägeli — A morphologically distinctive cyanobacterial genus, Chroococcus has relatively large, spherical cells that divide at consecutive right angles to produce small packets of cells, often within concentric, gelatinous sheath layers. A number of reports, compiled by Tschermak-Woess (Reference Tschermak-Woess and Galun1988a), attribute thallus and cephalodial photobionts of various lichens to this genus or merely to Chroococcaceae, or Chroococcales. Many are anecdotal and most await reinvestigation with molecular sequence comparisons. The photobionts of certain Dictyonema species, once attributed to Chroococcus, have been shown to belong instead to Rhizonema, a usually filamentous taxon that may be greatly altered morphologically in certain lichen associations (Lücking et al. Reference Lücking, Lawrey, Sikaroodi, Gilleve, Chaves, Sipman and Bungartz2009). The circumscription of Chroococcus and its status as a lichen photobiont genus remain uncertain at present.

Gloeocapsa Kützing — This colonial cyanobacterium has roundish to oblong cells surrounded individually and communally by successive layers of dense mucilage, reflecting the sequence of cell divisions. Morphologically defined at present, Gloeocapsa commonly occurs free-living in moist terrestrial habitats and is also reported as thallus photobiont in several genera of Lichinaceae, and as cephalodial symbiont in certain species of Stereocaulon and Amygdalaria (Tschermak-Woess Reference Tschermak-Woess and Galun1988a). In the lichen Gonohymenia, contacting mycobiont hyphae broadly invaginate the cells of its photobiont, identified as Gloeocapsa (Paran et al. Reference Paran, Ben-Shaul and Galun1971). Geitler (Reference Geitler1933) described appressorial hyphae in the lichen Synalissa that branch in synchrony with the binary fission of its Gloeocapsa photobiont.

Molecular sequence data are much needed to understand the relationship among taxa currently assigned to Gloeocapsa.

Hyella É. Bornet & C. Flahault — The filamentous cyanobacterium Hyella is a widespread inhabitant of the marine intertidal zone, where it colonizes calcareous substrata such as mollusc shells. The substratum is penetrated by threads arising from a basal system at the surface; endospore-like baeocytes may be formed (Fritsch Reference Fritsch1945). Genomic analysis shows Hyella phylogenetically nearest to the genus Chroococcidiopsis (Brito et al. Reference Brito, Vieira, Vieira, Zhu, Leão, Ramos, Lu, Vasconcelos, Gugger and Tamagnini2020). Hyella is reported to be the photobiont of some species of fungi now assigned to Collemopsidium (Mohr et al. Reference Mohr, Ekman and Heegaard2004). However, details of the symbiotic interaction are few; other genera of cyanobacteria, such as Gloeocapsa and Nostoc, are also said to be photobionts for Collemopsidium [=Pyrenocollema] (Purvis et al. Reference Purvis, Coppins, Hawksworth, James and Moore1992).

Hyphomorpha A. Borzi — These seldom encountered cyanobacteria occur as epiphytes upon tropical liverworts and tree bark, where they form a prostrate filament system. The filaments have an apical cell producing derivatives that may later divide periclinally to become pluriseriate, as do structurally similar species of Stigonema. Cells of these older portions tend to fall out of alignment and become jumbled into a ‘chroococcoid stage’ (Fritsch Reference Fritsch1945). Hyphomorpha was first identified as photobiont in two species of Spilonema lichens by Henssen (Reference Henssen1981), who reported confirmation of the alga's identity by eminent phycologist Lothar Geitler. One of these mycobiont species has been recently reclassified as Erinacellus dendroides (Spribille et al. Reference Spribille, Tønsberg, Stabentheiner and Muggia2014). At present, the algal genus Hyphomorpha is phenotypically defined; it is currently placed in Fischerellaceae (Büdel & Kauff Reference Büdel, Kauff and Frey2012) or included under Hapalosiphonaceae (Komárek et al. Reference Komárek, Kaštovský, Mareš and Johansen2014) within the Nostocales.

Nostoc Vaucher ex É. Bornet & C. Flahault — This genus accommodates cyanobacteria occurring worldwide in fresh water and upon soil, bark and low-growing plants, with some strains highly desiccation-tolerant (Dodds et al. Reference Dodds, Gudder and Mollenhauer1995). Phenotypically defined at present, taxa attributed to Nostoc fall within several distinct clades of the Nostocales, making the genus polyphyletic (Rajaniemi et al. Reference Rajaniemi, Hrouzek, Kaštovská, Willame, Rantala, Hoffmann, Komárek and Sivonen2005; Gagunashvili & Andrésson Reference Gagunashvili and Andrésson2018). These algae typically form darkly pigmented, mucilaginous macrocolonies of highly variable size and shape, ranging from spheres to irregularly pustulose mats to tangles of cord-like axes. Embedded within the gelatinous matrix are uniseriate trichomes markedly constricted at the cross walls, giving individual cells an almost spherical to barrel-shaped form and the filaments a characteristic string-of-beads appearance. Cell division is diffuse, without apical cells or directional polarity. At intervals along the chain of vegetative cells are slightly larger, thicker-walled, lighter-coloured heterocytes (heterocysts) that specialize as centres of nitrogen fixation. Since the enzyme involved in this process is inhibited by the presence of oxygen, heterocytes lack oxygen-generating Photosystem II (Wolk et al. Reference Wolk, Ernst, Elhai and Bryant1994); electron donors are imported and fixed nitrogen is exported via microplasmodesmatal connections with neighbouring vegetative cells (Giddings & Staehelin Reference Giddings TH and Staehelin1981; Kumar et al. Reference Kumar, Mella-Herrera and Golden2010). Thus, prokaryotic Nostoc and its heterocytic relatives show degrees of cell specialization and intercellular transport characteristic of true multicellular organization (Garcia-Pichel Reference Garcia-Pichel and Schaechter2009).

Nostoc, like many filamentous cyanobacteria, has a motile phase. Short filament segments known as hormogonia are produced by multiple divisions of the vegetative cells between two heterocytes, then break free (Boissière et al. Reference Boissière, Boissière, Champion-Arnaud, Lallmant and Wagner1987; Paulsrud Reference Paulsrud2001). The segments disperse or migrate directionally by a gliding motion that involves secretion of polysaccharide, against which proteinaceous pili appear to push or pull the trichome (Khayatan et al. Reference Khayatan, Meeks and Risser2015). Under favourable conditions, the hormogonia lose motility and differentiate heterocytes as they transition to vegetative filaments (Paulsrud Reference Paulsrud2001). It is conceivable that motile hormogonia might facilitate symbiont encounters in the formation of cyanolichens, as also suspected of flagellate stages in eukaryotic photobionts, but direct evidence is lacking. In the establishment of plant-Nostoc symbioses, the role of hormogonia as infective agents is well known (Adams et al. Reference Adams, Duggan, Jackson and Whitton2012), and genes related to hormogonial function have been identified in lichen-symbiotic strains (Gagunashvili & Andrésson Reference Gagunashvili and Andrésson2018). Nostoc may also disperse temporally by forming akinetes, a kind of resistant spore that develops from a vegetative cell and endures adverse conditions.

Nostoc is photobiont in the majority of cyanophilic lichens. In the Peltigerales, Nostoc serves as principal thallus photobiont, or as secondary photobiont specialized for nitrogen fixation within discrete structures known as cephalodia; these are formed upon or within a thallus that has a green alga as principal photobiont. In a number of cases, Nostoc may serve as both principal and secondary photobiont of a single mycobiont species or individual; this results in cyanomorph and cephalodiate chloromorph thalli that may be either separate or conjoined (James & Henssen Reference James, Henssen, Brown, Hawksworth and Bailey1976; Brodo & Richardson Reference Brodo and Richardson1978; Tønsberg & Holtan-Hartwig Reference Tønsberg and Holtan-Hartwig1983; Armaleo & Clerc Reference Armaleo and Clerc1991; Stenroos et al. Reference Stenroos, Stocker-Wörgötter, Yoshimura, Myllys, Thell and Hyvönen2003; Moncada et al. Reference Moncada, Coca and Lücking2013; Simon et al. Reference Simon, Goffinet, Magain and Sérusiaux2018). The same strain of Nostoc may occur in both morphs (Paulsrud et al. Reference Paulsrud, Rikkinen and Lindblad1998, Reference Paulsrud, Rikkinen and Lindblad2001). In many such instances, chloromorph and cyanomorph are both foliose, but in some species of Lobaria and Sticta, the Nostoc-containing cyanomorph is a branching, fruticose growth that bears no resemblance to the foliose chloromorph (Jordan Reference Jordan1972; James & Henssen Reference James, Henssen, Brown, Hawksworth and Bailey1976; Tønsberg & Goward Reference Tønsberg and Goward2001; Magain et al. Reference Magain, Goffinet and Sérusiaux2012); when growing separately, the two morphs were long presumed to represent very different taxa. Thallus morphology would appear to be influenced by the distinct photobionts in such cases. In certain species of Pseudocyphellaria on the other hand, the independently growing ‘cyanomorphs’ include numerous clusters of the green algal symbiont (probably Symbiochloris) spread among the Nostoc within the algal layer (Henskens et al. Reference Henskens, Green and Wilkins2012), with no visible alterations to thallus morphology. Even when Nostoc serves as a secondary (cephalodial) photobiont in a mature lichen, it may be acquired at a very early stage of lichen formation through contact and capture by the developing mycobiont prothallus (Ott Reference Ott1988; de los Ríos et al. Reference de los Ríos, Raggio, Pérez-Ortega, Vivas, Pintado, Green, Ascaso and Sancho2011). Once organized, thallus lobes containing green algae may secondarily encounter and incorporate compatible Nostoc on the lower surface (Jordan Reference Jordan1970; Jordan & Rickson Reference Jordan and Rickson1971), or either the upper or lower surface (Cornejo & Scheidegger Reference Cornejo and Scheidegger2013). Mycobiont selectivity for particular strains of Nostoc can be very high (Paulsrud et al. Reference Paulsrud, Rikkinen and Lindblad2001). The Nostoc-containing cyanomorph may in turn capture compatible green algal symbionts that contact the tomentum hyphae of the lower cortex, from which chloromorph lobes arise (Sanders Reference Sanders2001).

In most lichens where it is primary photobiont, Nostoc is confined to a discrete algal layer; its filaments are often broken up or contorted into cell clusters with little secretion of mucilaginous sheath material (Fig. 1J). When isolated into culture, it reverts to the morphology and growth pattern typical of its free-living state (Kardish et al. Reference Kardish, Kessel and Galun1989). However, in many of the so-called gelatinous lichens, the form of the Nostoc is not fundamentally altered in lichenization; it maintains the necklace-like filaments and extensive surrounding gelatinous sheath, through which the mycobiont hyphae penetrate (Fig. 1I). In such cases, the photobiont constitutes the main structural component of the lichen, which may maintain an appearance and texture rather similar to that of free-living Nostoc macrocolonies. A recent study suggests that these differences in phenotypic expression, leading to stratified versus gelatinous lichens, may be associated with different genetic strains of Nostoc (Magain & Sérusiaux Reference Magain and Sérusiaux2014). This would appear to be another example where major differences in thallus structure may be correlated with photobiont identity.

Cyanophilic mycobionts can be highly selective of their Nostoc partner strains, often overriding geographical factors (Paulsrud et al. Reference Paulsrud, Rikkinen and Lindblad1998, Reference Paulsrud, Rikkinen and Lindblad2000; Stenroos et al. Reference Stenroos, Högnabba, Myllys, Hyvönen and Thell2006; Myllys et al. Reference Myllys, Stenroos, Thell and Kuusinen2007), although a considerably lower selectivity was observed in lichen communities in maritime Antarctica (Wirtz et al. Reference Wirtz, Lumbsch, Green, Türk, Pintado, Sancho and Schroeter2003). Within a single clade of Peltigera, both highly selective and less discriminating generalist species can be recognized (Magain et al. Reference Magain, Miadlikowska, Goffinet, Sérusiaux and Lutzoni2017, Reference Magain, Truong, Goward, Niu, Goffinet, Sérusiaux, Vitikainen, Lutzoni and Miadlikowska2018). A study of temperate and boreal communities reported genetically distinct terricolous and epiphytic pools of Nostoc, from which Peltigera and Nephroma spp. colonizing those respective substrata select their photobionts (Rikkinen et al. Reference Rikkinen, Oksanen and Lohtander2002). Using a larger data set, Stenroos et al. (Reference Stenroos, Högnabba, Myllys, Hyvönen and Thell2006) found Nostoc photobiont strains to be correlated with mycobiont identity rather than ecological guild. However, fungal preference for the Nostoc photobiont strains of other community members over those sampled from the substratum has been reported in other lichen communities (Cardós et al. Reference Cardós, Prieto, Jylhä, Aragón, Molina, Martínez and Rikkinen2019). In other studies, involving Pannaria and other cyanophilic lichens, both corticolous and saxicolous species sometimes chose closely related strains of Nostoc, and more complex combinations of variable mycobiont selectivity and ecological factors were observed (Elvebakk et al. Reference Elvebakk, Papaefthimiou, Robertsen and Liaimer2008).

Nostoc participates in a range of symbioses besides those it forms with lichen-forming fungi (Adams et al. Reference Adams, Duggan, Jackson and Whitton2012). It is taken up by the locally emergent protoplast of the coenocytic, glomeromycete fungus Geosiphon pyriformis, which then produces a swollen bladder within which the endosymbiotic (endocytobiotic) Nostoc is housed. The intracellular location of the algal symbiont and the close affinities of the fungal component to arbuscular mycorrhizal fungi make the Geosiphon-Nostoc symbiosis quite distinct from fungal-algal symbioses treated under the lichen concept (Kluge et al. Reference Kluge, Mollenhauer, Wolf, Schüßler, Rai, Bergman and Rasmussen2002; Schüßler Reference Schüßler and Hock2012). Nostoc also includes obligatory partners of plants representing several major clades of embryophytes; motile hormogonia are the usual infective agent, and fixed nitrogen, usually in the form of ammonium, is supplied to the host from the numerous heterocytes that differentiate in the symbiotic state (Meeks Reference Meeks1998). In hornworts and the liverwort Blasia, hormogonia enter and inhabit specialized, mucilage-secreting chambers within the gametophytes (Adams & Duggan Reference Adams and Duggan2002). Branched filamentous outgrowths from the inner surfaces of these chambers then develop and increase surface contact between the host and the cyanobacterial colonies (Rodgers & Stewart Reference Rodgers and Stewart1977). In cycad gymnosperms, Nostoc colonizes radial cavities in the cortex of specialized, upward-growing coralloid roots (Costa & Lindblad Reference Costa, Lindblad, Rai, Bergman and Rasmussen2002). Symbiosis with the floating aquatic fern Azolla is unique in that the Nostoc (or Anabaena; Svenning et al. Reference Svenning, Eriksson and Rasmussen2005) is vertically inherited through plant generations, obviating the need for new symbiont capture; the principal cyanobacterium involved cannot be cultivated separately, since its genome shows considerable gene degradation (Ran et al. Reference Ran, Larsson, Vigil-Stenman, Nylander, Ininbergs, Zheng, Lapidus, Lowry, Haselkorn and Bergman2010). In the angiosperm Gunnera, symbiotic Nostoc occurs intracellularly in leaf base tissue (Bergman et al. Reference Bergman, Johanssen and Söderbäck1992). Some of these symbiotic strains, as well as free-living isolates, appear to be similar or closely related to those occurring within lichen thalli or cephalodia, whereas certain other Nostoc strains might be more specialized as lichen photobionts (O'Brien et al. Reference O'Brien, Miadlikowsa and Lutzoni2005; Stenroos et al. Reference Stenroos, Högnabba, Myllys, Hyvönen and Thell2006). Recent genomic comparisons identified certain genes of potential relevance to symbiosis in Nostoc, suggesting also that symbiotic strains may have larger genomes than non-symbiotic ones (Gagunashvili & Andrésson Reference Gagunashvili and Andrésson2018).

Rhizonema Lücking & Barrie — This cyanobacterial genus was resurrected recently to accommodate filamentous, heterocyte-producing photobionts previously assumed to belong to Scytonema, but distinct from that lineage in their 16S rRNA sequences (Lücking et al. Reference Lücking, Lawrey, Sikaroodi, Gilleve, Chaves, Sipman and Bungartz2009). Rhizonema species may be boreal as well as tropical; they are at present known mainly from lichen symbioses but free-living or liverwort-associated populations have also been reported (Cornejo et al. Reference Cornejo, Nelson, Stepanchikova, Himelbrant, Jørgensen and Scheidegger2016). The filaments may be broken up into cell clusters or remain as discrete trichomes (Fig. 1E & F), with sporadic lateral proliferation that has been interpreted as true branching based on the appearance of a mature branch junction (Lücking et al. Reference Lücking, Barrie and Genney2014). This would presumably distinguish Rhizonema from Scytonema, which shows false branching. Thus, when Võ (Reference Võ2016) observed paired false branching in photobionts of Vietnamese Cyphellostereum and Dictyonema, she concluded that the algae were Scytonema rather than Rhizonema, apparently without corroborating molecular data. However, recent observations of Rhizonema, isolated into culture from Dictyonema and identified with genetic sequence comparisons, show branching that appears distinctly false (Fig. 1D). Interestingly, a 19th century illustration of a Dictyonema sericeum thallus (Bornet Reference Bornet1873: plate 12) depicted the photobiont with both double-false branching and seemingly true branching with a junction similar to that shown in Lücking et al. (Reference Lücking, Barrie and Genney2014). The range of branch development modes possible in Rhizonema strains clearly requires further study in both lichenized and aposymbiotic material.

Major genera of lichen-forming fungal partners known so far include Coccocarpia, Erioderma (Peltigerales), and the basidiomycetes Acantholichen, Dictyonema, Cora, Corella and Cyphellostereum (all Hygrophoraceae). In those basidiolichens, the Rhizonema trichome is usually penetrated longitudinally by a single, central mycobiont haustorium quite unlike anything reported in other lichen groups (Roskin Reference Roskin1970; Oberwinkler Reference Oberwinkler, Schwemmler and Schenk1980, Reference Oberwinkler, Hertel and Oberwinkler1984, Reference Oberwinkler and Hock2012; Slocum Reference Slocum1980; Tschermak-Woess Reference Tschermak-Woess1983). Such elaborate intrusive structures differ dramatically from the very limited penetrations known in other lichenized algae and might represent specialized absorptive structures. Carbon transfer has not yet been studied in basidiolichens.

Rivularia C. Agardh ex É. Bornet & C. Flahault — The trichomes of this cyanobacterial genus occur in clusters, often on submerged rocks; each filament has a heterocyte at the base and tends to taper gradually towards the apex. The genus includes the photobionts of a couple of maritime species of Lichina, whose algal symbionts were previously attributed to the morphologically similar genus Calothrix (Ortiz-Álvarez et al. Reference Ortiz-Álvarez, de los Ríos, Fernández-Mendoza, Torralba-Burrial and Pérez-Ortega2015).

Scytonema C. Agardh ex É. Bornet & C. Flahault — This aquatic or aerophilic genus of cyanobacteria has trichome walls unconstricted at the septa, with vegetative cells usually wider than long, prominent heterocytes, and thick sheaths that are often darkly pigmented. Scytonema is traditionally recognized by the frequently paired (‘double’) false branches, where segments created by a break in the trichome continue linear growth by simply reorienting laterally and emerging from their formerly common sheath. Trichome breaks may arise where intercellular material is deposited as a separation disc, or one or more cells degenerate, or at intercalary heterocyte positions (Bhâradwâja Reference Bhâradwâja1933). Once considered a significant photobiont genus, including both principal and secondary (cephalodial) lichen symbionts, Scytonema in its current sense encompasses an uncertain but much reduced number of lichen algae. Photobionts previously ascribed to Scytonema have been shown by DNA sequence analyses to belong to a quite distinct clade, now designated Rhizonema (Lücking et. al. Reference Lücking, Lawrey, Sikaroodi, Gilleve, Chaves, Sipman and Bungartz2009). Nevertheless, at least one recent photobiont sequence (16s rRNA), from a Heppia thallus, appears to fall within Scytonema in the strict sense (Võ Reference Võ2016). This may provide some corroboration for previous attributions of Heppia photobionts to Scytonema based on morphology of cultured isolates (Wetmore Reference Wetmore1970). The cell shape and division planes of the Heppia photobionts are radically transformed to produce cell clusters in the lichenized state, reverting quickly to typical filamentous growth when cultured aposymbiotically (Marton & Galun Reference Marton and Galun1976). In Pyrenothrix nigra, the lichenized filamentous cyanobiont shows the double-false branching typical of Scytonema (Tschermak-Woess et al. Reference Tschermak-Woess, Bartlett and Peveling1983), although Lücking et al. (Reference Lücking, Lawrey, Sikaroodi, Gilleve, Chaves, Sipman and Bungartz2009) suggested that its photobiont might be Rhizonema. This is quite plausible, since there is some doubt as to whether the two cyanobacterial genera can be reliably distinguished by their mode of branching (see comments under Rhizonema). More sequence data are clearly needed to clarify the extent to which lichen symbioses may involve the genus Scytonema in its current, more restricted sense.

Stigonema C. Agardh ex É. Bornet & C. Flahault — This cyanobacterial genus is recognized by its complex, branching axes with cells dividing in perpendicular planes as in true parenchyma. Filaments are uniseriate at the apex but become locally multiseriate proximally by periclinal divisions, often but not necessarily associated with the formation of true branches laterally. After division, cells retain continuity at the central portion of the septum, where micropores traverse the septal wall (Butler & Allsopp Reference Butler and Allsopp1972). Stigonema has been reported as thallus photobiont in Ephebe and Spilonema, and also as cephalodial partner in numerous species of Stereocaulon (Tschermak-Woess Reference Tschermak-Woess and Galun1988a). The genus awaits molecular treatment, remaining morphologically defined for the time being.

Tolypothrix Kützing ex É. Bornet & C. Flahault — These are filamentous cyanobacteria resembling Scytonema but with usually single- rather than double-false branches emerging from filament breaks; one side of the break grows out as the false branch, the other usually differentiates as a heterocyte. Tolypothrix has been reported as photobiont of the ‘primitively lichenized’ Thermutopsis jamesii based on morphology in collected material (Henssen Reference Henssen1990). Molecular sequences obtained from cephalodia of Placopsis placed the cyanobionts in or near Tolypothrix (Raggio et al. Reference Raggio, Green, Crittenden, Pintado, Vivas, Pérez-Ortega, de los Ríos and Sancho2012).

Green algae (Viridiplantae – Archaeplastida)

Apatococcus F. Brand — Abundant and widely distributed as a free-living organism, Apatococcus has long been known as an omnipresent subaerial unicellular alga, inevitably encountered but not chosen by discriminating germling hyphae of lichen-forming fungi. Now it appears that Apatococcus includes lichen symbionts as well. Light microscopic observations of algal symbionts cultured from several maritime lichen species first implicated Apatococcus as a photobiont (Watanabe et al. Reference Watanabe, Nakano and Deguchi1997); molecular sequence comparisons later identified Apatococcus strains as partners of Scoliciosporum (Beck Reference Beck2002) and Fuscidea species (Zahradníková et al. Reference Zahradníková, Andersen, Tønsberg and Beck2017). Cells are spherical with alternating perpendicular planes of division, producing cuboidal packets of transiently adherent daughter cells. Autospores and biflagellate zoospores are also formed (Ettl & Gärtner Reference Ettl and Gärtner2014). Autospores may be of unequal size within a sporangium (Gärtner & Ingolić Reference Gärtner and Ingolić1989), as also occurs in Watanabalean genera such as Chloroidium and Jaagichlorella. As with Elliptochloris and Trebouxia, Apatococcus is facultatively heterotrophic: it is very slow growing in culture unless carbohydrate is supplied (Gustavs et al. Reference Gustavs, Schumann and Karsten2016). This observation is particularly interesting because the similarly heterotrophic behaviour of Trebouxia in culture was central to Ahmadjian's (Reference Ahmadjian1988, Reference Ahmadjian2002) argument that Trebouxia cannot exist free-living. The seemingly ubiquitous Apatococcus shows quite clearly that a photobiont exhibiting strongly heterotrophic tendencies in culture may nonetheless abound free-living in nature.

Asterochloris Tschermak-Woess — First described to accommodate the trebouxioid photobiont of a single lichen in the Pertusariaceae (Tschermak-Woess Reference Tschermak-Woess1980a), this major photobiont clade now encompasses the former Trebouxia subgenus Eleutherococcus (Tschermak-Woess Reference Tschermak-Woess1989; Škaloud & Peksa Reference Škaloud and Peksa2010). It corresponds roughly to Archibald's (Reference Archibald1975) restricted concept of genus Trebouxia, a source of continual confusion. Asterochloris species produce aplanospores, or zoospores in culture, but most strains do not form the appressed, low-number autospores characteristic of Trebouxia in the current sense. Its deeply-lobed chloroplast becomes flattened and parietal during cell division, while that of Trebouxia remains more or less central (Tschermak-Woess Reference Tschermak-Woess1989). Pyrenoids are present; in TEM they may be distinguished as the irregularis-, erici-, or magna-types of Friedl (Reference Friedl1989). Chloroplast morphology is highly variable and its utility as a marker in species delimitation was emphasized by Škaloud et al. (Reference Škaloud, Steinová, Řidká and Vančurová2015). As with Trebouxia, a considerable amount of genetic diversity is revealed at the molecular level in Asterochloris (Škaloud & Peksa Reference Škaloud and Peksa2010; Peksa & Škaloud Reference Peksa and Škaloud2011).

Sexual fusion of biflagellate isogametes to form a quadriflagellate zygote has been documented in cultures of A. woessiae (Škaloud et al. Reference Škaloud, Steinová, Řidká and Vančurová2015). The detection of genes specific to meiosis in A. glomerata (Armaleo et al. Reference Armaleo, Müller, Lutzoni, Andrésson, Blanc, Bode, Collart, Dal Grande, Dietrich and Grigoriev2019) provides further support for a functioning sexual cycle in Asterochloris.

Asterochloris is associated principally with mycobionts of the Cladoniaceae, Stereocaulaceae, and the genus Lepraria. These fungal partners appear to range from moderately to rather highly selective of their Asterochloris symbionts; there is also some indication that mycobionts of different clades are choosing particular Asterochloris lineages, showing distinct climatic preferences related to rainfall regime (Peksa & Škaloud Reference Peksa and Škaloud2011).

Auxenochlorella (I. Shihira & R. W. Krauss) T. Kalina & M. Puncochárová — Within the Chlorellaceae, Auxenochlorella is related to the fully heterotrophic genus Prototheca, and its type species, A. protothecoides, is also known for its heterotrophic tendencies in culture (Darienko & Pröschold Reference Darienko and Pröschold2015). Auxenochlorella has been implicated in regard to the identity of the photobiont associated with Psoroglaena stigonemoides in the Verrucariaceae (Nyati et al. Reference Nyati, Beck and Honegger2007; Thüs et al. Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011). Unlike Chlorella, Auxenochlorella lacks a pyrenoid. The genus also includes ‘zoochlorellae’ symbionts of the cnidarian Hydra that are now considered a new species, A. symbiontica (Darienko & Pröschold Reference Darienko and Pröschold2015).

Bracteacoccus Tereg — Bracteacoccus are small, spherical unicells that have a multinucleate stage as they mature, and reproduce by zoospores or aplanospores; chloroplasts lack pyrenoids (Kouwets Reference Kouwets1996). Currently included in the Sphaeropleales (Fučíková et al. Reference Fučíková, Lewis and Lewis2014), Bracteacoccus appears at present to be the only genus of the class Chlorophyceae into which lichen photobionts have been placed with supporting DNA sequence data. The corresponding mycobionts are two species of the basidiomycete Sulzbacheromyces in the Lepidostromatales (Hodkinson et al. Reference Hodkinson, Moncada and Lücking2014; Masumoto Reference Masumoto2020).

Cephaleuros Kunze ex E. M. Fries — These foliicolous relatives of Trentepohlia form macroscopic, multicellular thalli visible as small, fuzzy yellow-orange patches on leaves and fruit in tropical and subtropical climates. Cephaleuros species typically grow beneath the cuticle of the leaf substratum, forming rounded to lobed thalli of more or less integrated horizontal filaments. These give rise to the erect setae and sporangiophores that emerge through the overlying cuticle. Usually, Cephaleuros develops within the space it excavates between the host cuticle and epidermis; in some cases, filaments penetrate deeper among the epidermal or mesophyll cells of the leaf, provoking a localized phellogen wound response. The alga can therefore be mildly pathogenic, but it is more often described as ‘parasitic’, despite an absence of information concerning any nutritional exchange with the plant host. Occasionally, the alga may develop upon the leaf cuticle, like other epiphylls. The behaviour may vary according to the species of Cephaleuros or that of the host plant (Ward Reference Ward1884; Suto & Ohtani Reference Suto and Ohtani2009; Brooks et al. Reference Brooks, Rindi, Suto, Ohtani and Green2015). Lichenization by foliicolous Strigula fungi is said to curb the alga's invasion of host tissue and its localized pathogenic effects (Joubert & Rijkenberg Reference Joubert and Rijkenberg1971).

Morphologically, Cephaleuros can somewhat resemble the related foliicolous genus Phycopeltis, which is not subcuticular and generally lacks vertical hairs and complex, long-stalked sporangiophores. According to molecular sequence analyses, however, the nearest relatives of Cephaleuros lie not within Phycopeltis but rather Stomatochroon (Zhu et al. Reference Zhu, Hu and Liu2017), a microscopic colonizer of the leaf's substomatal cavities. While trentepohliaceous taxa currently ascribed to Phycopeltis and Trentepohlia are phylogenetically intertwined, Cephaleuros appears to be essentially monophyletic (López-Bautista et al. Reference López-Bautista, Rindi and Guiry2006; Rindi et al. Reference Rindi, Lam and López-Bautista2009; Nelsen et al. Reference Nelsen, Rivas Plata, Andrew, Lücking and Lumbsch2011; Zhu et al. Reference Zhu, Hu and Liu2017).

Cephaleuros is one of the few lichen photobiont genera for which life cycle events have been observed in some detail. Thompson & Wujek (Reference Thompson and Wujek1997) describe a haplodiplontic life cycle with heteromorphic multicellular phases. The familiar thallus corresponds to the gametophyte; fusion of gametes produces a zygote that germinates into a short-stalked, dwarf sporophyte bearing a putative meiosporangium. Flagellate meiospores presumably develop into new gametophytic thalli. The Cephaleuros gametophyte is the phase known to serve as phycobiont for the fungus Strigula. Whether or not the sporophytes can also be lichenized is unknown. Perhaps they are too highly reduced or short-lived, but the question does not seem to have been explored. On gametophyte thalli, two distinct structures produce flagellate zoospores or gametes (often called zoïds, or swarmers, when their function is uncertain or polyvalent). Zoosporangia are elevated in groups upon vertical stalks; they produce quadriflagellate zoospores that have been observed to round off, germinate and reproduce the gametophyte thallus asexually (Ward Reference Ward1884; Thompson & Wujek Reference Thompson and Wujek1997). The mature sporangia detach readily as units of dispersal, for which both wind and insects act as vectors. On the horizontal filament system, single, usually terminal cells may enlarge to become what are referred to as gametangia; these produce biflagellate cells that may fuse sexually (Thompson & Wujek Reference Thompson and Wujek1997). However, other authors have been unable to observe any instances of sexual fusions in the taxa they studied, instead reporting that the biflagellate zoïds germinate directly as zoospores to form new gametophyte thalli (Suto & Ohtani Reference Suto and Ohtani2013).

Ward (Reference Ward1884) described in detail the course of lichenization of Cephaleuros by Strigula. Young germlings of Cephaleuros are often quickly overrun by the mycelium of Strigula, suppressing algal reproduction, while individuals contacted at more advanced stages of development may produce abundant sporangia from portions of its thalli remaining relatively free of mycobiont domination. Interestingly, both symbionts grow and sporulate independently upon the leaf substratum, although the fungus Strigula will produce pycnidia and perithecia only after successful lichenization. These observations highlight the flexibility of the symbionts in this particular association.

Chlamydomonas Ehrenberg — This well-known unicellular green algal genus chiefly encompasses aquatic taxa that are flagellate in the vegetative state and unlikely candidates for lichen symbiosis. However, a number of aeroterrestrial species are also known (Ettl & Gärtner Reference Ettl and Gärtner2014). One species of Chlamydomonas (C. augustae) was described in association with the ascomycete Pyronema laetissimum, growing on leaf litter in Latvia (Skuja Reference Skuja1943). It was included in Tschermak-Woess's (Reference Tschermak-Woess and Galun1988a) review of phycobionts as ‘facultatively lichenized’. However, Skuja (Reference Skuja1943) distinguished this association from lichen and lichenoid symbioses, making comparisons instead with green algae known to grow abundantly on the surfaces of perennial basidiocarps. The Chlamydomonas was abundantly present among the dense hyphae below the Pyronema apothecium, but no tissue layer was differentiated, nor were any distinctive contact interfaces noted between symbionts. Skuja also mentioned that other apothecia of the same fungus were fruiting nearby without the alga present. The operculate discomycetes (Pezizales), to which Pyronema belongs, are not otherwise known to include lichen-forming members. The Pyronema-Chlamydomonas association is worthy of further investigation but seems unlikely to fit the criteria usually ascribed to lichen symbioses.

Chlorella Beijerinck — A once-notorious miscellany of indistinguishable ‘little round green things’, this trebouxiophycean genus has been radically deconstructed, particularly with the help of DNA sequence comparisons (Huss et al. Reference Huss, Frank, Hartmann, Hirmer, Kloboucek, Seidel, Wenzeler and Kessler1999). Many formerly included species have been moved to different genera, orders, even classes, while taxa surrounding the type species C. vulgaris, and C. sorokiniana, are retained as true Chlorella. In TEM, they show a distinctive pyrenoid surrounded by a thick sheath of starch and bisected centrally by a single thylakoid (Ikeda & Takeda Reference Ikeda and Takeda1995; Němcová & Kalina Reference Němcová and Kalina2000; Hoshina et al. Reference Hoshina, Iwataki and Imamura2010). Flagellate cells and sexual reproduction are unknown. True Chlorella also includes a number of mucilaginous, colonial forms in its current circumscription (Luo et al. Reference Luo, Pröschold, Bock and Krienitz2010; Bock et al. Reference Bock, Krienitz and Pröschold2011). Many of the lichen photobionts previously attributed to Chlorella s. lat. (e.g. Tschermak-Woess Reference Tschermak-Woess1988b) are among those taxa moved to other genera, especially Chloroidium; others await re-examination. At present, only a couple of lichen-forming fungal species have photobionts of corroborated placement within Chlorella (Porpidia crustulata; Li et al. Reference Li, Feng and Xie2013) or Chlorellales. The genus has also been long identified with endosymbiotic algal symbionts of diverse protists and invertebrates. Molecular sequences confirm that true Chlorella occur as endosymbionts of the ciliate Paramecium bursaria (Hoshina et al. Reference Hoshina, Kamako and Imamura2004; Summerer et al. Reference Summerer, Sonntag and Sommaruga2008) and the cnidarian Hydra (Kovačević et al. Reference Kovačević, Franjević, Jelenčić and Kalafatić2010), which may also utilize Auxenochlorella as its algal symbiont (Pröschold et al. Reference Pröschold, Darienko, Silva, Reisser and Krienitz2011). Chloroplast ultrastructure likewise suggests that the green endosymbiont of the colonial ciliate Ophrydium versatile is a true Chlorella (Forsberg & Lindblad Reference Forsberg and Lindblad1996). However, the phylogenetic affinities of other ‘zoochlorellae’ symbionts appear to fall elsewhere in the Trebouxiophyceae (Lewis & Muller-Parker Reference Lewis and Muller-Parker2004; Kovačević et al. Reference Kovačević, Franjević, Jelenčić and Kalafatić2010; Pröschold et al. Reference Pröschold, Darienko, Silva, Reisser and Krienitz2011), while many have yet to be explored with molecular sequence comparisons.

Chloroidium Nadson — Resurrected to accommodate segregates from Chlorella s. lat. (Darienko et al. Reference Darienko, Gustavs, Mudimu, Menendez, Schumann, Karsten, Friedl and Pröschold2010), Chloroidium falls within the trebouxiophycean assemblage now formalized as Watanabeales (Li et al. Reference Li, Tan, Liu, Zhu, Hu and Liu2021). Cells have a parietal chloroplast with or without a pyrenoid; in C. saccharophilum, a prominent pyrenoid with surrounding plastoglobuli and traversing membranes has been observed (González et al. Reference González, Pröschold, Palacios, Aguayo, Inostroza and Gómez2013). Reproduction is by autospores, often of variable number and different sizes within a single sporangium. The genus encompasses diverse taxa found in a wide variety of habitats (Darienko et al. Reference Darienko, Lukešová and Pröschold2018), including extremophiles capable of using a variety of carbon sources (Nelson et al. Reference Nelson, Khraiwesh, Fu, Alseekh, Jaiswal, Chaiboonchoe, Hazzouri, O'Connor, Butterfoss and Dro2017). Since its recent emendation, Chloroidium includes photobiont partners of a growing number of lichen-forming fungi, including some species of Gomphillaceae, Verrucariaceae, Psora, Stereocaulon and Sticta.

Chlorosarcinopsis Herndon — In the course of her studies on lichen haustoria, Plessl (Reference Plessl1963) identified as Chlorosarcina [=Chlorosarcinopsis] minor the photobionts she isolated from two species of Lecidea, L. plana and L. lapicida. Chlorosarcinopsis has traditionally accommodated spherical unicellular green algae dividing to form cuboidal packets. According to Neustupa (Reference Neustupa and Frey2015), the genus is polyphyletic, with members scattered among the Chlamydomonadales (Chlorophyceae). As this clade is not otherwise known for lichen symbionts (but see Skuja Reference Skuja1943), the photobionts of the Lecidea species in question need further study.

Coccobotrys Chodat (now Uvulifera Molinari-Novoa) — This green alga forms irregular cuboidal cell packages or branched multiseriate filaments in culture (Neustupa Reference Neustupa and Frey2015). The genus Coccobotrys was described by Chodat (Reference Chodat1913) and emended by Vischer (Reference Vischer1960) to accommodate the putative photobiont C. verrucariae isolated from a thallus of Verrucaria nigrescens. Thüs et al. (Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011), on the other hand, reported Diplosphaera as photobiont of the V. nigrescens thallus they sampled. Coccobotrys verrucariae was also cited among algae isolated from soil crusts (Flechtner et al. Reference Flechtner, Johansen and Belnap2009), and a photobiont identified with microscopy as ‘probably Coccobotrys’ (Canals et al. Reference Canals, Hernández-Mariné, Gómez-Bolea and Llimona1997) was isolated from Botrylepraria lesdainii, another member of the Verrucariales (Kukwa & Pérez-Ortega Reference Kukwa and Pérez-Ortega2010). A second species of Coccobotrys was described by Warén (Reference Warén1920) as the photobiont of Lecidea fuliginosa, but Tschermak-Woess (Reference Tschermak-Woess and Galun1988a) expressed doubt that the alga he described belongs in Coccobotrys. The photobiont status of species in this genus should be corroborated. Genetic sequence analyses place Coccobotrys in the Trebouxiophyceae (e.g. Thüs et al. Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011; Mikhailyuk et al. Reference Mikhailyuk, Holzinger, Tsarenko, Glaser, Demchenko and Karsten2020), but its affinities among the defined clades within this class remain uncertain. Molinari-Novoa (Reference Molinari-Novoa2016) recently found Coccobotrys Chodat to be a later homonym of a name applied to an anamorphic basidiomycete and renamed the algal genus Uvulifera.

Coccomyxa Schmidle — This trebouxiophycean algal genus is notable for the diversity of habitats and ecological circumstances in which its species are known to occur. Environmental surveys have found Coccomyxa sequences to be among the most widely distributed OTUs, and notably well represented in cold high-latitude climates (Metz et al. Reference Metz, Singer, Domaizon, Unrein and Lara2019). It is commonly reported free-living on terrestrial substrata and in aquatic environments, including those highly polluted with heavy metals and radioactive materials (see Gustavs et al. Reference Gustavs, Schiefelbein, Darienko, Grube, Seckbach and Muggia2017). Coccomyxa species are subspherical to ovoid-ellipsoidal unicells, often embedded colonially in thick gelatinous sheath material with concentric layering that reflects the cell division pattern. The chloroplast is parietal, not markedly lobed, and lacks a pyrenoid. In TEM, thylakoid bands often show a distinctly longitudinal orientation over the length of the chloroplast, with interspersed starch grains (Peveling & Galun Reference Peveling and Galun1976; Palmqvist et al. Reference Palmqvist, de los Ríos, Ascaso and Samuelsson1997). Flagellate cells and sexual reproduction are unknown; cells subdivide into packages of 2–8 autospores (Tschermak-Woess Reference Tschermak-Woess and Galun1988a). Recent assessments of species number within the genus range from seven (Darienko et al. Reference Darienko, Gustavs, Eggert, Wolf and Pröschold2015) to as many as 27 (Malavasi et al. Reference Malavasi, Škaloud, Rindi, Tempesta, Paoletti and Pasqualetti2016). The genus is thought to include the photobionts of diverse lichen-forming fungi, such as species of Icmadophila, Micarea, Nephroma, Peltigera, Solorina, the stalked-apotheciate genera Baeomyces, Dibaeis and Phyllobaeis, and the basidiomycete Lichenomphalia (Table 1). Some of these reports await confirmation with genetic sequence data. The photobionts do not form a single clade but instead represent several distinct lineages within Coccomyxa, intermixed among free-living isolates (Darienko et al. Reference Darienko, Gustavs, Eggert, Wolf and Pröschold2015). In lichen symbiosis, the cells are often more spheroidal, and extensive gelatinous sheath material is not usually produced (Tschermak-Woess Reference Tschermak-Woess and Galun1988a). Interestingly, while the cells of Coccomyxa and Elliptochloris photobionts are tightly enveloped by mycobiont hyphae, their walls are usually not penetrated (Tschermak Reference Tschermak1941a; Geitler Reference Geitler1955; Plessl Reference Plessl1963; but see Coppins Reference Coppins1983: figs 2 & 55). This has been attributed to degradation-resistant polymers resembling sporopollenin in the multilayered cell wall (Honegger & Brunner Reference Honegger and Brunner1981; Brunner & Honegger Reference Brunner and Honegger1985). However, Coccomyxa cells are fully penetrated by Aphelidium collabens, a parasitoid basal within, or sister to, the kingdom Fungi (Seto et al. Reference Seto, Matsuzawa, Kuno and Kagami2020).

Species of the genus Coccomyxa also live in poorly understood symbioses within molluscs (Stevenson & South Reference Stevenson and South1974; Syasina et al. Reference Syasina, Kukhlevsky, Kovaleva and Vaschenko2012) and echinoderms, and endocytotically within ovules and other tissues of the gymnosperm Ginkgo biloba (Trémouillaux-Guiller et al. Reference Trémouillaux-Guiller, Rohr, Rohr and Huss2002; Trémouillaux-Guiller & Huss Reference Trémouillaux-Guiller and Huss2007). Molecular sequence comparisons have shown that some zoochlorellae isolated from certain strains of Paramecium bursaria correspond to Coccomyxa, while most others are true Chlorella (Hoshina & Imamura Reference Hoshina and Imamura2008).

Deuterostichococcus Pröschold & Darienko — A recent segregate of Stichococcus s. lat. (Pröschold & Darienko Reference Pröschold and Darienko2020), this trebouxiophycean genus currently includes, in addition to free-living isolates, the photobiont of two Placopsis species (Beck et al. Reference Beck, Bechteler, Casanova-Katny and Dzhilyanova2019) and Staurothele clopima (Hodač et al. Reference Hodač, Hallmann, Spitzer, Elster, Faßhauer, Brinkmann, Lepka, Diwan and Fried2016); the latter is also known to partner with Diplosphaera algae (Thüs et al. Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011).

Dictyochloropsis Geitler — See Symbiochloris.

Dilabifilum Tschermak-Woess — Polymorphic, unicellular to sarcinoid to filamentous algae with pyrenoids and quadriflagellate zoospores have been included in this ulvalean genus. They occur free-living, as photobionts, or both. Recently, Darienko & Pröschold (Reference Darienko and Pröschold2017) deconstructed Dilabifilum, recognizing at generic level several distinct clades resolved in their gene-based phylogenies. A number of photobionts previously contained therein are now distributed in Halofilum, Lithotrichon, Paulbroadya and Pseudendoclonium, while others await reassessment.

Diplosphaera Bialosuknia — This prasiolalean genus appears to contain the majority of the unicellular photobiont strains attributed until recently to the related Stichococcus. Apparently, the two morphologically plastic genera are often not distinguishable microscopically, although Diplosphaera may produce distinctive, adherent two-celled clusters in division. Pyrenoids may be absent (Pröschold & Darienko Reference Pröschold and Darienko2020) or weakly visible (Ettl & Gärtner Reference Ettl and Gärtner2014) but some taxa falling within the Diplosphaera clade, including lichen photobionts, show prominent pyrenoids in TEM (Fig. 2E). The main fungal partners of Diplosphaera are members of the Verrucariaceae. Photobiont strains compared so far appear to represent the same species and are similar to free-living collections (Pröschold & Darienko Reference Pröschold and Darienko2020).

In association with certain lichen genera, such as Endocarpon and Staurothele, Diplosphaera photobionts ‘escape’ vegetative hyphal contacts and penetrate into the hymenial layer of developing perithecia, where they freely intermix among the asci (Fig. 3A). These algal cells are typically much smaller than those within the algal layer of the vegetative thallus; they scatter everywhere when a hand-cut section is water-mounted, indicating that unlike the photobionts in the vegetative thallus, those entering the perithecia are not bound in place by lichenizing contacts with the mycobiont. The unassociated photobionts may adhere to the large ascospores as they are ejected, and can be dispersed with them. Many readily detach and divide aposymbiotically; they are available to the mycobiont if the spore germinates successfully (Stahl Reference Stahl1877; Bertsch & Butin Reference Bertsch and Butin1967; Ahmadjian & Heikkilä Reference Ahmadjian and Heikkilä1970) or might otherwise divide to form free-living populations. Potentially co-dispersable photobionts also occur in the conidiomata and ascomal epithecia of many foliicolous lichens of the Gomphillaceae and Pilocarpaceae (see Heveochlorella). Dispersal of liberated photobionts can thereby provide a direct connection between lichenized and free-living populations of the alga.

Elliptochloris Tschermak-Woess — Like its sister genus Coccomyxa, Elliptochloris has subspherical to ellipsoidal unicells with a parietal chloroplast, bearing two opposed indentations in the type species E. bilobata (Tschermak-Woess Reference Tschermak-Woess1980b). Sexual or flagellate stages are unknown; reproduction occurs by autospores, of which there are usually two morphologically distinct types. Autosporangia may contain a low number (usually four in cultured E. bilobata) of spherical spores appressed together at flattened junctions, or more numerous (16–32) cylindrical-ellipsoidal spores (Tschermak-Woess Reference Tschermak-Woess1980b; Darienko et al. Reference Darienko, Gustavs and Pröschold2016). The multilayered cell wall, as in Coccomyxa, is impregnated with degradation-resistant polymers, which are thought to explain the lack of haustorial penetration by their lichen-forming partners (Brunner & Honegger Reference Brunner and Honegger1985). However, haustoria have been noted in certain species of Micarea (Coppins Reference Coppins1983: figs 2 & 55), a lichen-forming genus known to partner with Elliptochloris and Coccomyxa photobionts. Unlike Coccomyxa, at least some species of Elliptochloris possess pyrenoids, and layered mucilaginous sheaths are typically lacking. However, gelatinous extracellular material may be copious in free-living populations, and was observed in association with Protothelenella thalli where the photobiont population grew beyond the reach of mycobiont hyphae (Tschermak-Woess Reference Tschermak-Woess1985).

Elliptochloris is somewhat less often reported than Coccomyxa but is known from a similarly diverse array of habitats. It is said to be quite strongly heterotrophic in culture, where it depends heavily on organic materials to thrive; this might in part account for its less frequent recovery in isolation procedures (Gustavs et al. Reference Gustavs, Schiefelbein, Darienko, Grube, Seckbach and Muggia2017). Species of Elliptochloris are known to partner with mycobionts of diverse genera including Catillaria, Catolechia, Fuscidea, Micarea, Sticta, Stictis, Verrucaria, and the basidiolichen-forming Bryoclavula and Multiclavula (see Table 1). They also occur as endosymbionts of the marine anenome Anthopleura (Letsch et al. Reference Letsch, Muller-Parker, Friedl and Lewis2009).

Gloeocystis Nägeli — Taxa treated under this genus are unicellular green algae that form occasionally macroscopic colonies of ellipsoidal cells with a parietal chloroplast possessing a pyrenoid. Thick, colourless, often layered gelatinous sheaths surround the cells. Reports of Gloeocystis as photobiont of Cryptodiscus [Bryophagus] gloeocapsa and Epigloea bactrospora were cited by Ahmadjian (Reference Ahmadjian1967) and Tschermak-Woess (Reference Tschermak-Woess and Galun1988a). There is doubt as to whether Epigloea is lichenized (Kirk et al. Reference Kirk, Cannon, David and Stalpers2001; not included in Lücking et al. (Reference Lücking, Hodkinson and Leavitt2017a)), although distinctly symbiotic contacts with living, unicellular green algae were illustrated by Jaag & Thomas (Reference Jaag and Thomas1934) and Döbbler (Reference Döbbler1984).

According to Neustupa (Reference Neustupa and Frey2015), the algal genus Gloeocystis is highly polyphyletic, encompassing members of both Chlorophyceae and Trebouxiophyceae. The identities of the photobionts associated with the mycobionts mentioned above will therefore require further study.

Halofilum Darienko & Pröschold — This is another genus that now accommodates taxa previously treated under Dilabifilum (Darienko & Pröschold Reference Darienko and Pröschold2017). These algae consist of branched filaments with parietal chloroplasts containing pyrenoids; flagellated stages are unknown. The species H. ramosum occurs as photobiont of Hydropunctaria maura and Wahlenbergiella striatula (Verrucariaceae), as well as free-living (Darienko & Pröschold Reference Darienko and Pröschold2017).

Heveochlorella J. Zhang et al. — Unicellular algae attributed to Heveochlorella have a prominent, somewhat lobed chloroplast with a central pyrenoid that is readily visible with light microscopy. TEM shows the pyrenoid surrounded by several irregular starch plates and penetrated centripetally by thylakoid-derived tubules that are lined with pyrenoglobuli (Fig. 2B). Cells reproduce by autospores, usually in low number (2–8) and not infrequently of unequal size within sporangia, at least in culture (Zhang et al. Reference Zhang, Huss, Sun, Chang and Pang2008; Ma et al. Reference Ma, Huss, Tan, Sun, Chun, Xie and Zhang2013; Sanders et al. Reference Sanders, Pérez-Ortega, Nelsen, Lücking and de los Ríos2016). Darienko & Pröschold (Reference Darienko and Pröschold2019) recently subsumed both Heveochlorella (Zhang et al. Reference Zhang, Huss, Sun, Chang and Pang2008) and the related Heterochlorella (Neustupa et al. Reference Neustupa, Němcová, Eliáš and Škaloud2009), which has not been reported from lichen symbioses, into the resurrected genus Jaagichlorella. These algae belong to the trebouxiophyceaen clade recently formalized as Watanabeales (Li et al. Reference Li, Tan, Liu, Zhu, Hu and Liu2021).

The first indication that lichen symbionts belonged in this group was the report of Heveochlorella isolated as photobiont from one specimen of Sticta and two of Pseudocyphellaria from Taiwan (Dal Grande et al. Reference Dal Grande, Beck, Cornejo, Singh, Cheenacharoen, Nelsen and Scheidegger2014b). Soon thereafter, the ‘trebouxioid’ photobionts associated with foliicolous Gomphillaceae and Pilocarpaceae were also attributed to this genus (Sanders et al. Reference Sanders, Pérez-Ortega, Nelsen, Lücking and de los Ríos2016). More recently, a study of Sticta lichens worldwide reported Heveochlorella to be the photobiont of numerous specimens from New Zealand and Indian Ocean islands, including six identified species and many undetermined collections (Lindgren et al. Reference Lindgren, Moncada, Lücking, Magain, Simon, Goffinet, Sérusiaux, Nelsen, Mercado-Díaz and Widhelm2020). In the opinion of Darienko & Pröschold (Reference Darienko and Pröschold2019), the algae encompassed by Jaagichlorella, though distributed worldwide, are rare taxa. While more surveys will be necessary to evaluate this view, a number of observations suggest that these algae might be in fact quite common and merely overlooked. We know, for example, that foliicolous lichens of the Gomphillaceae and Pilocarpaceae occur in abundance throughout much of the humid tropics (Santesson Reference Santesson1952; Lücking Reference Lücking2008), although it is not yet clear how consistently they harbour Heveochlorella (Jaagichlorella) photobionts. Recent sampling of the phyllosphere community in Asian tropical forests has revealed a major representation of Heveochlorella genotypes (Zhu et al. Reference Zhu, Li, Hu and Liu2018), as well as several new species in related genera (Li et al. Reference Li, Sun, Hu, Liu, Zhu, Hu and Liu2020, Reference Li, Tan, Liu, Zhu, Hu and Liu2021). Some of the most frequently detected OTUs in environmental surveys of marine habitats (Metz et al. Reference Metz, Singer, Domaizon, Unrein and Lara2019) were also attributed to Heveochlorella.

In many foliicolous lichens, dividing Heveochlorella photobionts may escape the lichenizing vegetative hyphae and proliferate among the spore-generating fungal structures, upon apothecia and within specialized conidiomata such as campylidia and hyphophores (Fig. 3B & C). They can be dispersed from these structures, as are the fungal spores or diahyphae to which the algal cells may adhere (Fig. 3D–F). Once dispersed, they may become lichenized by the germinating fungal propagules, or divide to produce independent populations on the substratum (Sanders Reference Sanders2014; Sanders & de los Ríos Reference Sanders and de los Ríos2015). Co-dispersal and relichenization thereby provide Heveochlorella with abundant opportunities for exchange between lichenized and free-living populations.

Interfilum Chodat — This genus of aeroterrestrial charophytes (Streptophyta) includes taxa that form single, paired or sarcinoid packets of cells or grow filamentously, often closely resembling unrelated Chlorophyta, such as Desmococcus (Mikhailyuk et al. Reference Mikhailyuk, Sluiman, Massalski, Mudimu, Demchenko, Kondratyuk and Friedl2008). It is sister to clades of the widely distributed Klebsormidium (Rindi et al. Reference Rindi, Mikhailyuk, Sluiman, Friedl and López-Bautista2011). Interfilum was reported by Voytsekhovich et al. (Reference Voytsekhovich, Dymytrova and Nadyeina2011) as a secondary photobiont within the algal layer of Micarea and Placynthiella thalli collected in Ukraine, based on light microscopic examination of thalli and cultured isolates. The principal photobionts in those lichens were reported to be Elliptochloris and Radiococcus, respectively. As the charophytes are not otherwise known as lichen symbionts, and other algal genera were cited as the main photobionts within the thalli in question, further study of the reported associations is warranted.

Jaagichlorella Reisigl — See Heveochlorella.

Leptosira A. Borzì — This photobiont grows as unicells tightly wrapped by mycobiont hyphae or separated by copious sheath material free of the mycobiont; in agar culture, it produces short filaments (Tschermak-Woess Reference Tschermak-Woess1953). Leptosira is a trebouxiophycean of uncertain placement, appearing in the vicinity of the Microthamniales clade in recent gene-based phylogenies (Lemieux et al. Reference Lemieux, Otis and Turmel2014; Neustupa Reference Neustupa and Frey2015; Hallmann et al. Reference Hallmann, Hoppert, Mudimu and Friedl2016). According to Mattox & Stewart (Reference Mattox, Stewart, Irvine and John1984), ‘Pleurastrum terrestre’ (a synonym of Leptosira obovata, now L. terrestris; Friedl Reference Friedl1996) is so similar ultrastructurally to the genus Trebouxia that they could be combined in the same genus. Ahmadjian (Reference Ahmadjian1988) went one step further, opining that Trebouxia was merely the lichenized form of this taxon. However, the aforementioned gene-based cladograms do not show a close relationship between Leptosira and the Trebouxiales.

Leptosira terrestris, in lichen symbiosis with Vezdaea aestivalis, grows subcuticularly (Tschermak-Woess & Poelt Reference Tschermak-Woess, Poelt, Brown, Hawksworth and Bailey1976), a distinction shared with the photobiont Cephaleuros. Leptosira is also among the very few photobiont genera (Phycopeltis and Cephaleuros are others) reported to produce zoospores in the lichenized state (Tschermak-Woess & Poelt Reference Tschermak-Woess, Poelt, Brown, Hawksworth and Bailey1976).

Lithotrichon Darienko & Pröschold — Another genus separated from the Dilabifilum (Ulvales) complex, Lithotrichon forms clustered cell packets as well as branching filaments and is distinguished from similar genera by SSU and ITS sequence data. The species L. pulchrum occurs as photobiont of the freshwater lichen Hydropunctaria rheitrophila (Darienko & Pröschold Reference Darienko and Pröschold2017).

Myrmecia Printz — These spherical to pyriform unicells have a parietal chloroplast, without a pyrenoid, extending around most of the cell, with 2–4 broad lobes defined by deep notches. Cells proliferate via zoospores, aplanospores, or autospores (Ettl & Gärtner Reference Ettl and Gärtner2014). Gene-based phylogenies consistently place Myrmecia in the Trebouxiales, sister to Trebouxia (Muggia et al. Reference Muggia, Nelsen, Kirika, Barreno, Beck, Lindgren, Lumbsch and Leavitt2020) or to the Asterochloris + Vulcanochloris clade (Vančurová et al. Reference Vančurová, Peksa, Němcová and Škaloud2015). Myrmecia occurs free-living as well as in lichen symbiosis. An aerophilic alga, originally described as Friedmannia from Negev Desert rocks, is now recognized as Myrmecia israelensis (Friedl Reference Friedl1995) and was recently reported as lichen photobiont (Thüs et al. Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011; Moya et al. Reference Moya, Chiva, Molins, Jadrná, Škaloud, Peksa and Barreno2018). Psora decipiens and a number of species in the Verrucariaceae are among the lichen-forming fungi known to partner with Myrmecia.

Nannochloris Naumann — The genus Nannochloris has encompassed simple, extremely tiny (1.5–2 μm) chlorophyte algae that reproduce by binary division or autospores. Circumscription of the genus has been controversial, but molecular data indicate that most of the species belong in Chlorellales (Henley et al. Reference Henley, Hironaka, Gillou, Buchheim, Buchheim, Fawley and Fawley2004). Tschermak-Woess (Reference Tschermak-Woess1981) recognized Nannochloris normandinae as the photobiont partner of lichen-forming Normandina pulchella; in other works, Nannochloris has been mentioned more indirectly in the context of photobionts (e.g. Lohtander et al. Reference Lohtander, Oksanen and Rikkinen2003). However, Thüs et al. (Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011) found only Diplosphaera as photobiont in the 10 Normandina thalli they examined and, more recently, Pröschold & Darienko (Reference Pröschold and Darienko2020) reduced Nannochloris normandinae to synonymy with Diplosphaera chodatii (Prasiolales). Thus, clear evidence of lichen photobionts belonging in Nannochloris appears to be lacking at present.

Neocystis F. Hindák — Members of this trebouxiophycean genus produce mucilaginous colonies of spherical to ellipsoidal or crescent-shaped cells that reproduce by autospores (Neustupa Reference Neustupa and Frey2015). Cultures assigned to Neocystis as well as other genera were recently reviewed with molecular sequence analyses, revealing considerable taxonomic redundancy assigned to only two closely related, genetically distinct but morphologically plastic species (Eliáš et al. Reference Eliáš, Neustupa, Pažoutová and Škaloud2013). An alga identified as Neocystis sp. was cited as ‘additional photobiont’ of Micarea misella, in thalli having Elliptochloris bilobata as principal photobiont (Voytsekhovich et al. Reference Voytsekhovich, Dymytrova and Nadyeina2011).

Paulbroadya Darienko & Pröschold — This recently recognized clade in the Ulvales is distinguished by SSU-ITS sequences from other taxa previously treated under Dilabifilum (Darienko & Pröschold Reference Darienko and Pröschold2017). The species Paulbroadya petersii occurs as photobiont of the marine intertidal lichen Wahlenbergiella mucosa (Darienko & Pröschold Reference Darienko and Pröschold2017).

Phycopeltis Millardet — Members of this trentepohliaceous genus are most often seen as coppery orange discs a few mm in diameter on leaf surfaces in humid subtropical and tropical regions, with one or two species extending to cooler regions such as oceanic Europe (Rindi et al. Reference Rindi, Menéndez, Guiry and Rico2004). Thallus discs consist of a monostromatic layer of closely appressed, bifurcating filaments (Fig. 6A). Unlike those of Cephaleuros, Phycopeltis thalli are supracuticular, non-pathogenic, and at least some species readily colonize other favourably displayed plant substrata besides leaves. Sporangia are usually borne erect on a very short stalk, and release quadriflagellate zoospores through a pore at the end opposite to the point of attachment. Gametangia are sessile and develop from intercalary compartments of the horizontal filament system in most species; gametes are biflagellate and isomorphic, and their fusion has been observed (Thompson & Wujek Reference Thompson and Wujek1997). The life cycle of Phycopeltis is believed to be haplodiplontic, with alternation of gametophytes and sporophytes that are isomorphic, rather than heteromorphic as in Cephaleuros and Stomatochroon (Thompson & Wujek Reference Thompson and Wujek1997). If this is the case, recognizing meiosporangia by the presence of tetrads might be the only means of distinguishing the phases phenotypically, but there do not appear to be such reports. Whether the gametophytes and sporophytes are equally susceptible to lichenization would be an interesting question to examine.

Although distinguishing Phycopeltis from Trentepohlia under current morphological concepts appears fairly straightforward, DNA sequence data show species of the two genera to be intertwined phylogenetically (Zhu et al. Reference Zhu, Zhao, Xia, Hu and Liu2015, Reference Zhu, Hu and Liu2017; Grube et al. Reference Grube, Muggia, Baloch, Hametner, Stocker-Wörgötter, Grube, Seckbach and Muggia2017a). Phycopeltis is particularly under-sampled at present. The morphological distinction between the two genera may also break down in the lichenized condition. Although Phycopeltis species can retain their placoid thallus characteristics when partnering with certain foliicolous mycobionts (Grube & Lücking Reference Grube and Lücking2002), in other lichens the algal filaments may be broken up into individual cells indistinguishable from those of Trentepohlia (see fig. 9; Lücking Reference Lücking2008). Using TEM, Matthews et al. (Reference Matthews, Tucker and Chapman1989) believed they could differentiate the two genera in such cases by features of the septal wall near plasmodesmata. It would be useful to test how well such traits correlate with molecular markers.

As widespread colonizers of the warm-temperate and tropical phyllosphere, species of Phycopeltis are important photobionts in foliicolous lichen communities, where they partner with diverse leaf-dwelling mycobionts including species of Arthonia, Chroodiscus, Mazosia, Opegrapha, Porina, Trichothelium, and supracuticular taxa of Strigulaceae, among others (Santesson Reference Santesson1952; Lücking Reference Lücking2008). Multiple Phycopeltis thalli may occur edge-to-edge within a single foliicolous lichen, as additional individuals are incorporated by the mycobiont's expanding prothallus (Sanders Reference Sanders2002). There are some reports of algal gametangia or sporangia being produced in the lichenized state, particularly in those taxa where the Phycopeltis thalli are sparsely covered by the mycobiont (Santesson Reference Santesson1952; Lücking Reference Lücking1994; Sanders Reference Sanders2002).

In at least one species of Phycopeltis (P. epiphyton), the highly degradation-resistant biopolymer sporopollenin was detected in the cell wall (Good & Chapman Reference Good and Chapman1978). Its presence in the walls of other photobionts (Coccomyxa and Elliptochloris) has been correlated with the absence of haustorial penetration by mycobionts (Honegger & Brunner Reference Honegger and Brunner1981; Brunner & Honegger Reference Brunner and Honegger1985). However, at least some strains of lichenized Phycopeltis may be deeply penetrated by mycobiont haustoria, such as those of Porina (Matthews et al. Reference Matthews, Tucker and Chapman1989).

Prasiola Meneghini — This trebouxiophycean seaweed of high-latitude supratidal zones is exceptional for its class in having a multicellular, macroscopic blade-like thallus. It is often abundant and readily visible both free-living and in symbiosis with mycobiont Mastodia tessellata (Verrucariaceae). Two or three distinct species of Prasiola appear to serve as photobiont to the bipolarly distributed Mastodia (Garrido-Benavent et al. Reference Garrido-Benavent, Pérez-Ortega and de los Ríos2017, Reference Garrido-Benavent, de los Ríos, Fernández-Mendoza and Pérez-Ortega2018). There has been some discussion, on structural grounds, as to whether this fungal-algal partnership ought to be considered a true lichen (Lud et al. Reference Lud, Huiskes and Ott2001; Kohlmeyer et al. Reference Kohlmeyer, Hawksworth and Volkmann-Kohlmeyer2004; Pérez-Ortega et al. Reference Pérez-Ortega, de los Ríos, Crespo and Sancho2010). There is no fungal cortex, nor does symbiosis substantially change algal thallus morphology; its anatomy, however, is significantly altered, as algal cells become well separated by a proliferation of encircling mycobiont hyphae (Kovačik & Batista Pereira Reference Kovačik and Batista Pereira2001; Lud et al. Reference Lud, Huiskes and Ott2001). The symbiosis therefore entails considerably more structural transformation than that produced by the marine fungus Mycophycias upon its seaweed host Ascophyllum (Xu et al. Reference Xu, Deckert and Garbary2008), or Turgidosculum upon Blidingia (Pérez-Ortega et al. Reference Pérez-Ortega, Miller and de los Ríos2018). From a phylogenetic perspective, it is worth noting that close relatives of both the mycobiont (Verrucariaceae) and the alga (Prasiolales) participate in symbioses that are unambiguously lichenic.

Pseudendoclonium Wille — These ulvalean algae have variably packet-forming to filamentous morphologies and may be differentiated into prostrate and erect filament systems. Darienko & Pröschold (Reference Darienko and Pröschold2017) moved into this genus a couple of photobionts previously treated under Dilabifilum, recognizing the photobiont of Arthopyrenia kelpii as Pseudendoclonium arthopyreniae, and the photobiont of Hydropunctaria maura as P. commune, which is also widespread as a free-living alga on intertidal rocks. Pseudendoclonium arthopyreniae has a pyrenoid surrounded by thick plates of starch and traversed by several narrow, thylakoid-derived membranes lacking pyrenoglobuli (Namba & Nakayama Reference Namba and Nakayama2021).

Pseudochlorella J. W. G. Lund — This unicellular genus of Chlorella-like algae is now placed in the Prasiolales. Molecular data support inclusion of the photobiont of at least one lichen-forming fungus, Trapelia coarctata (Darienko et al. Reference Darienko, Gustavs and Pröschold2016). Other reports attribute to Pseudochlorella the photobionts of certain Micarea, Placynthiella and Stereocaulon species (see Tschermak-Woess Reference Tschermak-Woess and Galun1988a; Voytsekhovich et al. Reference Voytsekhovich, Dymytrova and Nadyeina2011). However, molecular sequence studies have so far identified photobionts from Micarea as Coccomyxa and Elliptochloris, and those from the green algal layer of Stereocaulon as Asterochloris, Chloroidium and Vulcanochloris.

Pseudococcomyxa Korshikov — Isolates identified as Pseudococcomyxa simplex have been reported as photobionts of a maritime Leproloma sp. (Watanabe et al. Reference Watanabe, Nakano and Deguchi1997) and also Micarea prasina (Voytsekhovich et al. Reference Voytsekhovich, Dymytrova and Nadyeina2011), based on light microscopy and culture studies. The genus Pseudococcomyxa has been distinguished morphologically from Coccomyxa by the polarized secretion of mucilage to form a cap at one end of the cell. However, Darienko et al. (Reference Darienko, Gustavs, Eggert, Wolf and Pröschold2015) found this character to be culture-dependent, and the Pseudococcomyxa strains they analyzed phylogenetically appeared intermixed with those of Coccomyxa (see also Yahr et al. Reference Yahr, Florence, Škaloud and Voytsekhovich2015). Isolates attributed to P. simplex in particular occurred in several distinct clades. Darienko et al. (Reference Darienko, Gustavs, Eggert, Wolf and Pröschold2015) reassigned all these strains to Coccomyxa. While lichen photobiont isolates attributed to Pseudococcomyxa remain to be examined, support for distinction of the genus from Coccomyxa now appears to be lacking.

Pseudostichococcus L. Moewus — Morphologically similar to Stichococcus, this genus was recently revised with molecular data (Pröschold & Darienko Reference Pröschold and Darienko2020). It currently includes the photobiont partner of Neocatapyrenium rhizinosum (Hodač et al. Reference Hodač, Hallmann, Spitzer, Elster, Faßhauer, Brinkmann, Lepka, Diwan and Fried2016) in the Verrucariaceae.

Pseudotrebouxia P. A. Archibald — See Trebouxia.

Radiococcus Schmidle — Species of this genus have been reported to occur as principal photobiont in thalli of two species of Placynthiella (P. icmalea and P. uliginosa) from the Ukraine (Voytsekhovich et al. Reference Voytsekhovich, Dymytrova and Nadyeina2011). Corroboration with DNA sequence data is needed, particularly since diverse, unrelated taxa have been repeatedly ascribed to this genus in the past (Wolf et al. Reference Wolf, Hepperle and Krienitz2003). According to a recent taxonomic treatment, Radiococcaceae and Radiococcus belong in the order Sphaeropleales of the Chlorophyceae (Neustupa Reference Neustupa and Frey2015), although these names are still being applied to taxa falling in other groups, such as the Trebouxiophyceae (e.g. Metz et al. Reference Metz, Singer, Domaizon, Unrein and Lara2019).

Stichococcus Nägeli s. lat. — In its broad sense, Stichococcus (Prasiolales) has encompassed smallish unicellular to filamentous algae of notably labile morphology, the most commonly recognized form represented by short-cylindrical cells. The chloroplast is parietal, often extending to no more than half of the cell circumference, not markedly lobed, with or without a pyrenoid. Culture conditions appear to have a significant effect on cell form. The straight or slightly curved, rod-shaped cells may separate or remain together after division to form very short filaments or swell to more spherical shapes, and may or may not produce a surrounding gelatinous sheath (Ettl & Gärtner Reference Ettl and Gärtner2014). Quite a number of lichen photobionts have been ascribed to Stichococcus, but as their diversity is studied at the molecular level, these taxa are being placed in segregate genera or other prasiolalean clades. Some seven to nine clades have now been recognized within Stichococcus s. lat. (Hodač et al. Reference Hodač, Hallmann, Spitzer, Elster, Faßhauer, Brinkmann, Lepka, Diwan and Fried2016; Pröschold & Darienko Reference Pröschold and Darienko2020). All Stichococcus-like photobionts examined in a study of the Verrucariaceae by Thüs et al. (Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011), were shown to belong in Diplosphaera. Others now appear to fall within Pseudostichococcus, Deuterostichococcus or Tritostichococcus (Pröschold & Darienko Reference Pröschold and Darienko2020). It is not yet clear whether Stichococcus in the restricted sense (near to type species S. bacillaris) includes lichen photobionts.

Symbiochloris Škaloud et al. — Formally described by Škaloud et al. (Reference Škaloud, Friedl, Hallmann, Beck and Dal Grande2016), the genus corresponds to a distinct clade of Dictyochloropsis s. lat. previously recognized by Dal Grande et al. (Reference Dal Grande, Beck, Cornejo, Singh, Cheenacharoen, Nelsen and Scheidegger2014b). Symbiochloris is currently thought to include all lichen photobionts previously included in Dictyochloropsis, as well as some free-living taxa. Principal mycobiont partner genera are Lobaria, Pseudocyphellaria, Sticta and their recent segregates Crocodia, Dendriscosticta and Ricasolia, all members of the Lobariaceae. Other lichens reported harbouring Symbiochloris photobionts include species of Biatora, Brigantiaea, Chaenotheca, Megalospora and Phlyctis.

The net-like chloroplast of Symbiochloris, similar to that of Dictyochloropsis, has reticulations that vary in form, thickness and orientation according to species and developmental stage (Škaloud et al. Reference Škaloud, Neustupa, Radochová and Kubínová2005, Reference Škaloud, Friedl, Hallmann, Beck and Dal Grande2016). Lichenized populations reproduce by aplanospores, but zoospore production may be observed in isolated culture. Cells of free-living populations often attain much larger sizes and their surfaces may be covered with scales (Tschermak-Woess Reference Tschermak-Woess1995).

Trebouxia Puymaly — The principal crop of the alga-farming fungi, unicellular Trebouxia is thought to include the photobionts chosen by the largest proportion (nearly half) of known mycobiont species. Together with the closely related Asterochloris, Trebouxia is chlorobiont of most Lecanorales and Teloschistales, as well as many other taxa of the other species-rich lecanoromycetid orders (Miadlikowska et al. Reference Miadlikowska, Kauff, Högnabba, Oliver, Molnár, Fraker, Gaya, Hafellner, Hofstetter and Gueidan2014). Species of Trebouxia are spherical or occasionally ellipsoidal unicells with a variously lobed, axial chloroplast that fills much of the cell and bears a prominent pyrenoid (Fig. 2A & C). There is considerable diversity of pyrenoid ultrastructure within the genus, involving differences in the morphology of penetrating membranes and the distribution of starch deposits and pyrenoglobuli, when present (Peveling Reference Peveling1968, Reference Peveling1969; Fisher & Lang Reference Fisher and Lang1971; Friedl Reference Friedl1989). CO2-fixing Rubisco is concentrated in the pyrenoids, which in some instances also comprise additional, smaller, satellite substructures within the chloroplast (Ascaso et al. Reference Ascaso, Valladares and de los Ríos1995). Although pyrenoid types do not correspond precisely to the Trebouxia clades supported in molecular sequence analyses, and several are strikingly convergent with pyrenoids of distantly related algae, they can nonetheless be useful in distinguishing certain groupings of taxa at close range (Friedl Reference Friedl1989; but see also Muggia et al. (Reference Muggia, Zellnig, Rabensteiner and Grube2010)). Some 30 species of Trebouxia, distributed among four major clades, are currently recognized. However, this figure is believed to grossly underestimate the true genetic diversity present in the genus (Leavitt et al. Reference Leavitt, Kraichak, Vondrak, Nelsen, Altermann, Divakar, Alors, Esslinger, Crespo and Lumbsch2015; Muggia et al. Reference Muggia, Nelsen, Kirika, Barreno, Beck, Lindgren, Lumbsch and Leavitt2020). The boundaries among the formally described species remain largely unresolved, since much of the genetic diversity uncovered in recent studies is not fully congruent with the phenotypically defined taxa. Muggia et al. (Reference Muggia, Candotto-Carniel, Grube, Grube, Seckbach and Muggia2017) postulated that the application of a phylogenetic species concept would at least triple the number of species currently recognized in Trebouxia.

Two different groups were long distinguished within Trebouxia s. lat. (Ahmadjian Reference Ahmadjian1960), which was previously known as Cystococcus. Archibald (Reference Archibald1975) recognized two genera, Trebouxia and Pseudotrebouxia, based on differences in cell division which were judged sufficient to separate them into two distinct orders. However, Gärtner (Reference Gärtner1985) and Tschermak-Woess (Reference Tschermak-Woess1989) found Archibald's subdivision untenable and reunited the genus, while acknowledging that differences in cell division were present. Tschermak-Woess (Reference Tschermak-Woess1989) distinguished two subgenera: Trebouxia (corresponding roughly to Pseudotrebouxia), which forms aplanospores (or zoospores in culture) and also autospores, and Eleutherococcus (later Asterochloris), which produces aplanospores/zoospores but not autospores. Autospores are distinguishable from aplanospores in that they are produced in lower numbers and are tightly appressed together within the sporangium such that their walls form angular junctions between them (Tschermak-Woess Reference Tschermak-Woess1989). Another difference is the position of the chloroplast during cell division, which remains more or less central in Trebouxia but becomes parietal and flattened in Asterochloris (Ahmadjian Reference Ahmadjian1960; Tschermak-Woess Reference Tschermak-Woess1989). Molecular data firmly distinguish the two clades, which have been formally recognized as distinct genera for the past decade (Škaloud & Peksa Reference Škaloud and Peksa2010).

As a lichen symbiont, Trebouxia is abundant in a great diversity of habitats worldwide. It is said to be infrequently reported in the free-living state, although researchers who sample substrata with microscopy have often found it (Tschermak-Woess Reference Tschermak-Woess1978; Bubrick et al. Reference Bubrick, Galun and Frensdorff1984; Mukhtar et al. Reference Mukhtar, Garty and Galun1994; Sanders Reference Sanders2005; Handa et al. Reference Handa, Ohmura, Nakano and Nakahara-Tsubota2007; Uher Reference Uher2008; Neustupa & Štifterová Reference Neustupa and Štifterová2013), with the notable exception of Degelius (Reference Degelius1964). Clearly, the germinating spores of trebouxiophilic mycobionts manage to obtain it, often without needing to produce an extensive mycelium (Werner Reference Werner1931; Clayden Reference Clayden1998). Recent environmental sequencing studies have found Trebouxia on a variety of surfaces (Darienko et al. Reference Darienko, Gruber, Pröschold and Schagerl2013; Hallmann et al. Reference Hallmann, Stannek, Fritzlar, Hause-Reitner, Friedl and Hoppert2013, Reference Hallmann, Hoppert, Mudimu and Friedl2016; Yung et al. Reference Yung, Chan, Lacap, Pérez-Ortega, de los Ríos, Lee, Cary and Pointing2014) and well represented in soil, fresh water and even marine environments (Metz et al. Reference Metz, Singer, Domaizon, Unrein and Lara2019), although one cannot be certain that the detected sequences represent free-living individuals. By contrast, two other principal lichen photobiont genera, Trentepohlia and Nostoc, are uncontroversially well known in the free-living state. The comparison may not be fair, however, because Trentepohlia and Nostoc both form easily recognized macrocolonies (bright orange tufts and distinctive gelatinous globs, respectively) whereas Trebouxia cannot be distinguished without a microscope and some degree of effort. In any case, a shadow of doubt still seems to haunt the status of free-living Trebouxia populations, to judge from the cautious wording in even quite recent literature (e.g. Friedl & Büdel Reference Friedl, Büdel and Nash2008). Although he never claimed to have searched for it in nature, Ahmadjian (Reference Ahmadjian1988, Reference Ahmadjian1993, Reference Ahmadjian and Seckbach2001, Reference Ahmadjian2002) repeatedly affirmed that Trebouxia existed only in highly coevolved symbiosis with lichen fungi and did not occur free-living. Yet he acknowledged that aposymbiotic populations of Trebouxia could appear in nature. He even proposed, as have others, that they arose from the breakdown of lichenized propagules, such as soredia and isidia, that reach microhabitats unsuitable for the partners to develop symbiotically (Ahmadjian Reference Ahmadjian1988). Ahmadjian asserted, however, that such populations were not truly free-living, except in a ‘secondary sense’. Apparently, he meant that they were ephemeral rather than stably established, but stable or not, aposymbiotic populations of Trebouxia are likely to be significant. Like other micro-organisms, many algae take advantage of ephemeral resources and transiently favourable microenvironments, then complete their life cycles with sexual reproduction when conditions deteriorate. Some then survive as resistant spores; others may escape adversity by entering into lichen symbioses. Within a lichen thallus, an algal population may be perpetuated for many years, yet continually disperse via soredia, isidia, lichenized fragments and other propagules that can seed new free-living populations. This has been characterized as photobiont ‘escape’ from the lichen fungus (Werth Reference Werth and Fontaneto2010). It may be equally valid to view relichenization as photobiont escape from conditions that aposymbiotic populations might not long endure.

Although stages of flagellar development within a lichen thallus were reported (Slocum et al. Reference Slocum, Ahmadjian and Hildreth1980), authors have expressed scepticism that Trebouxia could produce motile or sexual cells in the symbiotic state (Tschermak-Woess Reference Tschermak-Woess1989), where all algal cells are held by one or more appressorial hyphae (Honegger Reference Honegger, Mendgen and Leseman1990). In aposymbiotic culture, by contrast, the production and release of Trebouxia zoids are well documented (Ahmadjian Reference Ahmadjian1960, Reference Ahmadjian1967; Tschermak-Woess Reference Tschermak-Woess1989; Takeshita Reference Takeshita2001). The huge genetic diversity present (Muggia et al. Reference Muggia, Nelsen, Kirika, Barreno, Beck, Lindgren, Lumbsch and Leavitt2020) and its structure within populations (Kroken & Taylor Reference Kroken and Taylor2000) suggest that Trebouxia is reproducing sexually, but virtually nothing is known about how or when the sexual cycle proceeds in nature. Although it is often said that sexual reproduction has not been observed in this genus, both Warén (Reference Warén1920) and Ahmadjian (Reference Ahmadjian1960) reported and illustrated the fusion of flagellate isogametes in Trebouxia cultures. However, Ahmadjian (Reference Ahmadjian1988, Reference Ahmadjian and Seckbach2001) believed that these features were vestiges of the alga's free-living ancestry that no longer play any role in their present life histories. Further investigation of aposymbiotic populations is needed, since considerable indirect evidence suggests that they may reveal key events in the Trebouxia life cycle.

Trentepohlia C. Martius — The filamentous taxa currently treated under this cosmopolitan genus are among the most familiar of subaerial algae, often forming readily visible yellowish orange tufts on bark, rocks and other substrata in a wide variety of environments. They are also among the phycobionts chosen by the most diverse lichen-forming ascomycetes, including members of the Arthoniomycetes, Coniocybomycetes, Dothidiomycetes, Eurotiomycetes (Pyrenulales) and ostropalean Lecanoromycetes such as the species-rich Graphidaceae. Members of the order Trentepohliales and its sole family Trentepohliaceae present a distinctive combination of features: phragmoplastic cell division with plasmodesmata (otherwise characteristic of charophycean algae), a uniquely structured flagellar apparatus, peculiar sporangiophores, and distinctive orange pigmentation. Consequently, widely divergent interpretations of their phylogenetic affinities have been proposed, with some authors even placing the group in a separate class of its own (van den Hoek et al. Reference van den Hoek, Mann and Jahns1995). However, rDNA sequence data firmly place the subaerial Trentepohliales among orders of mainly marine taxa within the Ulvophyceae (López-Bautista & Chapman Reference López-Bautista and Chapman2003; Leliaert et al. Reference Leliaert, Smith, Moreau, Herron, Verbruggen, Delwiche and De Clerck2012).

Among the taxa currently treated under Trentepohlia, a number of genera were described to accommodate the morphological diversity represented, most recently Printzina (Thompson & Wujek Reference Thompson and Wujek1992). However, DNA sequence analyses have so far shown that the phenotypic similarities recognized are unreliable indicators of phylogenetic affinity (López-Bautista et al. Reference López-Bautista, Rindi and Guiry2006; Rindi et al. Reference Rindi, Lam and López-Bautista2009). This also applies to some of the morphological traits currently used to distinguish Trentepohlia species from those of Phycopeltis. Free-living and lichenized isolates of Trentepohlia occur intermixed in gene-based phylogenies (Nelsen et al. Reference Nelsen, Rivas Plata, Andrew, Lücking and Lumbsch2011; Hametner et al. Reference Hametner, Stocker-Wörgötter and Grube2014a, Reference Hametner, Stocker-Wörgötter, Rindi and Grubeb; Kosecka et al. Reference Kosecka, Jabłońska, Flakus, Rodriguez-Flakus, Kukwa and Guzow-Krzemińska2020). Due to its visible and widespread presence in the free-living state, Trentepohlia is an excellent subject for studying the relationship between lichenized and aposymbiotic populations in nature (Fig. 1A–C). So far, however, the genus has been the focus of relatively few modern phylogenetic studies, despite its visibility and primary importance in lichen symbioses.

The pigments characteristic of the Trentepohliaceae, called ‘haematochrome’ in the older literature, are carotenoids that occur abundantly as lipidic globules in the cytoplasm. Some authors have attributed the colour to astaxanthin (Thompson & Wujek Reference Thompson and Wujek1997; Grube et al. Reference Grube, Muggia, Baloch, Hametner, Stocker-Wörgötter, Grube, Seckbach and Muggia2017a), a deep red carotenoid known from other chlorophytes such as Haematococcus and the snow alga Chlamydomonas nivalis. However, analyses of Trentepohlia haematochrome show principally beta-carotenes (Czeczuga & Maximov Reference Czeczuga and Maximov1996; Mukherjee et al. Reference Mukherjee, Borah and Goswami2010; Kharkongor & Ramanujam Reference Kharkongor and Ramanujam2015; Chen et al. Reference Chen, Zhang and Liu2016). Located outside the plastids, these secondary carotenoids do not participate in photosynthetic light-harvesting but are hypothesized to filter excess light and suppress any damaging reactive oxygen species thereby generated (Solovchenko Reference Solovchenko2013). This may contribute to the visible success of Trentepohlia in colonizing exposed substrata in diverse environments.

The sporangia of Trentepohlia are themselves units of dispersal, easily detached when mature and vectored by wind, rain or insects. They then initiate a second round of shorter-distance dispersal by releasing quadriflagellate zoospores (Thompson & Wujek Reference Thompson and Wujek1997). Trentepohlia also produces putative gametangia that are morphologically distinct from sporangia. However, the biflagellate zoïds released have most often been observed germinating as spores rather than fusing as gametes (Rindi & Guiry Reference Rindi and Guiry2002).

Cellular contacts between mycobionts and their trentepohliaceous photobionts often appear to be superficial. However, most of these lichens, when carefully examined microscopically, reveal haustorial penetration, often deeply into the algal cells (Tschermak Reference Tschermak1941a; Withrow & Ahmadjian Reference Withrow and Ahmadjian1983; Matthews et al. Reference Matthews, Tucker and Chapman1989; but see Meier & Chapman Reference Meier and Chapman1983).

Trentepohlia photobionts have been occasionally reported to grow out from the algal layer and emerge as free filaments projecting from the lichen thallus or thalline apothecial margin (Zahlbruckner Reference Zahlbruckner, Engler and Prantl1907: p. 126; McGee Reference McGee2002). In one case, such a filament was seen bearing a sporangium (Tschermak Reference Tschermak1941a: p. 289). However, some authors have expressed scepticism about this interpretation, suggesting that epiphytic Trentepohlia might instead develop upon, and then grow into, an already formed thallus (Henssen & Jahns Reference Henssen and Jahns1974: p. 196). More detailed observations are clearly required, but either explanation could represent another potentially significant mechanism by which exchange may occur between lichenized and free-living populations.

Tritostichococcus Pröschold & Darienko — This recent segregate of Stichococcus s. lat. (Pröschold & Darienko Reference Pröschold and Darienko2020) includes the Stichococcus-like photobionts that associate with Chaenotheca, a genus of lichen-forming fungi that partner with a remarkably broad spectrum of photobionts (Tibell Reference Tibell2001).

Trochiscia Kützing — This genus of unicellar algae is characterized by an often-thick cell wall with spine- or wart-like projections, an irregularly stellate chloroplast, and two forms of endogenous spore production, resulting in hundreds of small cylindrical autospores or just two rounded endospores (Tschermak Reference Tschermak1941b). Trochiscia currently appears to be placed among the Chlorophyceae, in or near Sphaeropleales (Fučíková et al. Reference Fučíková, Lewis, Neupane, Karol and Lewis2019). It was identified as photobiont of Polyblastia amota and P. hyperborea (Tschermak Reference Tschermak1941b; Ahmadjian Reference Ahmadjian1967) in the Verrucariaceae, but those reports appear to be in doubt (Ettl & Gärtner Reference Ettl and Gärtner2014) and further studies are needed. Trochiscia was not among the photobionts detected in the survey by Thüs et al. (Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011) of the algal partners of Verrucariaceae.

Vulcanochloris Vancurová et al. — This newest addition to the Trebouxia family encompasses three recently described species with a distinctive, highly dissected chloroplast structure, and molecular sequences that place them as sister to Asterochloris (Vančurová et al. Reference Vančurová, Peksa, Němcová and Škaloud2015). They are known mainly as principal photobionts from some thalli of Stereocaulon vesuvianum in the Canary Islands, although there is also a very recent report of Vulcanochloris from a Stereocaulon thallus collected in highland Bolivia (Kosecka et al. Reference Kosecka, Guzow-Krzemińska, Čemajová, Škaloud, Jabłońska and Kukwa2021). Most other Stereocaulon lineages investigated to date appear to associate with Asterochloris or Chloroidium (Vančurová et al. Reference Vančurová, Muggia, Peksa, Řídká and Škaloud2018). Vulcanochloris has also been recently reported as a minority photobiont in thalli of Ramalina farinacea (Moya et al. Reference Moya, Molins, Martínez-Alberola, Muggia and Barreno2017).

Stramenopila (Heterokontae)

Heterococcus Chodat — The yellow-green (xanthophyte) algae lack fucoxanthin, the golden brownish plastidial carotenoid otherwise characteristic of the photosynthetic stramenopiles. The absence of fucoxanthin makes them rather easy to confuse with green algae. Their zoospores, however, will have the characteristic stramenopilous flagellum bearing stiff, hollow, tripartite appendages (mastigonemes). Heterococcus forms irregular filaments and/or cell clusters when isolated into culture (Zeitler Reference Zeitler1954). Molecular sequences support the light microscope identification of Heterococcus as photobiont in thalli of three species of Verrucariaceae (Hydropunctaria rheitrophila, Verrucaria funckii and V. hydrela) that are each in separate clades and not closely related to one another (Thüs et al. Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube and Friedl2011). Another xanthophyte, Heterothrix (now Xanthonema; Silva Reference Silva1979) was identified via light microscopy as photobiont of Staurothele clopimoides (Pereira Riquelme Reference Pereira Riquelme1992) but that interesting report requires corroboration.

Petroderma Kuckuck — Petroderma maculiforme is a small crustose brown alga (Phaeophyceae) found on rocks in the lower intertidal zone of western North America and Europe. In San Francisco Bay, it is particularly common on discarded plastic (Sanders et al. Reference Sanders, Moe and Ascaso2004). The alga is a disc of tightly branched, radiating horizontal filaments, rather similar in morphology to the chlorophyte Phycopeltis but with a dense carpet of short, erect filaments arising proximally from the horizontal system. In the free-living state, these erect filaments may bear unilocular and/or plurilocular sporangia (zoidangia) terminally (Fritsch Reference Fritsch1945). Chloroplasts typically possess one or several large pyrenoids that are prominent in electron micrographs (Fig. 2D) but not readily visible with light microscopy. The pyrenoids are traversed by branching tubules arising from invagination of the plastidial boundary membranes (rather than thylakoids, as in Trebouxia and Heveochlorella); the position of the pyrenoid may be laminar, protruding to exserted, or enfolded by chloroplast lobes (Sanders et al. Reference Sanders, Moe and Ascaso2005). The alga was first brought to the attention of lichenologists by a footnote in a phycology dissertation (Wynne Reference Wynne1969) that reported it in symbiosis with a Verrucaria species on intertidal rocks in northern California. The lichen was not studied further until Moe (Reference Moe1997) re-collected it and formally described the fungal symbiont as Verrucaria tavaresiae (now Wahlenbergiella tavaresiae; Gueidan et al. Reference Gueidan, Thüs and Pérez-Ortega2011). When lichenized, the Petroderma filaments are separated by fungal tissue, into which they grow and branch downwards rather than upwards as in the free-living condition (Sanders et al. Reference Sanders, Moe and Ascaso2004). Petroderma is the only member of the Phaeophyceae known to enter into lichen symbiosis. However, certain larger brown seaweeds, such as Ascophyllum, have intimate, mutualistic associations with verrucariacean fungi (Garbary & London Reference Garbary and London1995; Garbary & MacDonald Reference Garbary and MacDonald1995) that are generally not considered to be lichens on structural grounds, since the fungus grows within algal tissues as a conventional mycelium (Hawksworth Reference Hawksworth1988).

Acknowledgements

The manuscript benefited from critical review by Dr Richard L. Moe and two anonymous referees, and the expert scrutiny of the journal production staff. Funding for colour in print was kindly provided by the College of Arts and Sciences and the Department of Biological Sciences, Florida Gulf Coast University.

Author ORCID

William B. Sanders, 0000-0001-9572-4244.

References

Adams, DG and Duggan, PS (2002) Cyanobacteria-bryophyte symbioses. Journal of Experimental Botany 59, 10471058.CrossRefGoogle Scholar
Adams, DG, Duggan, PS and Jackson, O (2012) Cyanobacterial symbiosis. In Whitton, BA (ed.), Ecology of Cyanobacteria II: Their Diversity in Space and Time. Heidelberg: Springer, pp. 593647.CrossRefGoogle Scholar
Ahmadjian, V (1960) Some new and interesting species of Trebouxia, a lichenized alga. American Journal of Botany 47, 677683.CrossRefGoogle Scholar
Ahmadjian, V (1962) Investigations on lichen synthesis. American Journal of Botany 49, 277283.CrossRefGoogle Scholar
Ahmadjian, V (1967) A guide to the algae occurring as lichen symbionts: isolation, culture, cultural physiology, and identification. Phycologia 6, 127160.CrossRefGoogle Scholar
Ahmadjian, V (1982) Holobionts have more parts. International Association for Lichenology Newsletter 15, 19.Google Scholar
Ahmadjian, V (1987) Coevolution in lichens. Annals of the New York Academy of Sciences 503, 307315.CrossRefGoogle Scholar
Ahmadjian, V (1988) The lichen alga Trebouxia: does it occur free-living? Plant Systematics and Evolution 158, 243247.CrossRefGoogle Scholar
Ahmadjian, V (1993) The Lichen Symbiosis. New York: John Wiley and Sons.Google Scholar
Ahmadjian, V (1995) Lichens are more important than you think. BioScience 45, 124.CrossRefGoogle Scholar
Ahmadjian, V (2001) Trebouxia: reflections on a perplexing and controversial lichen photobiont. In Seckbach, J (ed.), Symbiosis: Mechanisms and Model Systems. Dordrecht: Kluwer Academic Publishers, pp. 373383.Google Scholar
Ahmadjian, V (2002) Lingering lichen myths are hard to dispel. ISS Symbiosis International 2, 12.Google Scholar
Ahmadjian, V and Heikkilä, H (1970) The culture and synthesis of Endocarpon pusillum and Staurothele clopima. Lichenologist 4, 259267.CrossRefGoogle Scholar
Ahmadjian, A and Jacobs, JB (1981) Relationship between fungus and alga in the lichen Cladonia cristatella Tuck. Nature 289, 169172.CrossRefGoogle Scholar
Ahmadjian, A, Russell, LA and Hildreth, KC (1980) Artificial reestablishment of lichens. I. Morphological interactions between the phycobionts of different lichens and the mycobionts Cladonia cristatella and Lecanora chrysoleuca. Mycologia 72, 7389.CrossRefGoogle Scholar
Aoki, M, Nakano, T, Kanda, H and Deguchi, H (1998) Photobionts isolated from Antarctic lichens. Journal of Marine Biotechnology 6, 3943.Google Scholar
Archibald, PA (1975) Trebouxia de Puymaly (Chlorophyceae, Chlorococcales) and Pseudotrebouxia gen. nov. (Chlorophyceae, Chlorosarcinales). Phycologia 14, 125137.CrossRefGoogle Scholar
Armaleo, D and Clerc, P (1991) Lichen chimeras: DNA analysis suggests that one fungus forms two morphotypes. Experimental Mycology 15, 110.CrossRefGoogle Scholar
Armaleo, D, Müller, O, Lutzoni, F, Andrésson, ÓS, Blanc, G, Bode, HB, Collart, FR, Dal Grande, F, Dietrich, F, Grigoriev, IV, et al. (2019) The lichen symbiosis re-viewed through the genomes of Cladonia grayi and its algal partner Asterochloris glomerata. BMC Genomics 20, 605.CrossRefGoogle ScholarPubMed
Ascaso, C, Valladares, F and de los Ríos, A (1995) New ultrastructural aspect of pyrenoids of the lichen photobiont Trebouxia (Microthamniales, Chlorophyta). Journal of Phycology 31, 114119.CrossRefGoogle Scholar
Asplund, J and Wardle, DA (2013) The impact of secondary compounds and functional characteristics on lichen palatability and decomposition. Journal of Ecology 101, 689700.CrossRefGoogle Scholar
Bačkor, M, Peksa, O, Škaloud, P and Bačkorová, M (2010) Photobiont diversity in lichens from metal-rich substrata based on ITS rDNA sequences. Ecotoxicology and Environmental Safety 73, 603612.CrossRefGoogle ScholarPubMed
Bates, ST, Cropsey, GWG, Caporaso, JG, Knight, R and Fierer, N (2011) Bacterial communities associated with the lichen symbiosis. Applied and Environmental Microbiology 77, 13091314.CrossRefGoogle ScholarPubMed
Beck, A (1999) Photobiont inventory of a lichen community growing on heavy-metal-rich rock. Lichenologist 31, 501510.CrossRefGoogle Scholar
Beck, A (2002) Selektivität der symbionten schwermetalltoleranter flechten. Ph.D. thesis, Ludwig Maximilian University of Munich.Google Scholar
Beck, A and Koop, H-U (2001) Analysis of the photobiont population in lichens using a single-cell manipulator. Symbiosis 31, 5767.Google Scholar
Beck, A and Mayr, C (2012) Nitrogen and carbon isotope variability in the green-algal lichen Xanthoria parietina and their implications on mycobiont–photobiont interactions. Ecology and Evolution 2, 31323144.CrossRefGoogle ScholarPubMed
Beck, A, Friedl, T and Rambold, G (1998) Selectivity of photobiont choice in a defined lichen community: inferences from cultural and molecular studies. New Phytologist 139, 709720.CrossRefGoogle Scholar
Beck, A, Kasalicky, T and Rambold, G (2002) Myco-photobiontal selection in a Mediterranean cryptogam community with Fulgensia fulgida. New Phytologist 153, 317326.CrossRefGoogle Scholar
Beck, A, Bechteler, J, Casanova-Katny, A and Dzhilyanova, I (2019) The pioneer lichen Placopsis in maritime Antarctica: genetic diversity of their mycobionts and green algal symbionts, and their correlation with deglaciation time. Symbiosis 79, 124.CrossRefGoogle Scholar
Beckett, RP, Solhaug, KA, Gauslaa, Y and Minibayeva, F (2019) Improved photoprotection in melanized lichens is a result of fungal solar radiation screening rather than photobiont acclimation. Lichenologist 51, 483491.CrossRefGoogle Scholar
Bergman, B and Hällbom, L (1981) Nostoc of Peltigera canina when lichenized and isolated. Canadian Journal of Botany 60, 20922098.CrossRefGoogle Scholar
Bergman, B, Johanssen, C and Söderbäck, E (1992) The Nostoc-Gunnera symbiosis (Tansley Review No. 42). New Phytologist 122, 379400.CrossRefGoogle Scholar
Bertsch, A and Butin, H (1967) Die Kultur der Erdflechte Endocarpon pusillum im Labor. Planta 72, 2942.CrossRefGoogle Scholar
Bhâradwâja, Y (1933) False branching and sheath structure in the Myxophyceae, with special reference to the Scytonemataceae. Archiv für Protistenkunde 81, 243283.Google Scholar
Bitter, G (1904) Peltigeren-Studien II. Das Verhalten der oberseitigen Thallusschuppen der Peltigera lepidophora (Nyl). Berichte der Deutschen Botanischen Gesellschaft 22, 251254.Google Scholar
Blaha, J, Baloch, E and Grube, M (2006) High photobiont diversity associated with the euryoecious lichen-forming ascomycete Lecanora rupicola (Lecanoraceae, Ascomycota). Biological Journal of the Linnean Society 88, 283293.CrossRefGoogle Scholar
Boch, S, Prati, D, Werth, S, Rüetschi, J and Fischer, M (2011) Lichen endozoochory by snails. PLoS ONE 6, e18770.CrossRefGoogle ScholarPubMed
Bock, C, Krienitz, L and Pröschold, T (2011) Taxonomic reassessment of the genus Chlorella (Trebouxiophyceae) using molecular signatures (barcodes), including description of seven new species. Fottea 11, 293312.CrossRefGoogle Scholar
Boissière, J-C, Boissière, M-C, Champion-Arnaud, P, Lallmant, R and Wagner, J (1987) Le cycle des Nostoc des genres Peltigera et Collema en cultures in vitro et dans le thalle liquénique. Canadian Journal of Botany 65, 14681477.CrossRefGoogle Scholar
Bold, HC and Wynne, M (1985) Introduction to the Algae, 2nd Edn. Englewood Cliffs, New Jersey: Prentice Hall.Google Scholar
Bornet, E (1873) Recherches sur les gonidies des algues. Annales des Sciences Naturelles (5me Série, Botanique) 17, 45110.Google Scholar
Brito, A, Vieira, J, Vieira, CP, Zhu, T, Leão, PN, Ramos, V, Lu, X, Vasconcelos, VM, Gugger, M and Tamagnini, P (2020) Comparative genomics discloses the uniqueness and the biosynthetic potential of the marine cyanobacterium Hyella patelloides. Frontiers in Microbiology 11, 1527.CrossRefGoogle ScholarPubMed
Brodo, IM and Richardson, DHS (1978) Chimeroid association in the genus Peltigera. Lichenologist 10, 157170.CrossRefGoogle Scholar
Brodo, IM, Sharnoff, SD and Sharnoff, S (2001) Lichens of North America. New Haven and London: Yale University Press.Google Scholar
Brooks, F, Rindi, F, Suto, S, Ohtani, S and Green, M (2015) The Trentepohliales (Ulvophyceae, Chlorophyta): an unusual algal order and its novel plant pathogen, Cephaleuros. Plant Disease 99, 740753.CrossRefGoogle ScholarPubMed
Brunner, U and Honegger, R (1985) Chemical and ultrastructural studies on the distribution of sporopollenin-like biopolymers in six genera of lichen phycobionts. Canadian Journal of Botany 63, 22212230.CrossRefGoogle Scholar
Bubrick, P (1988) Effects of symbiosis on the photobiont. In Galun, M (ed.), CRC Handbook of Lichenology. Boca Raton, Florida: CRC Press, pp. 133144.Google Scholar
Bubrick, P, Galun, M and Frensdorff, A (1984) Observations on free-living Trebouxia de Puymaly and Pseudotrebouxia Archibald, and evidence that both symbionts from Xanthoria parietina (L.) Th. Fr. can be found free-living in nature. New Phytologist 97, 455462.CrossRefGoogle Scholar
Büdel, B (1985) Blue-green phycobionts in the lichen family Lichinaceae. Archiv für Hydrobiologie, Supplement 71 [Algological Studies 38/39], 355357.Google Scholar
Büdel, B (1999) Ecology and diversity of rock-inhabiting cyanobacteria in tropical regions. European Journal of Phycology 34, 361370.CrossRefGoogle Scholar
Büdel, B and Henssen, A (1983) Chroococcidiopsis (Cyanophyceae), a phycobiont in the lichen family Lichinaceae. Phycologia 22, 367375.CrossRefGoogle Scholar
Büdel, B and Henssen, A (1987) Trebouxia aggregata und Gloeocapsa sanguinea, phycobionten in Euopsis granatina (Lichinaceae). Plant Systematics and Evolution 158, 235241.CrossRefGoogle Scholar
Büdel, B and Kauff, F (2012) Blue-green algae. In Frey, G (series ed.), Syllabus of Plant Families: Adolf Engler's Syllabus der Pflanzenfamilien, Part VI. Stuttgart: Borntraeger Verlagsbuchhandlung, pp. 539.Google Scholar
Butler, RD and Allsopp, A (1972) Ultrastructural investigations in the Stigonemataceae (Cyanophyta). Archiv für Mikrobiologie 82, 283299.CrossRefGoogle Scholar
Canals, A, Hernández-Mariné, M, Gómez-Bolea, A and Llimona, X (1997) Botryolepraria, a new monotypic genus segregated from Lepraria. Lichenologist 29, 339345.CrossRefGoogle Scholar
Candotto, Carniel F, Zanelli, D, Bertuzzi, S and Tretiach, M (2015) Desiccation tolerance and lichenization: a case study with the aeroterrestrial microalga Trebouxia sp. (Chlorophyta). Planta 242, 493505.CrossRefGoogle Scholar
Cao, S, Zhang, F, Liu, C, Hao, Z, Tian, Y, Zhu, L and Zhou, Q (2015) Distribution patterns of haplotypes for symbionts from Umbilicaria esculenta and U. muehlenbergii reflect the importance of reproductive strategy in shaping population genetic structure. BMC Microbiology 15, 212.CrossRefGoogle ScholarPubMed
Cao, S, Zhang, F, Zheng, H, Peng, F, Liu, C and Zhou, Q (2018) Coccomyxa greatwallensis sp. nov. (Trebouxiophyceae, Chlorophyta), a lichen epiphytic alga from Fildes Peninsula, Antarctica. PhytoKeys 110, 3950.CrossRefGoogle Scholar
Cardós, JLH, Prieto, M, Jylhä, M, Aragón, G, Molina, MC, Martínez, I and Rikkinen, J (2019) A case study on the re-establishment of the cyanolichen symbiosis: where do the compatible photobionts come from? Annals of Botany 124, 379388.CrossRefGoogle Scholar
Casano, LM, del Campo, E, García-Breijo, FJ, Reig-Armiñana, J, Gasulla, F, del Hoyo, A, Guéra, A and Barreno, E (2011) Two Trebouxia algae with different physiological performances are ever-present in lichen thalli of Ramalina farinacea. Coexistence versus competition? Environmental Microbiology 13, 806818.CrossRefGoogle ScholarPubMed
Catalá, S, del Campo, EM, Barreno, E, García-Breijo, FJ, Reig-Armiñana, J and Casano, LM (2016) Coordinated ultrastructural and phylogenomic analyses shed light on the hidden phycobiont diversity of Trebouxia microalgae in Ramalina fraxinea. Molecular Phylogenetics and Evolution 94, 765777.CrossRefGoogle ScholarPubMed
Chen, L, Zhang, L and Liu, T (2016) Concurrent production of carotenoids and lipid by a filamentous microalga Trentepohlia arborum. Bioresource Technology 214, 567573.CrossRefGoogle ScholarPubMed
Chodat, R (1913) Monographie d'algues en culture pure. Matériaux por la Flore Cryptogamique Suisse [Beiträge zur Kryptogamenflora der Schweiz] 4, 1226.Google Scholar
Clayden, SR (1998) Thallus initiation and development in the lichen Rhizocarpon lecanorinum. New Phytologist 139, 685695.CrossRefGoogle Scholar
Collins, CR and Farrar, JF (1978) Structural resistances to mass transfer in the lichen Xanthoria parietina. New Phytologist 81, 7183.CrossRefGoogle Scholar
Coppins, BJ (1983) A taxonomic study of the lichen genus Micarea in Europe. Bulletin of the British Museum (Natural History), Botany Series 11, 17214.Google Scholar
Cordeiro, LM, Reis, RA, Cruz, LM, Stocker-Wörgötter, E, Grube, M and Iacomini, M (2005) Molecular studies of photobionts of selected lichens from the coastal vegetation of Brazil. FEMS Microbiology Ecology 54, 381390.CrossRefGoogle ScholarPubMed
Cornejo, C and Scheidegger, C (2013) New morphological aspects of cephalodium formation in the lichen Lobaria pulmonaria (Lecanorales Ascomycota). Lichenologist 45, 7787.CrossRefGoogle Scholar
Cornejo, C and Scheidegger, C (2016) Cyanobacterial gardens: the liverwort Frullania asagrayana acts as a reservoir of lichen photobionts. Environmental Microbiology Reports 8, 352357.CrossRefGoogle ScholarPubMed
Cornejo, C, Nelson, PR, Stepanchikova, I, Himelbrant, D, Jørgensen, P-M and Scheidegger, C (2016) Contrasting pattern of photobiont diversity in the Atlantic and Pacific populations of Erioderma pedicellatum (Pannariaceae). Lichenologist 48, 275291.CrossRefGoogle Scholar
Costa, J-L and Lindblad, P (2002) Cyanobacteria in symbiosis with cycads. In Rai, AN, Bergman, B and Rasmussen, U (eds), Cyanobacteria in Symbiosis. Dordrecht: Kluwer Academic Publishers, pp. 195205.Google Scholar
Crittenden, P, David, JC, Hawksworth, DL and Campbell, FS (1995) Attempted isolation and success in the culturing of a broad spectrum of lichen-forming and lichenicolous fungi. New Phytologist 130, 267297.CrossRefGoogle Scholar
Czeczuga, B and Maximov, OM (1996) Carotenoids in the cells of the alga Trentepohlia gobii Meyer. Acta Societatis Botanicorum Poloniae 65, 273276.CrossRefGoogle Scholar
Dahlkild, A, Kallersjö, M, Lohtander, K and Tehler, A (2001) Photobiont diversity in the Physciaceae (Lecanorales). Bryologist 104, 527536.CrossRefGoogle Scholar
Dal Forno, M, Lawrey, JD, Sikaroodi, M, Gillevet, PM, Schuettpelz, E and Lücking, R (2020) Extensive photobiont sharing in a rapidly radiating cyanolichen clade. Molecular Ecology 30, 17551776.CrossRefGoogle Scholar
Dal Grande, F, Widmer, I, Wagner, HH and Scheidegger, C (2012) Vertical and horizontal photobiont transmission within populations of a lichen symbiosis. Molecular Ecology 21, 31593172.CrossRefGoogle ScholarPubMed
Dal Grande, F, Alors, D, Divakar, PK, Bálint, M, Crespo, A and Schmitt, I (2014 a) Insights into intrathalline genetic diversity of the cosmopolitan lichen symbiotic green alga Trebouxia decolorans Ahmadjian using microsatellite markers. Molecular Phylogenetics and Evolution 72, 5460.CrossRefGoogle ScholarPubMed
Dal Grande, F, Beck, A, Cornejo, C, Singh, G, Cheenacharoen, S, Nelsen, MP and Scheidegger, C (2014 b) Molecular phylogeny and symbiotic selectivity of the green algal genus Dictyochloropsis s. l. (Trebouxiophyceae): a polyphyletic and widespread group forming photobiont-mediated guilds in the lichen family Lobariaceae. New Phytologist 202, 455470.CrossRefGoogle Scholar
Dal Grande, F, Rolshausen, G, Divakar, PK, Crespo, A, Otte, J, Schleuning, M and Schmitt, I (2018) Environment and host identity structure communities of green algal symbionts in lichens. New Phytologist 217, 277289.CrossRefGoogle ScholarPubMed
Darienko, T and Pröschold, T (2015) Genetic variability and taxonomic revision of the genus Auxenochlorella (Shihira et Krauss) Kalina et Puncocharova (Trebouxiophyceae, Chlorophyta). Journal of Phycology 51, 394400.CrossRefGoogle Scholar
Darienko, T and Pröschold, T (2017) Toward a monograph of non-marine Ulvophyceae using an integrative approach (Molecular phylogeny and systematics of terrestrial Ulvophyceae II.). Phytotaxa 324, 141.CrossRefGoogle Scholar
Darienko, T and Pröschold, T (2019) The genus Jaagichlorella Reisigl (Trebouxiophyceae, Chlorophyta) and its close relatives: an evolutionary puzzle. Phytotaxa 388, 4768.CrossRefGoogle Scholar
Darienko, T, Gustavs, L, Mudimu, O, Menendez, CR, Schumann, R, Karsten, U, Friedl, T and Pröschold, T (2010) Chloroidium, a common terrestrial coccoid green alga previously assigned to Chlorella (Trebouxiophyceae, Chlorophyta). European Journal of Phycology 45, 7995.CrossRefGoogle Scholar
Darienko, T, Gruber, M, Pröschold, T and Schagerl, M (2013) Terrestrial microalgae on Viennese buildings. Final report of project H-2081/2010. Vienna: Universität Wien.Google Scholar
Darienko, T, Gustavs, L, Eggert, A, Wolf, W and Pröschold, T (2015) Evaluating the species boundaries of green microalgae (Coccomyxa, Trebouxiophyceae, Chlorophyta) using integrative taxonomy and DNA barcoding with further implications for the species identification in environmental samples. PLoS ONE 10, e0127838.CrossRefGoogle ScholarPubMed
Darienko, T, Gustavs, L and Pröschold, T (2016) Species concept and nomenclatural changes within the genera Elliptochloris and Pseudochlorella (Trebouxiophyceae) based on an integrative approach. Journal of Phycology 52, 11251145.CrossRefGoogle Scholar
Darienko, T, Lukešová, A and Pröschold, T (2018) The polyphasic approach revealed new species of Chloroidium (Trebouxiophyceae, Chlorophyta). Phytotaxa 372, 5166.CrossRefGoogle Scholar
de los Ríos, A, Sancho, LG, Grube, M, Wierzchos, J and Ascaso, C (2005) Endolithic growth of two Lecidea lichens in granite from continental Antarctica detected by molecular and microscopy techniques. New Phytologist 165, 181190.CrossRefGoogle ScholarPubMed
de los Ríos, A, Raggio, J, Pérez-Ortega, S, Vivas, M, Pintado, A, Green, TGA, Ascaso, C and Sancho, LG (2011) Anatomical, morphological and ecophysiological strategies in Placopsis pycnotheca (lichenized fungi, Ascomycota) allowing rapid colonization of recently deglaciated soils. Flora 206, 857864.CrossRefGoogle Scholar
Degelius, G (1964) Biological studies of the epiphytic vegetation on twigs of Fraxinus excelsior. Acta Horti Gotoburgensis 27, 1155.Google Scholar
Delwiche, CF (1999) Tracing the thread of plastid diversity through the tapestry of time. American Naturalist 154, S164S177.CrossRefGoogle Scholar
Döbbler, P (1984) Symbiosen zwischen Gallertalgen und Gallertpilzen der Gattung Epigloea (Ascomycetes). Beiheft zur Nova Hedwigia 79, 203239.Google Scholar
Dodds, WK, Gudder, DA and Mollenhauer, D (1995) The ecology of Nostoc. Journal of Phycology 31, 218.CrossRefGoogle Scholar
Doering, JA, Booth, T, Wiersma, YF and Piercey-Normore, MD (2020) How do genes flow? Identifying potential dispersal mode for the semi-aquatic lichen Dermatocarpon luridum using spatial modelling and photobiont markers. BMC Ecology 20, 56.CrossRefGoogle ScholarPubMed
Doering, M and Piercey-Normore, MD (2009) Genetically divergent algae shape an epiphytic lichen community on Jack Pine in Manitoba. Lichenologist 41, 6980.CrossRefGoogle Scholar
Drew, EA and Smith, DC (1967) Studies in the physiology of lichens VII. The physiology of the Nostoc symbiont of Peltigera polydactyla compared with cultured and free-living forms. New Phytologist 66, 379388.CrossRefGoogle Scholar
Eliáš, M, Neustupa, J, Pažoutová, M and Škaloud, P (2013) A case of taxonomic inflation in coccoid algae: Ellipsoidion parvum and Neocystis vischeri are conspecific with Neocystis (=Nephrodiella) brevis (Chlorophyta, Trebouxiophyceae). Phytotaxa 76, 1527.CrossRefGoogle Scholar
Elshobary, ME, Osman, MEH, Abushady, AM and Piercey-Normore, MD (2015) Comparison of lichen-forming cyanobacterial and green algal photobionts with free-living algae. Cryptogamie, Algologie 36, 81100.CrossRefGoogle Scholar
Elvebakk, A, Papaefthimiou, D, Robertsen, EH and Liaimer, A (2008) Phylogenetic patterns among Nostoc cyanobionts within bi- and tripartite lichens of the genus Pannaria. Journal of Phycology 44, 10491059.CrossRefGoogle ScholarPubMed
Engelen, A, Convey, P and Ott, S (2010) Life history strategy of Lepraria borealis at an Antarctic inland site, Coal Nunatak. Lichenologist 42, 339346.CrossRefGoogle Scholar
Engelen, A, Convey, P, Popa, O and Ott, S (2016) Lichen photobiont diversity and selectivity at the southern limit of the maritime Antarctic region (Coal Nunatak, Alexander Island). Polar Biology 39, 24032410.CrossRefGoogle Scholar
Ertz, D, Guzow-Krzemińska, B, Thor, G, Łubek, A and Kukwa, M (2018) Photobiont switching causes changes in the reproduction strategy and phenotypic dimorphism in the Arthoniomycetes. Scientific Reports 8, 4952.CrossRefGoogle ScholarPubMed
Ettl, H and Gärtner, G (2014) Syllabus der Boden-, Luft-, und Flechtenalgen, 2nd Edn. Berlin and Heidelberg: Springer Spektrum.CrossRefGoogle Scholar
Feige, GB, Lumbsch, HT and Mies, B (1993) Morphological and chemical changes in Roccella thalli infected by Lecanactis grumulosa. Cryptogamic Botany 3, 101107.Google Scholar
Fernández-Marín, B, López-Pozo, M, Perera-Castro, AV, Irati Arzac, M, Sáenz-Ceniceros, A, Colesie, C, de los Ríos, A, Sancho, LG, Pintado, A, Laza, JM, et al. (2019) Symbiosis at its limits: ecophysiological consequences of lichenization in the genus Prasiola in Antarctica. Annals of Botany 124, 12111226.CrossRefGoogle Scholar
Fernández-Mendoza, F, Domaschke, S, García, MA, Jordan, P, Martín, MP and Printzen, C (2011) Population structure of mycobionts and photobionts of the widespread lichen Cetraria aculeata. Molecular Ecology 20, 12081232.CrossRefGoogle ScholarPubMed
Fewer, D, Friedl, T and Büdel, B (2002) Chroococcidiopsis and heterocyst-differentiating cyanobacteria are each other's closest living relatives. Molecular Phylogenetics and Evolution 23, 8290.CrossRefGoogle ScholarPubMed
Fiore, MF, Sant'Anna, CL, Azevedo, MTDP, Komárek, J, Kaštovský, J, Sulek, J and Lorenzi, AAS (2007) The cyanobacterial genus Brasilonema, gen. nov., a molecular and phenotypic evaluation. Journal of Phycology 43, 789798.CrossRefGoogle Scholar
Fisher, KA and Lang, NJ (1971) Comparative ultrastructure of cultured species of Trebouxia. Journal of Phycology 7, 155165.Google Scholar
Flechtner, VR, Johansen, JR and Belnap, J (2009) The biological soil crusts of the San Nicolas Island: enigmatic algae from a geographically isolated ecosystem. Western North American Naturalist 68, 405436.CrossRefGoogle Scholar
Fontaine, KM, Beck, A, Stocker-Wörgötter, E and Piercey-Normore, MD (2012) Photobiont relationships and phylogenetic history of Dermatocarpon luridum var. luridum and related Dermatocarpon species. Plants 1, 3960.CrossRefGoogle ScholarPubMed
Forsberg, A and Lindblad, P (1996) The Ophrydium-Chlorella (Chlorophyceae, Chlorophyta) symbiosis: an ultrastructural characterization. Phycologia 35, 4447.CrossRefGoogle Scholar
Francisco de Oliveira, PM, Timsina, B and Piercey-Normore, MD (2012) Diversity of Ramalina sinensis and its photobiont in local populations. Lichenologist 44, 649660.CrossRefGoogle Scholar
Friedl, T (1987) Thallus development and phycobionts of the parasitic lichen Diploschistes muscorum. Lichenologist 19, 183191.CrossRefGoogle Scholar
Friedl, T (1989) Comparative ultrastructure of pyrenoids in Trebouxia (Microthamniales, Chlorophyta). Plant Systematics and Evolution 164, 145159.CrossRefGoogle Scholar
Friedl, T (1995) Inferring taxonomic positions and testing genus level assignments in coccoid green lichen algae: a phylogenetic analysis of 18S ribosomal RNA sequences from Dictyochloropsis reticulata and from members of the genus Myrmecia (Chlorophyta, Trebouxiophyceae cl. nov.). Journal of Phycology 31, 632639.CrossRefGoogle Scholar
Friedl, T (1996) Evolution of the polyphyletic genus Pleurastrum (Chlorophyta): inferences from nuclear-encoded ribosomal DNA sequences and motile cell ultrastructure. Phycologia 35, 456469.CrossRefGoogle Scholar
Friedl, T and Büdel, B (2008) Photobionts. In Nash, TH III (ed.), Lichen Biology. New York: Cambridge University Press, pp. 926.CrossRefGoogle Scholar
Fritsch, FE (1945) The Structure and Reproduction of the Algae, Vol. II. Cambridge: Cambridge University Press.Google Scholar
Fröberg, L, Björn, LO, Baur, A and Baur, B (2001) Viability of lichen photobionts after passing through the digestive tract of a land snail. Lichenologist 33, 543545.CrossRefGoogle Scholar
Fučíková, K, Lewis, PO and Lewis, LA (2014) Putting incertae sedis taxa in their place: a proposal for ten new families and three new genera in Sphaeropleales (Chlorophyceae, Chlorophyta). Journal of Phycology 50, 1425.CrossRefGoogle Scholar
Fučíková, K, Lewis, PO, Neupane, S, Karol, KG and Lewis, LA (2019) Order, please! Uncertainty in the ordinal-level classification of Chlorophyceae. PeerJ 7, e6899.CrossRefGoogle ScholarPubMed
Gagunashvili, AN and Andrésson, ÓS (2018) Distinctive characters of Nostoc genomes in cyanolichens. BMC Genomics 19, 434.CrossRefGoogle ScholarPubMed
Galløe, O (1927, 1932) Natural History of the Danish Lichens, Parts I and V. Copenhagen: H. Aschehoug and Co, Levin & Munksgaard.Google Scholar
Gallun [sic], M, Ben-Shaul, Y and Paran, N (1971) The fungus-alga association in the Lecideaceae: an ultrastructural study. New Phytologist 70, 483485.CrossRefGoogle Scholar
Galun, M, Paran, N and Ben-Shaul, Y (1970) The fungus-alga association in the Lecanoraceae: an ultrastructural study. New Phytologist 69, 599603.CrossRefGoogle Scholar
Garbary, DJ and London, JF (1995) The Ascophyllum/Polysiphonia/Mycosphaerella symbiosis. V. Fungal infection protects A. nodosum from desiccation. Botanica Marina 38, 529533.CrossRefGoogle Scholar
Garbary, DJ and MacDonald, KA (1995) The Ascophyllum/Polysiphonia/Mycosphaerella symbiosis. IV. Mutualism in the Ascophyllum/Mycosphaerella interaction. Botanica Marina 38, 221225.CrossRefGoogle Scholar
Garcia-Pichel, F (2009) Cyanobacteria. In Schaechter, M (ed.), Encyclopedia of Microbiology, Vol. 2, 3rd Edn. Oxford and San Diego: Elsevier [Academic Press], pp. 107124.CrossRefGoogle Scholar
Garrido-Benavent, I, Pérez-Ortega, S and de los Ríos, A (2017) From Alaska to Antarctica: species boundaries and genetic diversity of Prasiola (Trebouxiophyceae), a foliose chlorophyte associated with the bipolar lichen-forming fungus Mastodia tessellata. Molecular Phylogenetics and Evolution 107, 117131.CrossRefGoogle Scholar
Garrido-Benavent, I, de los Ríos, A, Fernández-Mendoza, F and Pérez-Ortega, S (2018) No need for stepping stones: direct, joint dispersal of the lichen-forming fungus Mastodia tesselata (Ascomycota) and its photobiont explains their bipolar distribution. Journal of Biogeography 45, 213224.CrossRefGoogle Scholar
Garrido-Benavent, I, Pérez-Ortega, S, de los Ríos, A and Fernández-Mendoza, F (2020) Amphitropical variation of the algal partners of Pseudephebe (Parmeliaceae, lichenized fungi). Symbiosis 82, 3548.CrossRefGoogle Scholar
Gärtner, G (1985) Die Gattung Trebouxia Puymaly (Chlorellales, Chlorophyceae). Archiv für Hydrobiologie, Supplement 71 [Algological Studies 41], 495548.Google Scholar
Gärtner, G and Ingolić, E (1989) Ein Beitrag zur Kenntnis von Apatococcus lobatus (Chlorophyta, Chaetophorales, Leptosiroideae). Plant Systematics and Evolution 164, 133143.CrossRefGoogle Scholar
Garty, J and Delarea, J (1988) Evidence of liberation of lichen ascospores in clusters and reports on contact between free-living algal cells and germinating lichen ascospores under natural conditions. Canadian Journal of Botany 66, 21712177.CrossRefGoogle Scholar
Gasulla, F, Guéra, A, de los Ríos, A and Pérez-Ortega, S (2019) Differential responses to salt concentrations of lichen photobiont strains isolated from lichens occurring in different littoral zones. Plant and Fungal Systematics 64, 149162.CrossRefGoogle Scholar
Gasulla, F, Barrasa, JM, Casano, LM and del Campo, EM (2020) Symbiont composition of the basidiolichen Lichenomphalia meridionalis varies with altitude in the Iberian Peninsula. Lichenologist 52, 1726.CrossRefGoogle Scholar
Gauslaa, Y, Alam, MA, Lucas, P-L, Chowdhury, DP and Solhaug, KA (2017) Fungal tissue per se is stronger as a UV-B screen than secondary fungal extrolites in Lobaria pulmonaria. Fungal Ecology 26, 109113.CrossRefGoogle Scholar
Geitler, L (1933) Beiträge zur Kenntnis de Flechtensymbiose. I–III. Archiv für Protistenkunde 80, 378409.Google Scholar
Geitler, L (1934) Beiträge zur Kenntnis de Flechtensymbiose. IV, V. Archiv für Protistenkunde 82, 5185.Google Scholar
Geitler, L (1955) Clavaria mucida, eine extratropische Basidiolichene. Biologisches Zentralblatt 74, 145159.Google Scholar
Giddings TH, Jr, and Staehelin, LA (1981) Observations of microplasmodesmata in both heterocyst-forming and non-heterocyst forming filamentous cyanobacteria by freeze-fracture electron microscopy. Archives of Microbiology 129, 295298.CrossRefGoogle Scholar
González, MA, Pröschold, T, Palacios, Y, Aguayo, P, Inostroza, I and Gómez, PI (2013) Taxonomic identification and lipid production of two Chilean Chlorella-like strains isolated from a marine and an estuarine coastal environment. AoB PLANTS 5, plt020.CrossRefGoogle Scholar
Good, BH and Chapman, RL (1978) The ultrastructure of Phycopeltis (Chroolepidaceae: Chlorophyta). I. Sporopollenin in the cell walls. American Journal of Botany 65, 2733.CrossRefGoogle Scholar
Graham, LE, Graham, JM and Wilcox, LW (2009) Algae, 2nd Edn. San Francisco: Pearson Education Inc.Google Scholar
Green, MS (2012) The Trentepohliales (Ulvophyceae, Chlorophyta) from coastal South Carolina. Ph.D. thesis, Tennessee Technological University.Google Scholar
Green, TGA, Büdel, B, Heber, U, Meyer, A, Zellner, H and Lange, OL (1993) Differences in photosynthetic performance between cyanobacterial and green algal components of lichen photosymbiodemes measured in the field. New Phytologist 125, 723731.CrossRefGoogle ScholarPubMed
Green, TGA, Schlensog, M, Sancho, LG, Winkler, JB, Broom, FD and Schroeter, B (2002) The photobiont determines the pattern of photosynthetic activity within a single lichen thallus containing cyanobacterial and green algal sectors (photosymbiodeme). Oecologia 130, 191198.CrossRefGoogle Scholar
Greenhalgh, GN and Anglesea, D (1979) The distribution of algal cells in lichen thalli. Lichenologist 11, 283292.CrossRefGoogle Scholar
Grube, M and Berg, B (2009) Microbial consortia of bacteria and fungi with focus on the lichen symbiosis. Fungal Biology Reviews 23, 7285.CrossRefGoogle Scholar
Grube, M and Lücking, R (2002) Fine structure of foliicolous lichens and their lichenicolous fungi studied by epifluorescence. Symbiosis 32, 229246.Google Scholar
Grube, M, Cernava, T, Soh, J, Fuchs, S, Aschenbrenner, I, Lassek, C, Wegner, U, Becher, D, Riedel, K, Sensen, CW, et al. (2015) Exploring functional contexts of symbiotic sustain within lichen-associated bacteria by comparative omics. ISME Journal 9, 412424.CrossRefGoogle ScholarPubMed
Grube, M, Muggia, L, Baloch, E, Hametner, C and Stocker-Wörgötter, E (2017 a) Symbiosis of lichen-forming fungi with trentepohlialean algae. In Grube, M, Seckbach, J and Muggia, L (eds), Algal and Cyanobacterial Symbioses. London: World Scientific Publishing Europe Ltd, pp. 85110.CrossRefGoogle Scholar
Grube, M, Seckbach, J and Muggia, L (2017 b) Algal and Cyanobacterial Symbioses. London: World Scientific Publishing Europe Ltd.CrossRefGoogle Scholar
Gueidan, C, Thüs, H and Pérez-Ortega, S (2011) Phylogenetic position of the brown algae-associated lichenized fungus Verrucaria tavaresiae (Verrucariaceae). Bryologist 114, 563569.CrossRefGoogle Scholar
Gustavs, L, Eggert, A, Michalik, D and Karsten, U (2010) Physiological and biochemical responses of green microalgae from different habitats to osmotic and matric stress. Protoplasma 243, 314.CrossRefGoogle ScholarPubMed
Gustavs, L, Görs, M and Karsten, U (2011) Polyol patterns in biofilm-forming aeroterrestrial green algae (Trebouxiophyceae, Chlorophyta). Journal of Phycology 47, 533537.CrossRefGoogle Scholar
Gustavs, L, Schumann, R and Karsten, U (2016) Mixotrophy in the terrestrial green alga Apatococcus lobatus (Trebouxiophyceae, Chlorophyta). Journal of Phycology 52, 311314.CrossRefGoogle Scholar
Gustavs, L, Schiefelbein, U and Darienko, T (2017) Symbioses of the green algal genera Coccomyxa and Elliptochloris (Trebouxiophyceae, Chlorophyta). In Grube, M, Seckbach, J and Muggia, L (eds), Algal and Cyanobacterial Symbioses. London: World Scientific Publishing Europe Ltd, pp. 169208.CrossRefGoogle Scholar
Guzow-Krzemińska, B (2006) Photobiont flexibility in the lichen Protoparmeliopsis muralis as revealed by ITS rDNA analyses. Lichenologist 38, 469476.CrossRefGoogle Scholar
Hallmann, C, Stannek, L, Fritzlar, D, Hause-Reitner, D, Friedl, T and Hoppert, M (2013) Molecular diversity of phototrophic biofilms on building stone. FEMS Microbiology Ecology 84, 355372.CrossRefGoogle ScholarPubMed
Hallmann, C, Hoppert, M, Mudimu, O and Friedl, T (2016) Biodiversity of green algae covering artificial hard substrate surfaces in a suburban environment: a case study using molecular approaches. Journal of Phycology 52, 732744.CrossRefGoogle Scholar
Hametner, C, Stocker-Wörgötter, E and Grube, M (2014 a) New insights into diversity and selectivity of trentepohlialean lichen photobionts from the extratropics. Symbiosis 63, 3140.CrossRefGoogle ScholarPubMed
Hametner, C, Stocker-Wörgötter, E, Rindi, F and Grube, M (2014 b) Phylogenetic position and morphology of lichenized Trentepohliales (Ulvophyceae, Chlorophyta) from selected species of Graphidaceae. Phycological Research 62, 170186.CrossRefGoogle Scholar
Handa, S, Ohmura, Y, Nakano, T and Nakahara-Tsubota, M (2007) Airborne green microalgae (Chlorophyta) in snowfall. Hikobia 15, 109120 [In Japanese, with English abstract and figure legends].Google Scholar
Hauck, M, Helms, G and Friedl, T (2007) Photobiont selectivity in the epiphytic lichens Hypogymnia physodes and Lecanora conizaeoides. Lichenologist 39, 195204.CrossRefGoogle Scholar
Hawksworth, DL (1988) The variety of fungal-algal symbioses, their evolutionary significance, and the nature of lichens. Botanical Journal of the Linnean Society 96, 320.CrossRefGoogle Scholar
Hawksworth, DL, Coppins, BJ and James, PW (1979) Blarneya, a lichenized hyphomycete from southern Ireland. Botanical Journal of the Linnean Society 79, 357367.CrossRefGoogle Scholar
Helms, G, Friedl, T, Rambold, G and Mayrhofer, H (2001) Identification of photobionts from the lichen family Physciaceae using algal-specific ITS rDNA sequencing. Lichenologist 33, 7386.CrossRefGoogle Scholar
Henley, WJ, Hironaka, JL, Gillou, L, Buchheim, MA, Buchheim, JA, Fawley, MW and Fawley, KP (2004) Phylogenetic analysis of the ‘Nannochloris-like’ algae and diagnoses of Picochlorum oklahomensis gen. et sp. nov. (Trebouxiophyceae, Chlorophyta). Phycologia 43, 641652.CrossRefGoogle Scholar
Henskens, FL, Green, TGA and Wilkins, A (2012) Cyanolichens can have both cyanobacteria and green algae in a common layer as major contributors to photosynthesis. Annals of Botany 110, 555563.CrossRefGoogle Scholar
Henson, BJ, Watson, LE and Barnum, SR (2002) Molecular differentiation of the heterocystous cyanobacteria, Nostoc and Anabaena, based on complete NifD sequences. Current Microbiology 45, 161164.CrossRefGoogle ScholarPubMed
Henssen, A (1981) Hyphomorpha als Phycobiont in Flechten. Plant Systematics and Evolution 137, 139143.CrossRefGoogle Scholar
Henssen, A (1990) Thermutopsis jamesii, a new member of the Lichinaceae from Antigua. Lichenologist 22, 253259.CrossRefGoogle Scholar
Henssen, A (1994) Contribution to the morphology and species delimitation in Heppia sensu stricto (lichenized Ascomycotina). Acta Botanica Fennica 150, 5773.Google Scholar
Henssen, A (1995) Studies on the biology and structure of Dacampia (Dothideales), a genus with lichenized and lichenicolous species. Cryptogamic Botany 5, 149158.Google Scholar
Henssen, A and Jahns, HM (1974) Lichenes: Eine Einführung in die Flechtenkunde. Stuttgart: Georg Thieme.Google Scholar
Hessler, R and Peveling, E (1978) Die Lokalisation von 14C-Assimilaten in Flechtenthalli von Cladonia incrassata Floerke und Hypogymnia physodes (L.) Ach. Zeitschrift für Pflanzenphysiologie 86, 287302.CrossRefGoogle Scholar
Hestmark, G, Lutzoni, F and Miadlikowska, J (2016) Photobiont associations in co-occurring umbilicate lichens with contrasting modes of reproduction in coastal Norway. Lichenologist 48, 545557.CrossRefGoogle Scholar
Hill, DJ (1972) The movement of carbohydrate from the alga to the fungus in the lichen Peltigera polydactyla. New Phytologist 71, 3139.CrossRefGoogle Scholar
Hill, DJ (1985) Changes in photobiont dimensions and numbers during co-development of lichen symbionts. In Brown, DH (ed.), Lichen Physiology and Cell Biology. New York and London: Plenum Press, pp. 303317.CrossRefGoogle Scholar
Hill, DJ (1989) The control of the cell cycle in microbial symbionts. New Phytologist 112, 175184.CrossRefGoogle Scholar
Hill, DJ (2009) Asymmetrical co-evolution in the lichen symbiosis caused by a limited capacity for adaptation in the photobiont. Botanical Review 75, 326338.CrossRefGoogle Scholar
Hitch, CBJ and Millbank, JW (1975) Nitrogen metabolism in lichens VII. Nitrogenase activity and heterocyst frequency in lichens with blue-green phycobionts. New Phytologist 75, 239244.CrossRefGoogle Scholar
Hodač, L, Hallmann, C, Spitzer, K, Elster, J, Faßhauer, F, Brinkmann, N, Lepka, D, Diwan, V and Fried, T (2016) Widespread green algae Chlorella and Stichococcus exhibit polar-temperate and tropical-temperate biogeography. FEMS Microbiology Ecology 92, fiw122.CrossRefGoogle ScholarPubMed
Hodkinson, BP, Moncada, B and Lücking, R (2014) Lepidostromatales, a new order of lichenized fungi (Basidiomycota, Agaricomycetes), with two new genera, Ertzia and Sulzbacheromyces, and one new species, Lepidostroma winklerianum. Fungal Diversity 64, 165179.CrossRefGoogle Scholar
Honegger, R (1986) Ultrastructural studies in lichens I. Haustorial types and their frequencies in a range of lichens with trebouxioid photobionts. New Phytologist 103, 785795.CrossRefGoogle Scholar
Honegger, R (1987) Questions about pattern formation in the algal layer of lichens with stratified (heteromerous) thalli. Bibliotheca Lichenologica 25, 5971.Google Scholar
Honegger, R (1990) Haustoria-like structures and cell wall surface layers in lichens. In Mendgen, K and Leseman, DE (eds), Electron Microscopy of Plant Pathogens. Berlin: Springer-Verlag, pp. 277290.Google Scholar
Honegger, R (1991) Functional aspects of the lichen symbiosis. Annual Review of Plant Physiology and Plant Molecular Biology 42, 553578.CrossRefGoogle Scholar
Honegger, R (2012) The symbiotic phenotype of lichen-forming ascomycetes and their endo- and epibionts. In Hock, B (ed.), The Mycota IX: Fungal Associations. Berlin and Heidelberg: Springer-Verlag, pp. 287339.CrossRefGoogle Scholar
Honegger, R and Brunner, U (1981) Sporopollenin in the cell wall of Coccomyxa and Myrmecia phycobionts of various lichens: an ultrastructural and chemical investigation. Canadian Journal of Botany 59, 27132734.CrossRefGoogle Scholar
Hoshina, R and Imamura, N (2008) Multiple origins of the symbioses in Paramecium bursaria. Protist 159, 5363.CrossRefGoogle ScholarPubMed
Hoshina, R, Kamako, S-i and Imamura, N (2004) Phylogenetic position of endosymbiotic green algae in Paramecium bursaria Ehrenberg from Japan. Plant Biology 6, 447453.CrossRefGoogle ScholarPubMed
Hoshina, R, Iwataki, M and Imamura, N (2010) Chlorella variabilis and Micractinium reisseri sp. nov. (Chlorellaceae, Trebouxiophyceae): redescription of the endosymbiotic green algae of Paramecium bursaria (Peniculia, Oligohymenophorea) in the 120th year. Phycological Research 58, 188201.CrossRefGoogle Scholar
Huss, VAR, Frank, C, Hartmann, EC, Hirmer, M, Kloboucek, A, Seidel, B, Wenzeler, P and Kessler, E (1999) Biochemical taxonomy and molecular phylogeny of the genus Chlorella sensu lato (Chlorophyta). Journal of Phycology 35, 587598.CrossRefGoogle Scholar
Ihda, T, Nakano, T and Deguchi, H (1997) Photobionts of Japanese Sphaerophorus. Symbiosis 23, 18.Google Scholar
Ikeda, T and Takeda, H (1995) Species-specific differences of pyrenoids in Chlorella (Chlorophyta). Journal of Phycology 31, 813818.CrossRefGoogle Scholar
Jaag, O and Thomas, E (1934) Neue Untersuchungen über die Flechte Epigloea bactrospora Zukal. Berichte der Schweizerschen Botanischen Gesellschaft 43, 7789.Google Scholar
Jacobs, JB and Ahmadjian, V (1971) The ultrastructure of lichens. IV. Movement of carbon products from alga to fungus as demonstrated by high resolution autoradiography. New Phytologist 70, 4750.CrossRefGoogle Scholar
Jahns, HM, Mollenhauer, D, Jenniger, M and Schönborn, D (1979) Die Neubesiedlung von Baumrinde durch Flechten I. Natur und Museum 109, 4051.Google Scholar
Jaklitsch, W, Baral, H-O, Lücking, R and Lumbsch, HT (2016) Syllabus of Plant Families: A. Engler's Syllabus der Pflanzenfamilien. Part 1/2, Ascomycota. Stuttgart: Borntraeger Verlagsbuchhandlung.Google Scholar
James, PW and Henssen, A (1976) The morphological and taxonomic significance of cephalodia. In Brown, DH, Hawksworth, DL and Bailey, RH (eds), Lichenology: Progress and Problems. London: Academic Press, pp. 2777.Google Scholar
Jayalal, U, Aptroot, A and Hur, J-S (2012) The lichen genus Polychidium new to South Korea. Mycobiology 40, 252254.CrossRefGoogle ScholarPubMed
Jiang, SH, Hawksworth, DL, Lücking, R and Wei, JC (2020) A new genus and species of foliicolous lichen in a new family of Strigulales (Ascomycota: Dothideomycetes) reveals remarkable class-level homoplasy. IMA Fungus 11, 1.CrossRefGoogle Scholar
Joneson, S and O'Brien, H (2017) A molecular investigation of free-living and lichenized Nostoc sp. and symbiotic lifestyle determination. Bryologist 120, 371381.CrossRefGoogle Scholar
Jordan, WP (1970) The internal cephalodia of the genus Lobaria. Bryologist 73, 669681.CrossRefGoogle Scholar
Jordan, WP (1972) Erumpent cephalodia, an apparent case of phycobial influence on lichen morphology. Journal of Phycology 8, 112117.Google Scholar
Jordan, WP and Rickson, F (1971) Cyanophyte cephalodia in the lichen genus Nephroma. American Journal of Botany 58, 562568.CrossRefGoogle Scholar
Joubert, JJ and Rijkenberg, FJH (1971) Parasitic green algae. Annual Review of Phytopathology 9, 4564.CrossRefGoogle Scholar
Jülich, W (1972) Monographie der Atheliae (Corticiaceae, Basidiomycetes). Willdenowia, Beiheft 7, 1283.Google Scholar
Jülich, W (1978) A new lichenized Athelia from Florida. Persoonia 10, 149151.Google Scholar
Jung, P, Emrich, D, Briegel-Williams, L, Schermer, M, Weber, L, Baumann, K, Colesie, C, Clerc, P, Lehnert, LW, Achilles, S, et al. (2019) Ecophysiology and phylogeny of new terricolous and epiphytic chlorolichens in a fog oasis of the Atacama Desert. MicrobiologyOpen 8, e894.CrossRefGoogle Scholar
Jüriado, I, Kaasalainen, U, Jylhä, and Rikkinen, J (2019) Relationships between mycobiont identity, photobiont specificity and ecological preferences in the lichen genus Peltigera (Ascomycota) in Estonia (northeastern Europe). Fungal Ecology 39, 4554.CrossRefGoogle Scholar
Kardish, N, Kessel, M and Galun, M (1989) Characterization of symbiotic and cultured Nostoc of the lichen Nephroma laevigatum Ach. Symbiosis 7, 257266.Google Scholar
Keeling, PJ (2004) Diversity and evolutionary history of plastids and their hosts. American Journal of Botany 91, 14811493.CrossRefGoogle ScholarPubMed
Keeling, PJ (2013) The number, speed, and impact of plastid endosymbioses in eukaryotic evolution. Annual Review of Plant Biology 64, 583607.CrossRefGoogle ScholarPubMed
Kharkongor, D and Ramanujam, P (2015) Spatial and temporal variation of carotenoids in four species of Trentepohlia (Trentepohliales, Chlorophyta). Journal of Botany 2015, 201641.CrossRefGoogle Scholar
Khayatan, B, Meeks, J and Risser, DD (2015) Evidence that a modified type IV pilus-like system powers gliding motility and polysaccharide secretion in filamentous cyanobacteria. Molecular Microbiology 98, 10211036.CrossRefGoogle ScholarPubMed
Kim, JI, Nam, SW, So, JE, Hong, SG, Choi, HG and Shin, W (2017) Asterochloris sejongensis sp. nov. (Trebouxiophyceae, Chlorophyta) from King George Island, Antarctica. Phytotaxa 295, 6070.CrossRefGoogle Scholar
Kim, JI, Kim, YJ, Nam, SW, So, JE, Hong, SG, Choi, H-G and Shin, W (2020) Taxonomic study of three new Antarctic Asterochloris (Trebouxiophyceae) based on morphological and molecular data. Algae 35, 1732.CrossRefGoogle Scholar
Kirk, PM, Cannon, PF, David, JC and Stalpers, JA (2001) Ainsworth and Bisby's Dictionary of the Fungi, 9th Edn. Wallingford: CAB International.Google Scholar
Kluge, M, Mollenhauer, D, Wolf, E and Schüßler, A (2002) The Nostoc-Geosiphon endocytobiosis. In Rai, AN, Bergman, B and Rasmussen, U (eds), Cyanobacteria in Symbiosis. Dordrecht: Kluwer Academic Publishers, pp. 1930.Google Scholar
Kohlmeyer, J, Hawksworth, DL and Volkmann-Kohlmeyer, B (2004) Observations on two marine and maritime ‘borderline’ lichens: Mastodia tessellata and Collemopsidium pelvetiae. Mycological Progress 3, 5156.CrossRefGoogle Scholar
Komárek, J, Kaštovský, J, Mareš, J and Johansen, JR (2014) Taxonomic classification of cyanoprokaryotes (cyanobacterial genera) 2014, using a polyphasic approach. Preslia 86, 295335.Google Scholar
Kono, M, Kon, Y, Ohmura, Y, Satta, Y and Terai, Y (2020) In vitro resynthesis of lichenization reveals the genetic background of symbiosis-specific fungal-algal interaction in Usnea hakonensis. BMC Genomics 21, 116.CrossRefGoogle ScholarPubMed
Kosecka, M, Jabłońska, A, Flakus, A, Rodriguez-Flakus, P, Kukwa, M and Guzow-Krzemińska, B (2020) Trentepohlialean algae (Trentepohliales, Ulvophyceae) show preference to selected mycobiont lineages in lichen symbioses. Journal of Phycology 56, 979993.CrossRefGoogle ScholarPubMed
Kosecka, M, Guzow-Krzemińska, B, Čemajová, I, Škaloud, P, Jabłońska, A and Kukwa, M (2021) New lineages of photobionts in Bolivian lichens expand our knowledge on habitat preferences and distribution of Asterochloris algae. Scientific Reports 11, 8701.CrossRefGoogle ScholarPubMed
Kosugi, M, Arita, M, Shizuma, R, Moryama, Y, Kashino, Y, Koike, H and Satoh, K (2009) Responses to desiccation stress in lichens are different from those in their photobionts. Plant and Cell Physiology 50, 879888.CrossRefGoogle ScholarPubMed
Kouwets, FAC (1996) Ultrastructural studies of the cell cycle in a multicentriolar form of Bracteacoccus minor (Chlorophyceae, Chlorellales). Protoplasma 191, 191204.CrossRefGoogle Scholar
Kovačević, G, Franjević, D, Jelenčić, B and Kalafatić, M (2010) Isolation and cultivation of endosymbiotic algae from green hydra and phylogenetic analysis of 18S rDNA sequences. Folia Biologica (Kraków ) 58, 135143.CrossRefGoogle ScholarPubMed
Kovačik, L and Batista Pereira, A (2001) Green alga Prasiola crispa and its lichenized form Mastodia tesselata in Antarctic environment: general aspects. Nova Hedwigia 123, 465478.Google Scholar
Kranner, I, Cram, WJ, Zorn, M, Wornik, S, Yoshimura, I, Stabentheiner, E and Pfeifhofer, HW (2005) Antioxidants and photoprotection in a lichen as compared with its isolated symbiotic partners. Proceedings of the National Academy of Sciences of the United States of America 102, 31413146.CrossRefGoogle Scholar
Kranner, I, Beckett, R, Hochman, A and Nash, TH III (2008) Desiccation-tolerance in lichens: a review. Bryologist 111, 576593.CrossRefGoogle Scholar
Kroken, S and Taylor, JW (2000) Phylogenetic species, reproductive mode, and specificity of the green alga Trebouxia forming lichens with the fungal genus Letharia. Bryologist 103, 645660.CrossRefGoogle Scholar
Kukwa, M and Pérez-Ortega, S (2010) A second species of Botryolepraria from the Neotropics and the phylogenetic placement of the genus within Ascomycota. Mycological Progress 9, 345351.CrossRefGoogle Scholar
Kumar, K, Mella-Herrera, RA and Golden, JW (2010) Cyanobacterial heterocysts. Cold Spring Harbor Perspectives in Biology 2, a000315.CrossRefGoogle ScholarPubMed
Lakatos, M, Lange-Bertalot, H and Büdel, B (2004) Diatoms living inside the thallus of the green algal lichen Coenogonium linkii in neotropical lowland rain forests. Journal of Phycology 40, 7073.CrossRefGoogle Scholar
Lamb, IM (1959) Lichens. Scientific American 201, 144159.CrossRefGoogle Scholar
Lange, OL, Kilian, E and Ziegler, H (1986) Water vapor uptake and photosynthesis of lichens: performance differences in species with green and blue-green algae as phycobionts. Oecologia 71, 104110.CrossRefGoogle ScholarPubMed
Lange, OL, Pfanz, H, Kilian, E and Meyer, A (1990) Effect of low water potential on photosynthesis in intact lichens and their liberated algal components. Planta 182, 467472.CrossRefGoogle ScholarPubMed
Lange, OL, Büdel, B, Meyer, A and Kilian, E (1993) Further evidence that activation of net photosynthesis by dry cyanobacterial lichens requires liquid water. Lichenologist 25, 175189.CrossRefGoogle Scholar
Lavoie, C, Renaudin, M, McMullin, RT, Ganon, J, Roy, C, Beaulieu, M-E, Bellenger, JP and Villareal, JC (2020) Extremely low genetic diversity of Stigonema associated with Stereocaulon in eastern Canada. Bryologist 123, 188203.CrossRefGoogle Scholar
Leavitt, SD, Nelsen, MP, Lumbsch, HT, Johnson, LA and St Clair, LL (2013) Symbiont flexibility in subalpine rock shield lichen communities in the Southwestern USA. Bryologist 116, 149161.CrossRefGoogle Scholar
Leavitt, SD, Kraichak, E, Vondrak, J, Nelsen, MP, Altermann, S, Divakar, PK, Alors, D, Esslinger, TL, Crespo, A and Lumbsch, HT (2015) Fungal specificity and selectivity for algae play a major role in determining lichen partnerships across diverse ecogeographic regions in the lichen-forming family Parmeliaceae (Ascomycota). Molecular Ecology 24, 37793797.CrossRefGoogle Scholar
Leavitt, SD, Kraichak, E, Vondrak, J, Nelsen, MP, Sohrabi, M, Pérez-Ortega, S, St Clair, LL and Lumbsch, HT (2016) Cryptic diversity and symbiont interactions in rock-posy lichens. Molecular Phylogenetics and Evolution 99, 261274.CrossRefGoogle ScholarPubMed
Lee, RE (2018) Phycology, 5th Edn. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Leliaert, F, Smith, DR, Moreau, H, Herron, MD, Verbruggen, H, Delwiche, CF and De Clerck, O (2012) Phylogeny and molecular evolution of the green algae. Critical Reviews in Plant Sciences 31, 146.CrossRefGoogle Scholar
Lemieux, C, Otis, C and Turmel, M (2014) Chloroplast phylogenomic analysis resolves deep-level relationships within the green algal class Trebouxiophyceae. BMC Evolutionary Biology 14, 211.CrossRefGoogle ScholarPubMed
Letrouit-Galinou, M-A (1968) Les algues des lichens. Bulletin de la Société Botanique de France 115 (Supp. 2), 3577.CrossRefGoogle Scholar
Letrouit-Galinou, M-A and Asta, J (1994) Thallus morphogenesis in some lichens. Cryptogamic Botany 4, 274282.Google Scholar
Letsch, MR, Muller-Parker, G, Friedl, T and Lewis, LA (2009) Elliptochloris marina sp. nov. (Trebouxiophyceae, Chlorophyta), symbiotic green alga of the temperate pacific sea anemones Anthopleura xanthogrammica and A. elegantissima (Anthozoa, Cnidaria). Journal of Phycology 45, 11271135.CrossRefGoogle Scholar
Lewis, L and Muller-Parker, G (2004) Phylogenetic placement of ‘Zoochlorella’ (Chlorophyta), algal symbiont of the temperate sea anemone Anthopleura elegantissima. Biological Bulletin 207, 8792.CrossRefGoogle Scholar
Li, B, Feng, J and Xie, S (2013) Morphological and phylogenetic study of algal partners associated with the lichen-forming fungus Porpidia crustulata from the Guancen Mountains, northern China. Symbiosis 61, 3746.CrossRefGoogle Scholar
Li, S, Sun, H, Hu, Y, Liu, B, Zhu, H, Hu, Z and Liu, G (2020) Four new members of foliicolous green algae within the Watanabea clade (Trebouxiophyceae, Chlorophyta) from China. Journal of Eukaryotic Microbiology 67, 369382.CrossRefGoogle ScholarPubMed
Li, S, Tan, H, Liu, B, Zhu, H, Hu, Z and Liu, G (2021) Watanabeales ord. nov. and twelve novel species of Trebouxiophyceae (Chlorophyta). Journal of Phycology, doi: 10.1111/JPY.13165CrossRefGoogle Scholar
Lindgren, H, Velmala, S, Högnabba, F, Goward, T, Holien, H and Myllys, L (2014) High fungal selectivity for algal symbionts in the genus Bryoria. Lichenologist 46, 681695.CrossRefGoogle Scholar
Lindgren, H, Moncada, B, Lücking, R, Magain, N, Simon, A, Goffinet, B, Sérusiaux, E, Nelsen, M, Mercado-Díaz, J, Widhelm, T, et al. (2020) Cophylogenetic patterns in algal symbionts correlate with repeated symbiont switches during diversification and geographic expansion of lichen-forming fungi in the genus Sticta (Ascomycota: Peltigeraceae). Molecular Phylogenetics and Evolution 150, 106860.CrossRefGoogle Scholar
Liu, A, Zhu, T, Lu, X and Song, L (2013) Hydrocarbon profiles and phylogenetic analyses of diversified cyanobacterial species. Applied Energy 111, 383393.CrossRefGoogle Scholar
Lohtander, K, Oksanen, I and Rikkinen, J (2003) Genetic diversity of green algal and cyanobacterial photobionts in Nephroma (Peltigerales). Lichenologist 35, 325339.CrossRefGoogle Scholar
López-Bautista, JM and Chapman, RL (2003) Phylogenetic affinities of the Trentepohliales inferred from small-subunit rDNA. International Journal of Systematic and Evolutionary Microbiology 53, 20992106.CrossRefGoogle ScholarPubMed
López-Bautista, JM, Rindi, F and Guiry, MD (2006) Molecular systematics of the subaerial green algal order Trentepohliales: an assessment based on morphological and molecular data. International Journal of Systematic and Evolutionary Microbiology 56, 17091715.CrossRefGoogle ScholarPubMed
Lücking, R (1994) A new foliicolous species of Microtheliopsis (Lichens, Microtheliopsidaceae) from Costa Rica. Mycotaxon 51, 6973.Google Scholar
Lücking, R (2008) Foliicolous lichenized fungi. Flora Neotropica Monograph 103, 1866.Google Scholar
Lücking, R and Grube, M (2002) Facultative parasitism and reproductive strategies in Chroodiscus (Ascomycota, Ostropales). Stapfia 80, 267292.Google Scholar
Lücking, R, Lawrey, JD, Sikaroodi, M, Gilleve, PM, Chaves, JL, Sipman, HJM and Bungartz, F (2009) Do lichens domesticate photobionts like farmers domesticate crops? Evidence from a previously unrecognized lineage of filamentous cyanobacteria. American Journal of Botany 96, 14091418.CrossRefGoogle ScholarPubMed
Lücking, R, Barrie, FR and Genney, D (2014) Dictyonema coppinsii, a new name for the European species known as Dictyonema interruptum (Basidiomycota: Agaricales: Hygrophoraceae), with a validation of its photobiont Rhizonema (Cyanoprokaryota: Nostocales: Rhizonemataceae). Lichenologist 46, 261267.CrossRefGoogle Scholar
Lücking, R, Hodkinson, BP and Leavitt, SD (2017 a) The 2016 classification of lichenized fungi in the Ascomycota and Basidiomycota – approaching one thousand genera. Bryologist 119, 361416.CrossRefGoogle Scholar
Lücking, R, Thorn, RG, Saar, I, Piercey-Normore, MD, Moncada, B, Doering, J, Mann, H, Lebeuf, R, Voitk, M and Voitk, A (2017 b) A hidden basidiolichen rediscovered: Omphalina oreades is a separate species in the genus Lichenomphalia (Basidiomycota: Agaricales: Hygrophoraceae). Lichenologist 49, 467481.CrossRefGoogle Scholar
Lud, D, Huiskes, AHL and Ott, S (2001) Morphological evidence for the symbiotic character of Turgidosculum complicatulum Kohlm. & Kohlm. (= Mastodia tesselata Hook. f. & Harvey). Symbiosis 31, 141151.Google Scholar
Ludwig, LR (2015) The reproductive ecology of Icmadophila splachnirima, including aspects of the reproduction in additional members of Icmadophilaceae. Ph.D. thesis, University of Otago.Google Scholar
Luo, W, Pröschold, T, Bock, C and Krienitz, L (2010) Generic concept in Chlorella-related coccoid green algae (Chlorophyta, Trebouxiophyceae). Plant Biology 12, 545553.CrossRefGoogle Scholar
Lutsak, T, Fernández-Mendoza, F, Kirika, P, Wondafrash, M and Printzen, C (2016) Mycobiont-photobiont interactions of the lichen Cetraria aculeata in high alpine regions of East Africa and South America. Symbiosis 68, 2537.CrossRefGoogle Scholar
Lutzoni, FM and Brodo, IM (1995) A generic redelimitation of the Ionaspis-Hymenelia complex (lichenized Ascomycotina). Systematic Botany 20, 224258.CrossRefGoogle Scholar
Ma, S, Huss, VAR, Tan, D, Sun, X, Chun, J, Xie, Y and Zhang, J (2013) A novel species in the genus Heveochlorella (Trebouxiophyceae, Chlorophyta) witnesses the evolution from an epiphytic into an endophytic lifestyle in tree-dwelling green algae. European Journal of Phycology 48, 200209.CrossRefGoogle Scholar
Magain, N and Sérusiaux, E (2014) Do photobiont switch and cephalodia emancipation act as evolutionary drivers in the lichen symbiosis? A case study in the Pannariaceae (Peltigerales). PLoS ONE 9, e89876.CrossRefGoogle Scholar
Magain, N, Goffinet, B and Sérusiaux, E (2012) Further photomorphs in the lichen family Lobariaceae from Réunion (Mascarene archipelago) with notes on the phylogeny of Dendriscocaulon cyanomorphs. Bryologist 115, 243254.CrossRefGoogle Scholar
Magain, N, Miadlikowska, J, Goffinet, B, Sérusiaux, E and Lutzoni, F (2017) Macroevolution of specificity in cyanolichens of the genus Peltigera section Polydactylon (Lecanoromycetes, Ascomycota). Systematic Botany 66, 7499.Google Scholar
Magain, N, Truong, C, Goward, T, Niu, D, Goffinet, B, Sérusiaux, E, Vitikainen, O, Lutzoni, F and Miadlikowska, J (2018) Species delimitation at a global scale reveals high species richness with complex biogeography and patterns of symbiont association in Peltigera section Peltigera (lichenized Ascomycota: Lecanoromycetes). Taxon 67, 836870.CrossRefGoogle Scholar
Makra, N, Gell, G, Juhász, A, Soós, V, Kiss, T, Molnár, Z, Ördög, V, Vörös, L and Balázs, E (2019) Molecular taxonomic evaluation of Anabaena and Nostoc strains from the Mosonmagyaróvár Algal Culture Collection. South African Journal of Botany 124, 8086.CrossRefGoogle Scholar
Malavasi, V, Škaloud, P, Rindi, F, Tempesta, S, Paoletti, M and Pasqualetti, M (2016) DNA-based taxonomy in ecologically versatile microalgae: a re-evaluation of the species concept within the coccoid green algal genus Coccomyxa (Trebouxiophyceae, Chlorophyta). PLoS ONE 11, e0151137.CrossRefGoogle Scholar
Mansournia, MR, Wu, B, Matsushita, N and Hogetsu, T (2012) Genotypic analysis of the foliose lichen Parmotrema tinctorum using microsatellite markers: association of mycobiont and photobiont, and their reproductive modes. Lichenologist 44, 419440.CrossRefGoogle Scholar
Marini, L, Nascimbene, J and Nimis, PL (2011) Large-scale patterns of epiphytic lichen species richness: photobiont-dependent response to climate and forest structure. Science of the Total Environment 409, 43814386.CrossRefGoogle ScholarPubMed
Mark, K, Laanisto, L, Bueno, CG, Niinemets, Ü, Keller, C and Scheidegger, C (2020) Contrasting co-occurrence patterns of photobiont and cystobasidiomycete yeast associated with common epiphytic lichen species. New Phytologist 227, 13621375.CrossRefGoogle ScholarPubMed
Marton, K and Galun, M (1976) In vitro dissociation and reassociation of the symbionts of the lichen Heppia echinulata. Protoplasma 87, 135143.CrossRefGoogle Scholar
Masumoto, H (2020) Taxonomic studies on lichenized basidiomycetes and their photobionts in Japan: towards the establishment of a model co-culture system of lichen symbiosis. Ph.D. thesis, University of Tsukuba.Google Scholar
Masumoto, H, Ohmura, Y and Degawa, Y (2019) Lichenomphalia meridionalis (Hygrophoraceae, lichenized Basidiomycota) new to Asia. Opuscula Philolichenum 18, 379389.Google Scholar
Matthews, SW, Tucker, SC and Chapman, RL (1989) Ultrastructural features of mycobionts and trentepohliaceous phycobionts in selected tropical crustose lichens. Botanical Gazette 150, 417438.CrossRefGoogle Scholar
Mattox, KR and Stewart, KD (1984) Classification of the green algae: a concept based on comparative cytology. In Irvine, DEG and John, DM (eds), Systematics of the Green Algae. London: Academic Press, pp. 2972.Google Scholar
McCune, B, Arup, U, Breuss, O, Di Meglio, E, Di Meglio, J, Esslinger, TL, Magain, N, Miadlikowska, J, Miller, AE, Muggia, L, et al. (2018) Biodiversity and ecology of lichens of Katmai and Lake Clark National Parks and Preserves, Alaska. Mycosphere 9, 859930.CrossRefGoogle Scholar
McGee, M (2002) Back cover. Bulletin of the California Lichen Society 9, 2526.Google Scholar
Meeks, JC (1998) Symbiosis between nitrogen-fixing cyanobacteria and plants. BioScience 48, 266276.CrossRefGoogle Scholar
Meier, FA, Scherrer, S and Honegger, R (2002) Faecal pellets of lichenivorous mites contain viable cells of the lichen-forming ascomycete Xanthoria parietina and its green algal photobiont, Trebouxia arboricola. Biological Journal of the Linnean Society 76, 259268.CrossRefGoogle Scholar
Meier, JL and Chapman, RL (1983) Ultrastructure of the lichen Coenogonium interplexum Nyl. American Journal of Botany 70, 400407.CrossRefGoogle Scholar
Metz, S, Singer, D, Domaizon, I, Unrein, F and Lara, E (2019) Global distribution of Trebouxiophyceae diversity explored by high-throughput sequencing and phylogenetic approaches. Environmental Microbiology 21, 38853895.CrossRefGoogle ScholarPubMed
Miadlikowska, J, Kauff, F, Högnabba, F, Oliver, JC, Molnár, K, Fraker, E, Gaya, E, Hafellner, J, Hofstetter, V, Gueidan, C, et al. (2014) A multigene phylogenetic synthesis for the class Lecanoromycetes (Ascomycota): 1307 fungi representing 1139 infrageneric taxa, 317 genera and 66 families. Molecular Phylogenetics and Evolution 79, 132168.CrossRefGoogle ScholarPubMed
Míguez, F, Schiefelbein, U, Karsten, U, García-Plazaola, JI and Gustavs, L (2017) Unraveling the photoprotective response of lichenized and free-living green algae (Trebouxiophyceae, Chlorophyta) to photochilling stress. Frontiers in Plant Science 8, 1144.CrossRefGoogle ScholarPubMed
Mikhailyuk, TI, Sluiman, HJ, Massalski, A, Mudimu, O, Demchenko, EM, Kondratyuk, SY and Friedl, T (2008) New streptophyte green algae from terrestrial habitats and an assessment of the genus Interfilum (Klebsormidiophyceae, Streptophyta). Journal of Phycology 44, 15861603.CrossRefGoogle Scholar
Mikhailyuk, T, Holzinger, A, Tsarenko, P, Glaser, K, Demchenko, E and Karsten, U (2020) Dictyosphaerium-like morphotype in terrestrial algae: what is Xerochlorella (Trebouxiophyceae, Chlorophyta)? Journal of Phycology 56, 671686.CrossRefGoogle ScholarPubMed
Millbank, JW and Kershaw, KA (1974 [‘1973’]) Nitrogen metabolism. In Ahmadjian, V and Hale, ME (eds), The Lichens. New York: Academic Press, pp. 289307.Google Scholar
Moe, RL (1997) Verrucaria tavaresiae sp. nov., a marine lichen with a brown algal photobiont. Bulletin of the California Lichen Society 4, 711.Google Scholar
Mohr, F, Ekman, S and Heegaard, E (2004) Evolution and taxonomy of the marine Collemopsidium species (lichenized Ascomycota) in north-west Europe. Mycological Research 108, 515532.CrossRefGoogle Scholar
Molinari-Novoa, EA (2016) Uvulifera, a new generic name for Coccobotrys (Chaetophoraceae). Notulae Algarum 5, 12.Google Scholar
Molins, A, García-Breijo, F, Reig-Armiñana, J, del Campo, E, Casano, L and Barreno, E (2013) Coexistence of different intrathalline symbiotic algae and bacterial biofilms in the foliose Canarian lichen Parmotrema pseudotinctorum. Vieraea 41, 249270.Google Scholar
Molins, A, Moya, P, García-Breijo, F, Reig-Armiñana, J and Barreno, E (2018) Molecular and morphological diversity of Trebouxia microalgae in sphaerothallioid Circinaria spp. lichens. Journal of Phycology 54, 494504.CrossRefGoogle Scholar
Molins, A, Chiva, S, Calatayud, A, Marco, F, García-Breijo, F, Reig-Armiñana, J, Carrasco, P and Moya, P (2020) Multidisciplinary approach to describe Trebouxia diversity within lichenized fungi Buellia zoharyi from the Canary Islands. Symbiosis 82, 1934.CrossRefGoogle Scholar
Moncada, B, Coca, LF and Lücking, R (2013) Neotropical members of Sticta (lichenized Ascomycota: Lobariaceae) forming photosymbiodemes, with the description of seven new species. Bryologist 116, 169200.CrossRefGoogle Scholar
Moniz, MBJ, Rindi, F and Guiry, MD (2012) Phylogeny and taxonomy of Prasiolales (Trebouxiophyceae, Chlorophyta) from Tasmania, including Rosenvingiella tasmanica sp. nov. Phycologia 51, 8697.CrossRefGoogle Scholar
Moniz, MBJ, Guiry, MD and Rindi, F (2014) TufA phylogeny and species boundaries in the green algal order Prasiolales (Trebouxiophyceae, Chlorophyta). Phycologia 53, 396406.CrossRefGoogle Scholar
Moya, P, Škaloud, P, Chiva, S, García-Breijo, FJ, Reig-Armiñana, J, Vančurová, L and Barreno, E (2015) Molecular phylogeny and ultrastructure of the lichen microalga Asterochloris mediterranea sp. nov. from Mediterranean and Canary Islands ecosystems. International Journal of Systematic and Evolutionary Microbiology 65, 18381854.CrossRefGoogle ScholarPubMed
Moya, P, Molins, A, Martínez-Alberola, F, Muggia, L and Barreno, E (2017) Unexpected associated microalgal diversity in the lichen Ramalina farinacea is uncovered by pyrosequencing analyses. PLoS ONE 12, e0175091.CrossRefGoogle ScholarPubMed
Moya, P, Chiva, S, Molins, A, Jadrná, I, Škaloud, P, Peksa, O and Barreno, E (2018) Myrmecia israelensis as the primary symbiotic microalga in squamulose lichens growing in European and Canary Island terricolous communities. Fottea 18, 7285.CrossRefGoogle Scholar
Muggia, L and Grube, M (2018) Fungal diversity in lichens: from extremotolerance to interactions with algae. Life 8, 15.CrossRefGoogle ScholarPubMed
Muggia, L, Grube, M and Tretiach, M (2008) Genetic diversity and photobiont association in selected taxa of the Tephromela atra group (Lecanorales, lichenized Ascomycota). Mycological Progress 7, 147160.CrossRefGoogle Scholar
Muggia, L, Zellnig, G, Rabensteiner, J and Grube, M (2010) Morphological and phylogenetic study of algal partners associated with the lichen-forming fungus Tephromela atra from the Mediterranean region. Symbiosis 51, 149160.CrossRefGoogle Scholar
Muggia, L, Baloch, E, Stabentheiner, E, Grube, M and Wedin, M (2011) Photobiont association and genetic diversity of the optionally lichenized fungus Schizoxylon albescens. FEMS Microbiology Ecology 75, 255272.CrossRefGoogle ScholarPubMed
Muggia, L, Vancurova, L, Škaloud, P, Peksa, O, Wedin, M and Grube, M (2013) The symbiotic playground of lichen thalli – a highly flexible photobiont association in rock-inhabiting lichens. FEMS Microbiology Ecology 85, 313323.CrossRefGoogle ScholarPubMed
Muggia, L, Pérez-Ortega, S, Kopun, T, Zellnig, G and Grube, M (2014) Photobiont selectivity leads to ecological tolerance and evolutionary divergence in a polymorphic complex of lichenized fungi. Annals of Botany 114, 463475.CrossRefGoogle Scholar
Muggia, L, Candotto-Carniel, F and Grube, M (2017) The lichen photobiont Trebouxia: towards and appreciation of species diversity and molecular studies. In Grube, M, Seckbach, J and Muggia, L (eds), Algal and Cyanobacterial Symbioses. London: World Scientific Publishing Europe Ltd, pp. 111146.CrossRefGoogle Scholar
Muggia, L, Leavitt, S and Barreno, E (2018) The hidden diversity of lichenised Trebouxiophyceae (Chlorophyta). Phycologia 57, 503524.CrossRefGoogle Scholar
Muggia, L, Nelsen, MP, Kirika, PM, Barreno, E, Beck, A, Lindgren, H, Lumbsch, HT, Leavitt, SD and Trebouxia working group (2020) Formally described species woefully underrepresent phylogenetic diversity in the common lichen photobiont genus Trebouxia (Trebouxiophyceae, Chlorophyta): an impetus for developing an integrated taxonomy. Molecular Phylogenetics and Evolution 149, 106821.CrossRefGoogle ScholarPubMed
Mukherjee, R, Borah, SP and Goswami, BC (2010) Biochemical characterization of carotenoids in two species of Trentepohlia (Trentepohliales, Chlorophyta). Journal of Applied Phycology 22, 569571.CrossRefGoogle Scholar
Mukhtar, A, Garty, J and Galun, M (1994) Does the lichen alga Trebouxia occur free-living in nature: further immunological evidence. Symbiosis 17, 247253.Google Scholar
Myllys, L, Stenroos, S, Thell, A and Kuusinen, M (2007) High cyanobiont selectivity of epiphytic lichens in old growth boreal forest of Finland. New Phytologist 173, 621629.CrossRefGoogle ScholarPubMed
Nakano, T (1988) Phycobionts of some Japanese species of the Graphidaceae. Lichenologist 20, 353360.CrossRefGoogle Scholar
Nakano, T and Iguchi, K (1994) Photobionts isolated from some Japanese species of Cladonia (lichens). Symbiosis 17, 6573.Google Scholar
Nakano, T and Ihda, T (1996) The identity of photobionts from the lichen Pyrenula japonica. Lichenologist 28, 437442.Google Scholar
Namba, N and Nakayama, T (2021) Taxonomic study of a new green alga, Annulotesta cochlephila gen. et sp. nov. (Kornmanniaceae, Ulvales, Ulvophyceae), growing on the shells of door snails. Journal of Plant Research 134, 7789.CrossRefGoogle Scholar
Nash, TH III, Kappen, L, Lösch, R, Larson, DW and Matthes-Sears, U (1987) Cold resistance of lichens with Trentepohlia- or Trebouxia-photobionts from the North American west coast. Flora 170, 241251.CrossRefGoogle Scholar
Nelsen, MP and Gargas, A (2008) Dissociation and horizontal transmission of codispersing lichen symbionts in the genus Lepraria (Lecanorales: Stereocaulaceae). New Phytologist 177, 264275.CrossRefGoogle Scholar
Nelsen, MP and Gargas, A (2009) Symbiont flexibility in Thamnolia vermicularis (Pertusariales: Icmadophilaceae). Bryologist 104, 404417.CrossRefGoogle Scholar
Nelsen, MP, Rivas Plata, E, Andrew, CJ, Lücking, R and Lumbsch, HT (2011) Phylogenetic diversity of trentepohlialean algae associated with lichen-forming fungi. Journal of Phycology 47, 282290.CrossRefGoogle ScholarPubMed
Nelson, DR, Khraiwesh, B, Fu, W, Alseekh, S, Jaiswal, A, Chaiboonchoe, A, Hazzouri, KM, O'Connor, MJ, Butterfoss, GL, Dro, N, et al. (2017) The genome and phenome of the green alga Chloroidium sp. UTEX 3007 reveal adaptive traits for desert acclimatization. eLife 6, e25783.CrossRefGoogle ScholarPubMed
Němcová, Y and Kalina, T (2000) Cell wall development, microfibril and pyrenoid structure in type strains of Chlorella vulgaris, C. kessleri, C sorokiniana compared with C. luteoviridis (Trebouxiophyceae, Chlorophyta). Algological Studies 100, 95105.Google Scholar
Neustupa, J (2015) Division Chlorophyta. In Frey, W (series ed.), Syllabus of Plant Families: 2/1 Photoautotrophic Eukaryotic Algae. Stuttgart: Borntraeger Verlagsbuchhandlung, pp. 191247.Google Scholar
Neustupa, J and Štifterová, S (2013) Distribution patterns of subaerial corticolous microalgae in two European regions. Plant Ecology and Evolution 146, 279289.CrossRefGoogle Scholar
Neustupa, J, Němcová, Y, Eliáš, M and Škaloud, P (2009) Kalinella bambusicola gen. et sp. nov. (Trebouxiophyceae, Chlorophyta), a novel coccoid Chlorella-like subaerial alga from Southeast Asia. Phycological Research 57, 159169.CrossRefGoogle Scholar
Nyati, S, Beck, A and Honegger, R (2007) Fine structure and phylogeny of green algal photobionts in the microfilamentous genus Psoroglaena (Verrucariaceae, lichen-forming ascomycetes). Plant Biology 9, 390399.CrossRefGoogle Scholar
Nyati, S, Bhattacharya, D, Werth, S and Honegger, R (2013) Phylogenetic analysis of LSU and SSU rDNA group I introns of lichen photobionts associated with the genera Xanthoria and Xanthomendoza (Teloschistaceae, lichenized Ascomycetes). Journal of Phycology 49, 11541166.CrossRefGoogle Scholar
Nyati, S, Scherrer, S, Werth, S and Honegger, R (2014) Green-algal photobiont diversity (Trebouxia spp.) in representatives of Teloschistaceae (Lecanoromycetes, lichen-forming ascomycetes). Lichenologist 46, 189212.CrossRefGoogle Scholar
O'Brien, H, Miadlikowsa, J and Lutzoni, F (2005) Assessing host specialization in symbiotic cyanobacteria associated with four closely related species of the lichen fungus Peltigera. European Journal of Phycology 40, 363378.CrossRefGoogle Scholar
O'Brien, H, Miadlikowsa, J and Lutzoni, F (2013) Assessing population structure and host specialization in lichenized cyanobacteria. New Phytologist 198, 557566.CrossRefGoogle ScholarPubMed
Oberwinkler, F (1980) Symbiotic relationship between fungus and alga in basidiolichens. In Schwemmler, W and Schenk, HEA (eds), Endocytobiology: Endosymbiosis and Cell Biology, Vol. I. Berlin: Walter de Gruyter, pp. 305315.CrossRefGoogle Scholar
Oberwinkler, F (1984) Fungus-alga interactions in basidiolichens. In Hertel, H and Oberwinkler, F (eds), Festschrift Josef Poelt: Beiheft Beiträge zur Lichenologie. Vaduz: J. Cramer, pp. 739774.Google Scholar
Oberwinkler, F (2012) Basidiolichens. In Hock, B (ed.), The Mycota IX. Fungal Associations, 2nd Edn. Berlin: Springer-Verlag, pp. 341362.CrossRefGoogle Scholar
Ohmura, Y, Kawachi, M, Kasai, F, Watanabe, MM and Takeshita, S (2006) Genetic combinations of symbionts in a vegetatively reproducing lichen, Parmotrema tinctorum, based on ITS rDNA sequences. Bryologist 109, 4359.CrossRefGoogle Scholar
Ohmura, Y, Takeshita, S and Kawachi, M (2019) Photobiont diversity within populations of a vegetatively reproducing lichen, Parmotrema tinctorum, can be generated by photobiont switching. Symbiosis 77, 5972.CrossRefGoogle Scholar
Oksanen, I, Lohtander, K, Paulsrud, P and Rikkinen, J (2002) A molecular approach to cyanobacterial diversity in a rock-pool community involving gelatinous lichens and free-living Nostoc colonies. Annales Botanici Fennici 39, 9399.Google Scholar
Onuţ-Brännström, I, Tibell, L and Johannesson, H (2017) A worldwide phylogeography of the whiteworm lichens Thamnolia reveals three lineages with distinct habitats and evolutionary histories. Ecology and Evolution 7, 36023615.CrossRefGoogle ScholarPubMed
Onuţ-Brännström, I, Benjamin, M, Scofield, DG, Heiđmarsson, S, Andersson, MGI, Lindström, ES and Johannesson, H (2018) Sharing of photobionts in sympatric populations of Thamnolia and Cetraria lichens: evidence from high-throughput sequencing. Scientific Reports 8, 4406.CrossRefGoogle ScholarPubMed
Ortiz-Álvarez, R, de los Ríos, A, Fernández-Mendoza, F, Torralba-Burrial, A and Pérez-Ortega, S (2015) Ecological specialization of two photobiont-specific maritime cyanolichen species of the genus Lichina. PLoS ONE 10, e0132718.CrossRefGoogle ScholarPubMed
Osyczka, P, Lenart-Borón, A, Borón, P and Rola, K (2021) Lichen-forming fungi in postindustrial habitats involve alternative photobionts. Mycologia 113, 4355.CrossRefGoogle ScholarPubMed
Otálora, MAG, Martínez, I, O'Brien, H, Molina, MC, Aragón, G and Lutzoni, F (2010) Multiple origins of high reciprocal symbiotic specificity at an intercontinental spatial scale among gelatinous lichens (Collemataceae, Lecanoromycetes). Molecular Phylogenetics and Evolution 56, 10891095.CrossRefGoogle Scholar
Otálora, MAG, Salvador, C, Martínez, I and Aragón, G (2013) Does the reproductive strategy affect the transmission and genetic diversity of bionts in cyanolichens? A case study using two closely related species. Microbial Ecology 65, 517530.CrossRefGoogle ScholarPubMed
Ott, S (1988) Photosymbiodemes and their development in Peltigera venosa. Lichenologist 20, 361368.CrossRefGoogle Scholar
Palmqvist, K, de los Ríos, A, Ascaso, C and Samuelsson, G (1997) Photosynthetic carbon acquisition in the lichen photobionts Coccomyxa and Trebouxia (Chlorophyta). Physiologia Plantarum 101, 6776.CrossRefGoogle Scholar
Paran, N, Ben-Shaul, Y and Galun, M (1971) Fine structure of the blue-green phycobiont and its relation to the mycobiont in two Gonohymenia lichens. Archiv für Mikrobiologie 76, 103113.CrossRefGoogle ScholarPubMed
Park, CH, Kim, KM, Elvebakk, A, Kim, O-S, Jeong, G and Hong, SG (2015) Algal and fungal diversity in Antarctic lichens. Journal of Eukaryotic Microbiology 62, 196205.CrossRefGoogle ScholarPubMed
Parra, OO and Redón, J (1977) Aislamiento de Heterococcus caespitosus Vischer ficobionte de Verrucaria maura. Boletin de la Sociedad de Biología de Concepción 51, 219224.Google Scholar
Paul, F, Otte, J, Schmitt, I and Dal Grande, F (2018) Comparing Sanger sequencing and high-throughput metabarcoding for inferring photobiont diversity in lichens. Scientific Reports 8, 8624.CrossRefGoogle ScholarPubMed
Paulsrud, P (2001) The Nostoc symbiosis of lichens: diversity, specificity and cellular modifications. Acta Universitatis Upsaliensis: Comprehensive Summaries of Uppsala Dissertations from the Faculty of Science and Technology 662, 155.Google Scholar
Paulsrud, P and Lindblad, P (1998) Sequence variation of the tRNALeu intron as a marker for genetic diversity and specificity of symbiotic cyanobacteria in some lichens. Applied and Environmental Microbiology 64, 310315.CrossRefGoogle ScholarPubMed
Paulsrud, P, Rikkinen, J and Lindblad, P (1998) Cyanobiont specificity in some Nostoc-containing lichens and in a Peltigera aphthosa photosymbiodeme. New Phytologist 139, 517524.CrossRefGoogle Scholar
Paulsrud, P, Rikkinen, J and Lindblad, P (2000) Spatial patterns of photobiont diversity in some Nostoc-containing lichens. New Phytologist 146, 291299.CrossRefGoogle ScholarPubMed
Paulsrud, P, Rikkinen, J and Lindblad, P (2001) Field investigations on cyanobacterial specificity in Peltigera aphthosa. New Phytologist 152, 117123.CrossRefGoogle Scholar
Peksa, O and Škaloud, P (2011) Do photobionts influence the ecology of lichens? A case study of environmental preferences in symbiotic green alga Asterochloris (Trebouxiophyceae). Molecular Ecology 20, 39363948.CrossRefGoogle Scholar
Pereira Riquelme, I (1992) Flora, vegetación y ecología de los líquenes acuáticos de España. Ph.D. thesis, University of Barcelona.Google Scholar
Pérez-Ortega, S, de los Ríos, A, Crespo, A and Sancho, LG (2010) Symbiotic lifestyle and phylogenetic relationships of the bionts of Mastodia tessellata (Ascomycota, incertae sedis). American Journal of Botany 97, 738752.CrossRefGoogle Scholar
Pérez-Ortega, S, Ortiz-Álvarez, R, Green, TGA and de los Ríos, A (2012) Lichen myco- and photobiont diversity and their relationships at the edge of life (McMurdo Dry Valleys, Antarctica). FEMS Microbiology Ecology 82, 429448.CrossRefGoogle Scholar
Pérez-Ortega, S, Miller, KA and de los Ríos, A (2018) Challenging the lichen concept: Turgidusculum ulvae (Verrucariaceae) represents an independent photobiont shift to a multicellular blade-like alga. Lichenologist 50, 341356.CrossRefGoogle Scholar
Peršoh, D, Beck, A and Rambold, G (2004) The distribution of ascus types and photobiontal selection in Lecanoromycetes (Ascomycota) against the background of a revised SSU nrDNA phylogeny. Mycological Progress 3, 103121.CrossRefGoogle Scholar
Peters, AF and Moe, RL (2001) DNA sequences confirm that Petroderma maculiforme (Phaeophyceae) is the brown algal phycobiont of the marine lichen Verrucaria tavaresiae (Verrucariales, Ascomycota) from central California. Bulletin of the California Lichen Society 8, 4143.Google Scholar
Peveling, E (1968) Pyrenoidstrukturen in symbiotisch lebenden Trebouxia-Arten. Zeitschrift für Pflanzenphysiologie 59, 393396.Google Scholar
Peveling, E (1969) Electronenoptische Untersuchungen an Flechten. II. Die Feinstruktur von Trebouxia-Phycobionten. Planta 87, 6985.CrossRefGoogle Scholar
Peveling, E and Galun, M (1976) Electron-microscopical studies on the phycobiont Coccomyxa Schmidle. New Phytologist 77, 713718.CrossRefGoogle Scholar
Piercey-Normore, MD (2004) Selection of algal genotypes by three species of lichen fungi in the genus Cladonia. Canadian Journal of Botany 82, 947961.CrossRefGoogle Scholar
Piercey-Normore, MD (2006) The lichen-forming ascomycete Evernia mesomorpha associates with multiple genotypes of Trebouxia jamesii. New Phytologist 169, 331344.CrossRefGoogle ScholarPubMed
Piercey-Normore, MD (2009) Vegetatively reproducing fungi in three genera of the Parmeliaceae share divergent algal partners. Bryologist 112, 773785.CrossRefGoogle Scholar
Piercey-Normore, MD and DePriest, PT (2001) Algal switching among lichen symbioses. American Journal of Botany 88, 14901498.CrossRefGoogle ScholarPubMed
Pino-Bodas, R and Stenroos, A (2020) Global biodiversity patterns of the photobionts associated with the genus Cladonia (Lecanorales, Ascomycota). Microbial Ecology, doi:10.1007/s00248-020-01633-3Google Scholar
Plessl, A (1963) Über die Beziehung von Haustorientypus und Organisationshöhe bei Flechten. Österreichische Botanische Zeitschrift 110, 194269.CrossRefGoogle Scholar
Poelt, J (1958) Über parasitische Flechten. II. Planta 51, 288307.CrossRefGoogle Scholar
Poelt, J and Baumgärtner, H (1964) Über Rhizinenstränge bei placodialen Flechten. Österreichische Botanische Zeitschrift 111, 118.CrossRefGoogle Scholar
Prieto, M, Schultz, M, Olariaga, I and Wedin, M (2019) Lichinodium is a new lichenized lineage in the Leotiomycetes. Fungal Diversity 94, 2339.CrossRefGoogle Scholar
Pröschold, T and Darienko, T (2020) The green puzzle Stichococcus (Trebouxiophyceae, Chlorophyta): new generic and species concept among this widely distributed genus. Phytotaxa 441, 113142.CrossRefGoogle Scholar
Pröschold, T, Darienko, T, Silva, P, Reisser, W and Krienitz, L (2011) The systematics of Zoochlorella revisited employing a molecular approach. Environmental Microbiology 13, 350364.CrossRefGoogle Scholar
Purvis, OW, Coppins, BJ, Hawksworth, DL, James, PW and Moore, DM (1992) The Lichen Flora of Great Britain and Ireland. London: British Lichen Society.Google Scholar
Raggio, J, Green, TGA, Crittenden, PD, Pintado, A, Vivas, M, Pérez-Ortega, S, de los Ríos, A and Sancho, LG (2012) Comparative ecology of three Placopsis species, pioneer lichens in recently exposed Chilean glacial forelands. Symbiosis 56, 5566.CrossRefGoogle Scholar
Rajaniemi, P, Hrouzek, P, Kaštovská, K, Willame, R, Rantala, A, Hoffmann, L, Komárek, J and Sivonen, K (2005) Phylogenetic and morphological evaluation of the genera Anabaena, Aphanizomenon, Trichormus and Nostoc (Nostocales, Cyanobacteria). International Journal of Systematic and Evolutionary Microbiology 55, 1126.CrossRefGoogle Scholar
Rambold, G, Friedl, T and Beck, A (1998) Photobionts in lichens: possible indicators of phylogenetic relationships? Bryologist 101, 392397.CrossRefGoogle Scholar
Ran, L, Larsson, J, Vigil-Stenman, T, Nylander, JAA, Ininbergs, K, Zheng, W-W, Lapidus, A, Lowry, S, Haselkorn, R and Bergman, B (2010) Genome erosion in a nitrogen-fixing vertically transmitted endosymbiotic multicellular cyanobacterium. PLoS ONE 5, e11486.CrossRefGoogle Scholar
Richardson, DHS, Hill, DJ and Smith, DC (1968) Lichen physiology. XI. The role of the alga in determining the pattern of carbohydrate movement between lichen symbionts. New Phytologist 67, 469486.CrossRefGoogle Scholar
Řidká, T, Peksa, O, Rai, H, Upreti, DK and Škaloud, P (2014) Photobiont diversity in Indian Cladonia lichens, with special emphasis on the geographical patterns. In Rai, H and Upreti, DK (eds), Terricolous Lichens in India. New York: Springer, pp. 5371.CrossRefGoogle Scholar
Rikkinen, J (2003) Ecological and evolutionary role of photobiont-mediated guilds in lichens. Symbiosis 34, 99110.Google Scholar
Rikkinen, J (2013) Molecular studies on cyanobacterial diversity in lichen symbiosis. MycoKeys 6, 332.CrossRefGoogle Scholar
Rikkinen, J (2017) Symbiotic cyanobacteria in lichens. In Grube, M, Seckbach, J and Muggia, L (eds), Algal and Cyanobacterial Symbioses. London: World Scientific Publishing Europe Ltd, pp. 147167.CrossRefGoogle Scholar
Rikkinen, J, Oksanen, I and Lohtander, K (2002) Lichen guilds share related cyanobacterial symbionts. Science 297, 357.CrossRefGoogle ScholarPubMed
Rindi, F and Guiry, MD (2002) Diversity, life history, and ecology of Trentepohlia and Printzina (Trentepohliales, Chlorophyta) in urban habitats in western Ireland. Journal of Phycology 38, 3954.CrossRefGoogle Scholar
Rindi, F, Menéndez, JL, Guiry, MD and Rico, JM (2004) The taxonomy and distribution of Phycopeltis (Trentepohliaceae, Chlorophyta) in Europe. Cryptogamie, Algologie 25, 317.Google Scholar
Rindi, F, Lam, DW and López-Bautista, JM (2009) Phylogenetic relationships and species circumscription in Trentepohlia and Printzina (Trentepohliales, Chlorophyta). Molecular Phylogenetics and Evolution 52, 329339.CrossRefGoogle Scholar
Rindi, F, Mikhailyuk, TI, Sluiman, HJ, Friedl, T and López-Bautista, JM (2011) Phylogenetic relationships in Interfilum and Klebsormidium (Klebsormidiophyceae, Streptophyta). Molecular Phylogenetics and Evolution 58, 218231.CrossRefGoogle Scholar
Rodgers, GA and Stewart, WDP (1977) The cyanophyte-hepatic symbiosis. New Phytologist 78, 441458.CrossRefGoogle Scholar
Rola, K, Lenart-Borón, A, Borón, P and Osyczka, P (2021) Heavy-metal pollution induces changes in the genetic composition and anatomical properties of photobionts in pioneer lichens colonizing post-industrial habitats. Science of the Total Environment 750, 141439.CrossRefGoogle Scholar
Romeike, J, Friedl, T, Helm, G and Ott, S (2002) Genetic diversity of algal and fungal partners in four species of Umbilicaria (lichenized ascomycetes) along a transect of the Antarctic Peninsula. Molecular Biology and Evolution 19, 12091217.CrossRefGoogle ScholarPubMed
Roskin, PA (1970) Ultrastructure of the host-parasite interaction in the basidiolichen Cora pavonia (Web.) E. Fries. Archiv für Mikrobiologie 70, 176182.CrossRefGoogle Scholar
Ruprecht, U, Brunauer, G and Printzen, C (2012) Genetic diversity of photobionts in Antarctic lecideoid lichens from an ecological viewpoint. Lichenologist 44, 661678.CrossRefGoogle Scholar
Ruprecht, U, Brunauer, G and Türk, R (2014) High photobiont diversity in the common European soil crust lichen Psora decipiens. Biodiversity and Conservation 23, 17711785.CrossRefGoogle ScholarPubMed
Ruprecht, U, Fernández-Mendoza, F, Türk, R and Fryday, AM (2020) High levels of endemism and local differentiation in the fungal and algal symbionts of saxicolous lecideoid lichens along a latitudinal gradient in southern South America. Lichenologist 52, 287303.CrossRefGoogle ScholarPubMed
Sadowska-Deś, AD, Dal Grande, F, Lumsch, HT, Beck, A, Otte, J, Hur, J-S, Kim, JA and Schmitt, I (2014) Integrating coalescent and phylogenetic approaches to delimit species in the lichen photobiont Trebouxia. Molecular Phylogenetics and Evolution 76, 202210.CrossRefGoogle ScholarPubMed
Sadowsky, A and Ott, S (2016) Symbiosis as a successful strategy in continental Antarctica: performance and protection of Trebouxia photosystem II in relation to lichen pigmentation. Polar Biology 39, 139151.CrossRefGoogle Scholar
Saini, KC, Nayaka, S and Bast, F (2019) Diversity of lichen photobionts: their coevolution and bioprospecting potential. In Satyanarayana, T, Johri, BN and Das, SK (eds), Microbial Diversity in Ecosystem Sustainability and Biotechnological Applications. Singapore: Springer Singapore, pp. 307323.CrossRefGoogle Scholar
Sanders, WB (1994) Role of lichen rhizomorphs in thallus propagation and substrate colonization. Cryptogamic Botany 4, 283289.Google Scholar
Sanders, WB (2001) Composite thalli of Sticta sp. from Brazil with morphologically similar lobes containing either a chlorobiont or a cyanobiont layer. Symbiosis 31, 4755.Google Scholar
Sanders, WB (2002) In situ development of the foliicolous lichen Phyllophiale (Trichotheliaceae) from propagule germination to propagule production. American Journal of Botany 89, 17411746.CrossRefGoogle ScholarPubMed
Sanders, WB (2005) Observing microscopic phases of lichen life cycles on transparent substrata placed in situ. Lichenologist 37, 373382.CrossRefGoogle Scholar
Sanders, WB (2014) Complete life cycle of the lichen fungus Calopadia puiggarii (Pilocarpaceae, Ascomycetes) documented in situ: propagule dispersal, establishment of symbiosis, thallus development, and formation of sexual and asexual reproductive structures. American Journal of Botany 101, 18361848.CrossRefGoogle ScholarPubMed
Sanders, WB and de los Ríos, A (2015) Structure and in situ development of the microlichen Gyalectidium paolae (Gomphillaceae, Ascomycota), an overlooked colonist on palm leaves in southwest Florida. American Journal of Botany 102, 14031412.CrossRefGoogle Scholar
Sanders, WB and Lücking, R (2002) Reproductive strategies, relichenization and thallus development observed in situ in leaf-dwelling lichen communities. New Phytologist 155, 425435.CrossRefGoogle ScholarPubMed
Sanders, WB and Rico, VJ (1992) Lichenizing rhizomorphs and thallus development in the squamulose lichen Aspicilia crespiana Rico ined. (Lecanorales, Ascomycetes). Botanica Acta 105, 449456.CrossRefGoogle Scholar
Sanders, WB, Wierzchos, J and Ascaso, C (1994) Physical interactions of two rhizomorph-forming lichens with their rock substrate. Acta Botanica 107, 432439.CrossRefGoogle Scholar
Sanders, WB, Moe, RL and Ascaso, C (2004) The intertidal marine lichen formed by the pyrenomycete fungus Verrucaria tavaresiae and the brown alga Petroderma maculiforme (Phaeophyceae): thallus organization and symbiont interaction. American Journal of Botany 91, 511522.CrossRefGoogle ScholarPubMed
Sanders, WB, Moe, RL and Ascaso, C (2005) Ultrastructural study of the brown alga Petroderma maculiforme (Phaeophyceae) in the free-living state and in lichen symbiosis with the intertidal marine fungus Verrucaria tavaresiae (Ascomycotina). European Journal of Phycology 40, 353361.CrossRefGoogle Scholar
Sanders, WB, Pérez-Ortega, S, Nelsen, MP, Lücking, R and de los Ríos, A (2016) Heveochlorella (Trebouxiophyceae): a little-known genus of unicellular green algae outside the Trebouxiales emerges unexpectedly as a major clade of lichen photobionts in foliicolous communities. Journal of Phycology 52, 840853.CrossRefGoogle ScholarPubMed
Santesson, R (1952) Foliicolous lichens I. A revision of the taxonomy of the obligately foliicolous, lichenized fungi. Symbolae Botanicae Upsaliensis 12, 1590.Google Scholar
Schaper, T and Ott, S (2003) Photobiont selectivity and interspecific interactions in lichen communities. I. Culture experiments with the mycobiont Fulgensia bracteata. Plant Biology 5, 441450.CrossRefGoogle Scholar
Scheidegger, C (1995) Reproductive strategies in Vezdaea (Lecanorales, lichenized Ascomycetes): a low-temperature scanning electron microscope study of a ruderal species. Cryptogamic Botany 5, 163171.Google Scholar
Schmitt, I and Lumbsch, HT (2001) Identification of the photobionts in Trapeliopsis and Pertusaria using SSU ribosomal DNA sequences obtained from PCR amplification with a non-green-algal primer. Mycotaxon 78, 407411.Google Scholar
Schüßler, A (2012) The GeosiphonNostoc endosymbiosis and its role as a model for arbuscular mycorrhiza research. In Hock, B (ed.), The Mycota IX. Fungal Associations, 2nd Edn. Berlin: Springer-Verlag, pp. 7791.CrossRefGoogle Scholar
Schuster, G (1992) Development of adventive thalli in Umbilicaria Hoffm. Flora 187, 201207.CrossRefGoogle Scholar
Schuster, G, Ott, S and Jahns, HM (1985) Artificial culture of lichens in the natural environment. Lichenologist 17, 147153.CrossRefGoogle Scholar
Schwendener, S (1869) Die Algentypen der Flechtengonidien. Basel: Universitaetsbuchdruckerei von C. Schultze.Google Scholar
Scott, GD (1957) Lichen terminology. Nature 179, 486487.CrossRefGoogle ScholarPubMed
Scott, GD (1960) Studies of the lichen symbiosis I. The relationship between nutrition and moisture content in the maintenance of the symbiotic state. New Phytologist 59, 374381.CrossRefGoogle Scholar
Seto, K, Matsuzawa, T, Kuno, H and Kagami, M (2020) Morphology, ultrastructure, and molecular phylogeny of Aphelidium collabens sp. nov. (Aphelida), a parasitoid of a green alga Coccomyxa sp. Protist 171, 125728.CrossRefGoogle Scholar
Silva, PC (1979) Review of the taxonomic history and nomenclature of the yellow-green algae. Archiv für Protistenkunde 121, 2063.CrossRefGoogle Scholar
Simon, A, Goffinet, B, Magain, N and Sérusiaux, E (2018) High diversity, high insular endemism and recent origin in the lichen genus Sticta (lichenized Ascomycota, Peltigerales) in Madagascar and the Mascarenes. Molecular Phylogenetics and Evolution 122, 1528.CrossRefGoogle ScholarPubMed
Singh, G, Dal Grande, F, Divakar, PK, Otte, J, Crespo, A and Schmitt, I (2017) Fungal-algal association patterns in lichen symbiosis linked to macroclimate. New Phytologist 214, 317329.CrossRefGoogle ScholarPubMed
Škaloud, P and Peksa, O (2010) Evolutionary inferences based on ITS rDNA and actin sequences reveal extensive diversity of the common lichen alga Asterochloris (Trebouxiophyceae, Chlorophyta). Molecular Phylogenetics and Evolution 54, 3646.CrossRefGoogle Scholar
Škaloud, P, Neustupa, J, Radochová, B and Kubínová, L (2005) Confocal microscopy of chloroplast morphology and ontogeny in three strains of Dictyochloropsis (Trebouxiophyceae, Chlorophyta). Phycologia 44, 261269.CrossRefGoogle Scholar
Škaloud, P, Steinová, J, Řidká, T and Vančurová, L (2015) Assembling the challenging puzzle of algal biodiversity: species delimitation within the genus Asterochloris (Trebouxiophyceae, Chlorophyta). Journal of Phycology 51, 507527.CrossRefGoogle Scholar
Škaloud, P, Friedl, T, Hallmann, C, Beck, A and Dal Grande, F (2016) Taxonomic revision and species delimitation of coccoid green algae currently assigned to the genus Dictyochloropsis (Trebouxiophyceae, Chlorophyta). Journal of Phycology 52, 599617.CrossRefGoogle Scholar
Škaloud, P, Moya, P, Molins, A, Peksa, O, Santos-Guerra, A and Barreno, E (2018) Untangling the hidden intrathalline microalgal diversity in Parmotrema pseudotinctorum: Trebouxia crespoana sp. nov. Lichenologist 50, 357369.CrossRefGoogle Scholar
Skuja, H (1943) Ein Fall von fakultativer Symbiose zwischen operculatem Discomycet und einer Chlmydomonade. Archiv für Protistenkunde 96, 365376.Google Scholar
Slocum, RD (1980) Light and electron microscopic investigations in the Dictyonemataceae (basidiolichens) II. Dictyonema irpicinum. Canadian Journal of Botany 58, 10051015.CrossRefGoogle Scholar
Slocum, RD, Ahmadjian, V and Hildreth, KC (1980) Zoosporogenesis in Trebouxia gelatinosa: ultrastructure potential for zoospore release and implications for the lichen association. Lichenologist 12, 173187.CrossRefGoogle Scholar
Smith, DC (1974) Transport from symbiotic algae and symbiotic chloroplasts to host cells. Symposia of the Society for Experimental Biology 28, 485520.Google Scholar
Smith, DC (2019) What can lichens tell us about ‘real fungi’? Symbiosis 77, 9398.Google Scholar
Smith, DC and Drew, EA (1965) Studies in the physiology of lichens. V. Translocation from the algal layer to the medulla in Peltigera polydactyla. New Phytologist 64, 195200.CrossRefGoogle Scholar
Smith, EC and Griffiths, H (1996) The occurrence of the chloroplast pyrenoid is correlated with the activity of a CO2-concentrating mechanism and carbon isotope discrimination in lichens and bryophytes. Planta 198, 616.CrossRefGoogle Scholar
Smith, H, Dal Grande, F, Muggia, L, Keuler, R, Divakar, PK, Grewe, F, Schmitt, I, Lumbsch, HT and Leavitt, SD (2020) Metagenomic data reveal diverse fungal and algal communities associated with the lichen symbiosis. Symbiosis 82, 133147.CrossRefGoogle Scholar
Solhaug, KA and Gauslaa, Y (1996) Parietin, a photoprotective secondary product of the lichen Xanthoria parietina. Oecologia 108, 412418.CrossRefGoogle ScholarPubMed
Solovchenko, AE (2013) Physiology and adaptive significance of secondary carotenogenesis in green microalgae. Russian Journal of Plant Physiology 60, 113.CrossRefGoogle Scholar
Spribille, T, Tønsberg, T, Stabentheiner, E and Muggia, L (2014) Reassessing evolutionary relationships in the filamentous cyanolichen genus Spilonema (Peltigerales, Lecanoromycetes). Lichenologist 46, 373388.CrossRefGoogle Scholar
Spribille, T, Tuovinen, V, Resl, P, Vanderpool, D, Wolinski, H, Aime, MC, Schneider, K, Stabentheiner, E, Toome-Heller, M, Thor, G, et al. (2016) Basidiomycete yeasts in the cortex of ascomycete macrolichens. Science 353, 488492.CrossRefGoogle ScholarPubMed
Stahl, E (1877) Beiträge zur Entwicklungsgeschichte der Flechten. II. Ueber die Bedeutung der Hymenialgonidien. Leipzig: Arthur Felix.Google Scholar
Steinová, J, Škaloud, P, Yahr, R, Bestová, H and Muggia, L (2019) Reproductive and dispersal strategies shape the diversity of mycobiont-photobiont association in Cladonia lichens. Molecular Phylogenetics and Evolution 134, 226237.CrossRefGoogle ScholarPubMed
Stenroos, S, Stocker-Wörgötter, E, Yoshimura, I, Myllys, L, Thell, A and Hyvönen, J (2003) Culture experiments and DNA sequence data confirm the identity of Lobaria photomorphs. Canadian Journal of Botany 81, 232247.CrossRefGoogle Scholar
Stenroos, S, Högnabba, F, Myllys, L, Hyvönen, J and Thell, A (2006) High selectivity in symbiotic associations of lichenized ascomycetes and cyanobacteria. Cladistics 22, 230238.CrossRefGoogle Scholar
Stevenson, RN and South, GR (1974) Coccomyxa parasitica sp. nov. (Coccomyxaceae, Chlorococcales), a parasite of giant scallops in Newfoundland. British Phycological Journal 9, 319329.CrossRefGoogle Scholar
Stocker-Wörgötter, E (2001) Experimental lichenology and microbiology of lichens: culture experiments, secondary chemistry of cultured mycobionts, resynthesis, and thallus morphogenesis. Bryologist 104, 576581.CrossRefGoogle Scholar
Stocker-Wörgötter, E and Hager, A (2008) Culture methods for lichens and lichen symbionts. In Nash, TH III (ed.), Lichen Biology. New York: Cambridge University Press, pp. 353363.CrossRefGoogle Scholar
Summerer, M, Sonntag, B and Sommaruga, R (2008) Ciliate-symbiont specificity of freshwater endosymbiotic Chlorella (Trebouxiophyceae, Chlorophyta). Journal of Phycology 44, 7784.CrossRefGoogle Scholar
Summerfield, TC and Eaton-Rye, JJ (2006) Pseudocyphellaria crocata, P. neglecta and P. perpetua from the Northern and Southern Hemispheres are a phylogenetic species and share cyanobionts. New Phytologist 170, 597607.CrossRefGoogle ScholarPubMed
Suto, Y and Ohtani, S (2009) Morphology and taxonomy of five Cephaleuros species (Trentepohliaceae, Chlorophyta) from Japan, including three new species. Phycologia 48, 213236.CrossRefGoogle Scholar
Suto, Y and Ohtani, S (2013) Seasonal development of five Cephaleuros species (Trentepohliaceae, Chlorophyta) on the leaves of woody plants and the behaviors of their gametes and zoospores. Phycological Research 61, 105115.CrossRefGoogle Scholar
Svenning, MM, Eriksson, T and Rasmussen, U (2005) Phylogeny of symbiotic cyanobacteria within the genus Nostoc based on 16S rDNA sequence analyses. Archives of Microbiology 183, 1926.CrossRefGoogle ScholarPubMed
Syasina, IG, Kukhlevsky, AD, Kovaleva, AL and Vaschenko, MA (2012) Phylogenetic and morphological characterization of the green alga infesting the horse mussel Modiolus from Vityaz Bay (Peter the Great Bay, Sea of Japan). Journal of Invertebrate Pathology 111, 175181.CrossRefGoogle Scholar
Takeshita, S (2001) A taxonomic revision of the genus Trebouxia. Hikobia 13, 425455.Google Scholar
Takeshita, S, Nakano, T and Iwatsuki, Z (1989) Phycobionts of some Japanese species of Pertusaria (Pertusariaceae). Plant Systematics and Evolution 165, 4954.CrossRefGoogle Scholar
Takeshita, S, Tokizawa, M, Handa, S and Okamoto, T (2010) A report of the photobiont isolated from Multiclavula clara (Berk. & Curt.) R. H. Peterson (Clavariaceae, lichenized Basidiomycetes). Hikobia 15, 493497. [In Japanese with English abstract and figure legends]Google Scholar
Thomas, EA (1939) Über die Biologie von Flechtenbildnern. Beiträge zur Kryptogamenflora der Schweiz 9, 1208 (+ 6 tab.).Google Scholar
Thompson, RH and Wujek, DE (1992) Printzina gen. nov. (Trentepohliaceae), including a description of a new species. Journal of Phycology 28, 232237.CrossRefGoogle Scholar
Thompson, RH and Wujek, DE (1997) Trentepohliales: Cephaleuros, Phycopeltis and Stomatochroon. Morphology Taxonomy and Ecology. Enfield, New Hampshire: Science Publishers.Google Scholar
Thüs, H, Muggia, L, Pérez-Ortega, S, Favero-Longo, SE, Joneson, S, O'Brien, H, Nelsen, MP, Duque-Thüs, R, Grube, M, Friedl, T, et al. (2011) Revisiting photobiont diversity in the lichen family Verrucariaceae (Ascomycota). European Journal of Phycology 46, 399415.CrossRefGoogle Scholar
Tibell, L (2001) Photobiont association and molecular phylogeny of the lichen genus Chaenotheca. Bryologist 104, 191198.CrossRefGoogle Scholar
Tibell, L and Beck, A (2002) Morphological variation, photobiont association and ITS phylogeny of Chaenotheca phaeocephala and C. subroscida (Coniocybaceae, lichenized ascomycetes). Nordic Journal of Botany 21, 651660.CrossRefGoogle Scholar
Tønsberg, T and Goward, T (2001) Sticta oroborealis sp. nov., and other Pacific North American lichens forming dendriscocauloid cyanotypes. Bryologist 104, 1223.CrossRefGoogle Scholar
Tønsberg, T and Holtan-Hartwig, J (1983) Phycotype pairs in Nephroma, Peltigera and Lobaria in Norway. Nordic Journal of Botany 3, 681688.CrossRefGoogle Scholar
Trembley, ML, Ringli, C and Honegger, R (2002 a) Differential expression of hydrophobins DGH1, DGH2 and DGH3 and immunolocalization of DGH1 in strata of the lichenized basidiocarp of Dictyonema glabratum. New Phytologist 154, 185195.CrossRefGoogle Scholar
Trembley, ML, Ringli, C and Honegger, R (2002 b) Morphological and molecular analysis of early stages in the resynthesis of the lichen Baeomyces rufus. Mycological Research 106, 768776.CrossRefGoogle Scholar
Trémouillaux-Guiller, J and Huss, VAR (2007) A cryptic intracellular green alga in Ginkgo biloba: ribosomal DNA markers reveal worldwide distribution. Planta 226, 553557.CrossRefGoogle ScholarPubMed
Trémouillaux-Guiller, J, Rohr, T, Rohr, R and Huss, VAR (2002) Discovery of an endophytic alga in Ginkgo biloba. American Journal of Botany 89, 727733.CrossRefGoogle ScholarPubMed
Tschaikner, A, Ingolić, E, Holzinger, A and Gärtner, G (2007) Phycobionts of some species of Evernia and Ramalina. Herzogia 20, 5360.Google Scholar
Tschermak, E (1941 a) Untersuchungen über die Beziehung von Pilz und Alge im Flechtenthallus. Österreichische Botanische Zeitschrift 90, 233307.CrossRefGoogle Scholar
Tschermak, L (1941 b) Beitrag zur Entwicklungsgeschichte und Morphologie der Protococcale Trochiscia granulata. Österreichische Botanische Zeitschrift 90, 6773.CrossRefGoogle Scholar
Tschermak-Woess, E (1953) Über wenig bekannte und neue Flechtengonidien III. Die Entwicklungsgeschichte von Leptosira thrombii nov. spec., der Gonidie von Thrombium epigaeum. Österreichische Botanische Zeitschrift 100, 203216.CrossRefGoogle Scholar
Tschermak-Woess, E (1970) Über wenig bekannte und neue Flechtengonidien V. Der Phycobiont von Verrucaria aquatilis und die Fortpflanzung von Pseudopleurococcus arthopyreniae. Österreichische Botanische Zeitschrift 118, 443455.CrossRefGoogle Scholar
Tschermak-Woess, E (1976) Algal taxonomy and the taxonomy of lichens: the phycobiont of Verrucaria adriatica. In Brown, DH, Hawksworth, DL and Bailey, RL (eds), Lichenology: Progress and Problems. London: Academic Press, pp. 7988.Google Scholar
Tschermak-Woess, E (1978) Myrmecia reticulata as a phycobiont and free-living – free-living Trebouxia – the problem of Stenocybe septata. Lichenologist 10, 6979.CrossRefGoogle Scholar
Tschermak-Woess, E (1980 a) Asterochloris phycobiontica, gen. et spec. nov., der Phycobiont der Flechte Varicellaria carneonivea. Plant Systematics and Evolution 135, 279294.CrossRefGoogle Scholar
Tschermak-Woess, E (1980 b) Elliptochloris bilobata, gen. et spec. nov., der Phycobiont von Catolechia wahlenbergii. Plant Systematics and Evolution 136, 6372.CrossRefGoogle Scholar
Tschermak-Woess, E (1981) Zur Kenntnis der Phycobionten von Lobaria linita und Normandina pulchella. Nova Hedwigia 35, 6373.Google Scholar
Tschermak-Woess, E (1983) Das Haustorialsystem von Dictyonema kennzeichend für die Gattung. Plant Systematics and Evolution 143, 109115.CrossRefGoogle Scholar
Tschermak-Woess, E (1985) Elliptochloris bilobata kein ganz seltener Photobiont. Herzogia 7, 105116.Google Scholar
Tschermak-Woess, E (1988 a) The algal partner. In Galun, M (ed.), CRC Handbook of Lichenology. Boca Raton, Florida: CRC Press, pp. 3992.Google Scholar
Tschermak-Woess, E (1988 b) New and known taxa of Chlorella (Chlorophyceae): occurrence as lichen phycobionts and observations on living dictyosomes. Plant Systematics and Evolution 159, 123139.CrossRefGoogle Scholar
Tschermak-Woess, E (1989) Developmental studies in trebouxioid algae and taxonomical consequences. Plant Systematics and Evolution 164, 161195.CrossRefGoogle Scholar
Tschermak-Woess, E (1995) Dictyochloropsis splendida (Chlorophyta), the correct phycobiont of Phlyctis argena and the high degree of selectivity involved. Lichenologist 27, 169187.CrossRefGoogle Scholar
Tschermak-Woess, E and Poelt, J (1976) Vezdaea, a peculiar lichen genus, and its phycobiont. In Brown, DH, Hawksworth, DL and Bailey, RL (eds), Lichenology: Progress and Problems. London: Academic Press, pp. 89105.Google Scholar
Tschermak-Woess, E, Bartlett, J and Peveling, E (1983) Lichenothrix riddlei is an ascolichen and also occurs in New Zealand – light and electron microscopical investigation. Plant Systematics and Evolution 143, 293309.CrossRefGoogle Scholar
Tzovaras, BG, Segers, FHID, Bicker, A, Dal Grande, F, Otte, J, Anvar, SY, Hankeln, T, Schmitt, I and Ebersberger, I (2020) What is in Umbilicaria pustulata? A metagenomic approach to reconstruct the holo-genome of a lichen. Genome Biology and Evolution 12, 309324.CrossRefGoogle Scholar
Uher, B (2008) Spatial distribution of cyanobacteria and algae from the tombstone in a historic cemetery in Bratislava, Slovakia. Fottea 9, 8192.CrossRefGoogle Scholar
van den Hoek, C, Mann, DG and Jahns, HM (1995) Algae: An Introduction to Phycology. Cambridge: Cambridge University Press.Google Scholar
Vančurová, L (2012) Diverzita fotobiontů ve stélkách lišejníků rodu Stereocaulon [Photobiont diversity in the lichen genus Stereocaulon (Lecanoromycetes, Ascomycota)]. Master's thesis, Charles University, Prague. [In Czech, with English summary].Google Scholar
Vančurová, L, Peksa, O, Němcová, Y and Škaloud, P (2015) Vulcanochloris (Trebouxiales, Trebouxiophyceae), a new genus of lichen photobiont from La Palma, Canary Islands, Spain. Phytotaxa 219, 118132.CrossRefGoogle Scholar
Vančurová, L, Muggia, L, Peksa, O, Řídká, T and Škaloud, P (2018) The complexity of symbiotic interactions influences the ecological amplitude of the host: a case study in Stereocaulon (lichenized Ascomycota). Molecular Ecology 27, 30163033.CrossRefGoogle Scholar
Vargas Castillo, R and Beck, A (2012) Photobiont selectivity and specificity in Caloplaca species in a fog-induced community in the Atacama Desert, northern Chile. Fungal Biology 116, 665676.CrossRefGoogle Scholar
Villanueva, CD, Hašler, P, Dvořák, P, Poulíčková, A and Casamatta, DA (2018) Brasilonema lichenoides sp. nov. and Chroococcidiopsis lichenoides sp. nov. (Cyanobacteria): two novel cyanobacterial constituents isolated from a tripartite lichen of headstones. Journal of Phycology 54, 224233.CrossRefGoogle ScholarPubMed
Vischer, W (1960) Reproduktion und systematische Stellung einiger Rinden- und Bodenalgen. Schweizerische Zeitschrift für Hydrologie 22, 330349.Google Scholar
, GTP (2016) Cyanobacterial lichenized fungi and their photobionts in Vietnam. Ph.D. thesis, Technische Universität Kaiserslautern.Google Scholar
Voytsekhovich, A and Beck, A (2016) Lichen photobionts of the rocky outcrops of Karadag massif (Crimean Peninsula). Symbiosis 68, 924.CrossRefGoogle Scholar
Voytsekhovich, A, Dymytrova, L and Nadyeina, O (2011) Photobiont composition of some taxa of the genera Micarea and Placynthiella (Lecanoromycetes, lichenized Ascomycota) from Ukraine. Folia Cryptogamica Estonica 48, 135148.Google Scholar
Wagner, J and Létrouit-Galinou, M-A (1988) Structure et ontogenèse du thalle squamuleux du lichen Endocarpon pusillum. Canadian Journal of Botany 66, 21182127.CrossRefGoogle Scholar
Wagner, M, Bathke, AC, Cary, CS, Green, TGA, Junker, RR, Trutschnig, W and Ruprecht, U (2020) Myco- and photobiont associations in crustose lichens in the McMurdo Dry Valleys (Antarctica) reveal high differentiation along an elevational gradient. Polar Biology 43, 19671983.CrossRefGoogle Scholar
Ward, HM (1884) On the structure, development and life history of a tropical epiphyllous lichen (Strigula complanata Fée, fide Rev. J. M. Crombie). Transactions of the Linnean Society of London 2, 87119.Google Scholar
Warén, H (1920) Reinkultur von Flechtengonidien. Finska Vetenskaps-Societetens Förhandlingar 61, 179.Google Scholar
Watanabe, S, Nakano, T and Deguchi, H (1997) Photobionts isolated from maritime lichens. Journal of Marine Biotechnology 5, 103112.Google Scholar
Waterbury, JB and Stanier, RY (1978) Pattern of growth and development in pleurocapsalean cyanobacteria. Microbiological Reviews 42, 244.CrossRefGoogle Scholar
Wedin, M, Maier, S, Fernández-Brime, S, Cronholm, B, Westberg, M and Grube, M (2016) Microbiome change by symbiotic invasion in lichens. Environmental Microbiology 18, 14281439.CrossRefGoogle ScholarPubMed
Werner, R-G (1931) Histoire de la synthèse liquénique. Mémoires de la Société des Science Naturelles du Maroc 27, 745.Google Scholar
Werth, S (2010) Biogeography and phylogeography of lichen fungi and their photobionts. In Fontaneto, D (ed.), Biogeography of Microscopic Organisms: Is Everything Small Everywhere? Cambridge: Cambridge University Press, pp. 191208.Google Scholar
Werth, S and Scheidegger, C (2012) Congruent genetic structure in the lichen-forming fungus Lobaria pulmonaria and its green-algal photobiont. Molecular Plant-Microbe Interactions 25, 220230.CrossRefGoogle ScholarPubMed
Werth, S and Sork, VL (2014) Ecological specialization in Trebouxia (Trebouxiophyceae) photobionts of Ramalina menziesii (Ramalinaceae) across six range-covering ecoregions of western North America. American Journal of Botany 101, 11271140.CrossRefGoogle ScholarPubMed
Wetmore, C (1970) The lichen family Heppiaceae in North America. Annals of the Missouri Botanical Gardens 57, 158209.CrossRefGoogle Scholar
Williams, L, Colesie, C, Ullmann, A, Westberg, M, Wedin, M and Büdel, B (2017) Lichen acclimation to changing environments: photobiont switching vs. climate-specific uniqueness in Psora decipiens. Ecology and Evolution 7, 25602574.CrossRefGoogle ScholarPubMed
Wirtz, N, Lumbsch, HT, Green, TGA, Türk, R, Pintado, A, Sancho, LG and Schroeter, B (2003) Lichen fungi have low cyanobiont selectivity in maritime Antarctica. New Phytologist 160, 177183.CrossRefGoogle ScholarPubMed
Withrow, K and Ahmadjian, V (1983) The ultrastructure of Chiodecton sanguineum. Mycologia 75, 337339.CrossRefGoogle Scholar
Wolf, M, Hepperle, D and Krienitz, L (2003) On the phylogeny of Radiococcus, Planktosphaeria and Schizochlamydella (Radiococcaceae, Chlorophyta). Biologia, Bratislava 58, 759765.Google Scholar
Wolk, CP, Ernst, A and Elhai, J (1994) Heterocyst metabolism and development. In Bryant, DA (ed.), The Molecular Biology of Cyanobacteria. Dordrecht: Kluwer Academic Publishers, pp. 769823.CrossRefGoogle Scholar
Wornik, S and Grube, M (2010) Joint dispersal does not imply maintenance of partnerships in lichen symbioses. Microbial Ecology 59, 150157.CrossRefGoogle Scholar
Wynne, MJ (1969) Life history and systematic studies of some Pacific North American Phaeophyceae (brown algae). University of California Publications in Botany 50, 188.Google Scholar
Xu, H, Deckert, RJ and Garbary, DJ (2008) Ascophyllum and its symbionts. X. Ultrastructure of the interaction between A. nodosum (Phaeophyceae) and Mycophycias ascophylli (Ascomycetes). Botany 86, 185193.CrossRefGoogle Scholar
Xu, M, De Boer, H, Olafssdottir, ES, Omarsdottir, S and Heiđmarsson, S (2020) Phylogenetic diversity of the lichenized algal genus Trebouxia (Trebouxiophyceae, Chlorophyta): a new lineage and novel insights from fungal-algal association patterns of Icelandic cetrarioid lichens (Parmeliaceae, Ascomycota). Botanical Journal of the Linnean Society 194, 460468.CrossRefGoogle Scholar
Yahr, R, Vilgalys, R and DePriest, PT (2004) Strong fungal specificity and selectivity for algal symbionts in Florida scrub Cladonia lichens. Molecular Ecology 13, 33673378.CrossRefGoogle ScholarPubMed
Yahr, R, Vilgalys, R and DePriest, PT (2006) Geographic variation in algal partners of Cladonia tenuis (Cladoniaceae) highlights the dynamic nature of a lichen symbiosis. New Phytologist 171, 847860.CrossRefGoogle Scholar
Yahr, R, Florence, A, Škaloud, P and Voytsekhovich, A (2015) Molecular and morphological diversity in photobionts associated with Micarea s. str. (Lecanorales, Ascomycota). Lichenologist 47, 403414.CrossRefGoogle Scholar
Yung, CCM, Chan, Y, Lacap, DC, Pérez-Ortega, S, de los Ríos, A, Lee, CK, Cary, SC and Pointing, S (2014) Characterization of chasmoendolithic community in Miers Valley, McMurdo Dry Valleys, Antarctica. Microbial Ecology 68, 351359.CrossRefGoogle ScholarPubMed
Zahlbruckner, A (1907) Specieller Teil: Ascolichenes (Schlauchflechten); Hymenolichenes (Basidiomycetenflechten). In Engler, A and Prantl, K (eds), Die Naturlichen Pflanzenfamilien. I. Teil, 1. Abteilung*. Leipzig: Wilhelm Engelmann, pp. 49243.Google Scholar
Zahradníková, M, Andersen, HL, Tønsberg, T and Beck, A (2017) Molecular evidence of Apatococcus, including A. fuscideae sp. nov., as photobiont in the genus Fuscidea. Protist 168, 425438.CrossRefGoogle ScholarPubMed
Zeitler, I (1954) Untersuchungen über die Morphologie, Entwicklungsgeschichte und Systematik von Flechtengonidien. Österreichische Botanische Zeitschrift 101, 453487.CrossRefGoogle Scholar
Zhang, J, Huss, VAR, Sun, X, Chang, K and Pang, D (2008) Morphology and phylogenetic position of a trebouxiophycean green alga (Chlorophyta) growing on the rubber tree, Hevea brasiliensis, with the description of a new genus and species. European Journal of Phycology, 43, 185193.CrossRefGoogle Scholar
Zhaxybayeva, O, Gogrten, JP, Charlebois, RL, Doolittle, WF and Papke, RT (2006) Phylogenetic analyses of cyanobacterial genomes: quantification of horizontal gene transfer events. Genome Research 16, 10991108.CrossRefGoogle ScholarPubMed
Zhu, H, Zhao, Z, Xia, S, Hu, Z and Liu, G (2015) Morphological examination and phylogenetic analyses of Phycopeltis spp. (Trentepohliales, Ulvophyceae) from tropical China. PLoS ONE 10, e0114936.Google ScholarPubMed
Zhu, H, Hu, Z and Liu, G (2017) Morphology and molecular phylogeny of Trentepohliales (Chlorophyta) from China. European Journal of Phycology 52, 330341.CrossRefGoogle Scholar
Zhu, H, Li, S, Hu, Z and Liu, G (2018) Molecular characterization of eukaryotic algal communities in the tropical phyllosphere based on real-time sequencing of the 18S rDNA gene. BMC Plant Biology 18, 365.CrossRefGoogle ScholarPubMed
Zoller, S and Lutzoni, F (2003) Slow algae, fast fungi: exceptionally high nucleotide substitution rate differences between lichenized fungi Omphalina and their symbiotic green algae Coccomyxa. Molecular Phylogenetics and Evolution 29, 629640.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. Three filamentous lichen photobiont genera in aposymbiotic and symbiotic states. A–C, Trentepohlia. A, branching filament free-living on bark. B, lichenized by Coenogonium hyphae (arrows) growing over morphologically unchanged algal filament and its new branches (horizontal arrow). C, lichenized by Arthonia rubrocincta; the alga is largely broken up into individual cells or short segments. D–F, Rhizonema. D, cultured isolate from Dictyonema; note false branching (arrowhead). E, trichome ensheathed by cells of mycobiont Dictyonema. F, contorted or broken filaments (arrow) within thallus of Coccocarpia palmicola. G–J. Nostoc. G, free-living thallus-like macrocolony on soil. H, cultured strain. I, more or less intact filaments (arrows) within thallus of Collema furfuraceum. J, contorted or broken up into cell groups (arrows) within cyanomorph of Sticta canariensis. Scales: A–F, H–J = 10 μm; G = 1 cm.

Figure 1

Table 1. Taxonomically grouped list of photobiont genera and mycobionts reported in association with them. The family names of the mycobionts are included in places where emphasis might be useful. id = procedures used in the study to identify the photobiont. LM = light microscopy, TEM = transmission electron microscopy. See table 1 in Tschermak-Woess (1988a) for a comprehensive list of photobiont reports prior to 1988. Taxon names follow those used in the original articles.

Figure 2

Fig. 2. TEM micrographs of some photobiont pyrenoids, with plastoglobuli (round black dots) and penetrating membranes in various positions and orientations. A, Trebouxia, within thallus of Lasallia pustulata. Note pyrenoid structure here more closely resembles that of distantly related Heveochlorella (B) than that of another species (C) of Trebouxia. B, Heveochlorella, within thallus of Calopadia. C, Trebouxia, within thallus of Ramalina usnea. D, bulging exserted pyrenoid of Petroderma maculiforme. E, Diplosphaera, within thallus of Endocarpon pusillum. S = starch grain or plates. Scales: A = 1 μm; B = 200 nm; C–E = 500 nm.

Figure 3

Fig. 3. Liberation and potential co-dispersal of photobionts from the spore-producing structures of certain mycobionts. A, Diplosphaera photobiont (arrows) within perithecium of Endocarpon pusillum; note much smaller size compared to photobiont cells within thalline tissue (t); s = ascospore. B, apothecial surface of foliicolous lichen colonizing plastic cover slip; note epithecial algal cells (arrows) among emerging ascospores (s). C, Heveochlorella photobionts (vertical arrow) within conidiogenous tissue of campylidia and intermixed among filiform macroconidia (oblique arrow). D, hyphophore of Gyalectidium paolae showing diahyphal propagules (bundles of conidial chains dispersed as a unit) with adhering or intermixed Heveochlorella photobionts (arrows). E, campylidial macroconidia, with co-dispersed Heveochlorella photobionts loosely encircled, germinating (arrowheads) on a plastic cover slip. F, diahyphal propagules of Gyalectidium germinating (arrowheads) on a plastic cover slip, with co-dispersed Heveochlorella photobionts. Scales: A, C & D = 20 μm; B = 50 μm; E & F = 10 μm.

Figure 4

Fig. 4. Dichotomously lobed chloromorphs of Sticta canariensis emerging from lower surfaces of cyanomorph thalli (arrows). Scale = 5 mm.

Figure 5

Fig. 5. Muriform ascospore (a), probably of Calopadia, germinating on a plastic cover slip placed in a south-west Florida oak hammock, and lichenizing a group of algal cells (arrow), most likely Heveochlorella. Scale = 20 μm.

Figure 6

Fig. 6. Phycopeltis free-living and in stages of lichenization. A, free-living. B, edge of developed Phycopeltis thallus (left) lichenized by a network of hyphae (probably foliicolous Porina sp.) that extend over substratum and capture additional young Phycopeltis germlings (arrows). Scales: A = 20 μm; B = 10 μm.