ArticlePDF Available

Abstract and Figures

Alcobiosis, the symbiosis of algae and corticioid fungi, frequently occurs on bark and wood. Algae form a layer in or below fungal basidiomata reminiscent of the photobiont layer in lichens. Identities of algal and fungal partners were confirmed by DNA barcoding. Algal activity was examined using gas exchange and chlorophyll fluorescence techniques. Carbon transfer from algae to fungi was detected as ¹³C, assimilated by algae, transferred to the fungal polyol. Nine fungal partners scattered across Agaricomycetes are associated with three algae from Trebouxiophycae: Coccomyxa sp. with seven fungal species on damp wood, Desmococcus olivaceus and Tritostichococcus coniocybes, both with a single species on bark and rain-sheltered wood, respectively. The fungal partner does not cause any obvious harm to the algae. Algae enclosed in fungal tissue exhibited a substantial CO2 uptake, but carbon transfer to fungal tissues was only detected in the Lyomyces-Desmococcus alcobiosis where some algal cells are tightly enclosed by hyphae in goniocyst-like structures. Unlike lichen mycobionts, fungi in alcobioses are not nutritionally dependent on the algal partner as all of them can live without algae. We consider alcobioses to be symbioses in various stages of co-evolution, but still quite different from true lichens.
This content is subject to copyright. Terms and conditions apply.
1
Vol.:(0123456789)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports
Alcobiosis, an algal‑fungal
association on the threshold
of lichenisation
Jan Vondrák 1,2, Stanislav Svoboda
1,2, Lucie Zíbarová
3, Lenka Štenclová 2,4, Jan Mareš 2,4,
Václav Pouska
5, Jiří Košnar 1,2 & Jiří Kubásek 6*
Alcobiosis, the symbiosis of algae and corticioid fungi, frequently occurs on bark and wood. Algae
form a layer in or below fungal basidiomata reminiscent of the photobiont layer in lichens. Identities
of algal and fungal partners were conrmed by DNA barcoding. Algal activity was examined using gas
exchange and chlorophyll uorescence techniques. Carbon transfer from algae to fungi was detected
as 13C, assimilated by algae, transferred to the fungal polyol. Nine fungal partners scattered across
Agaricomycetes are associated with three algae from Trebouxiophycae: Coccomyxa sp. with seven
fungal species on damp wood, Desmococcus olivaceus and Tritostichococcus coniocybes, both with a
single species on bark and rain‑sheltered wood, respectively. The fungal partner does not cause any
obvious harm to the algae. Algae enclosed in fungal tissue exhibited a substantial CO2 uptake, but
carbon transfer to fungal tissues was only detected in the Lyomyces-Desmococcus alcobiosis where
some algal cells are tightly enclosed by hyphae in goniocyst‑like structures. Unlike lichen mycobionts,
fungi in alcobioses are not nutritionally dependent on the algal partner as all of them can live without
algae. We consider alcobioses to be symbioses in various stages of co‑evolution, but still quite
dierent from true lichens.
Premises
(1) Denition of symbiosis. Symbiosis is a commonly used term in biology, but traditionally has two distinct
meanings1. In what follows, we use it in the sense of de Bary2 to refer to the close and long-term coexistence
of two dierent organisms. Symbiosis in this sense need not involve any benecial or harmful relationship,
merely close co-existence.
(2) Denition of lichen. Lücking etal.3 provided a chronologically ordered list of lichen denitions. None of
these is entirely satisfactory for our purposes, so here we dene a lichen as follows. A lichen is an association
of a fungus (mycobiont) and an alga or cyanobacterium (photobiont) with the following characteristics: (i)
e mycobiont is nutritionally dependent on its photobiont. (ii) e mycobiont is not obviously harmful to
its photobiont. (iii) e photobiont occurs within the mycobiont thallus. (iv) Mycobionts and photobionts
usually cannot persist over a long period outside the symbiosis.
Mutualistic relationship between photoautotrophs and fungi arose many times in evolution, had paramount
importance in the development of terrestrial life and still remains essential in the present-day ecosystems4. A
agship of these relationships is lichen symbiosis, a highly elaborated cooperation between fungi and green
algae and/or cyanobacteria where the fungal partner is nutritionally dependent on its photoautotroph5. Lichen
symbiosis has multiple independent origins6 and its complexity and the stage of lichenisation diers considerably
OPEN
1Institute of Botany of the Czech Academy of Sciences, Zámek 1, CZ252 43 Průhonice, Czech
Republic. 2Department of Botany, Faculty of Science, University of South Bohemia CZ, 370 05 České Budějovice,
Czech Republic. 3Independent Researcher, Resslova 26, 400 01 Usti Nad Labem, Czech Republic. 4Biology Centre
of the Czech Academy of Sciences, Institute of Hydrobiology, 370 05 České Budějovice, Czech Republic. 5Faculty of
Forestry and Wood Science, Czech University of Life Sciences Prague, Kamýcká 129, 165 00 Praha-Suchdol, Czech
Republic. 6Department of Experimental Plant Biology, Faculty of Science, University of South Bohemia, CZ370
05 České Budějovice, Czech Republic. *email: jirkak79@gmail.com
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
Vol:.(1234567890)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
in various examples7. In some cases, a single fungal species may be either lichenised or saprophytic depending
on conditions8,9, and the same is truth for algae10.
Numerous fungi are apparently associated with, and nutritionally dependent on algae, but their thallus is
inconspicuous, not stratied into the typical lichen thallus6. A list of such fungi, so called “semilichens, has been
recently provided by Vondrák etal.11. Apart from semilichens, other algal-fungal symbioses that do not fully
meet the denition of a lichen exist12. A remarkable but overlooked one, is linked to corticioid fungi, tradition-
ally dened as basidiomycetes with fruiting bodies (basidiomata) appressed to the substrate with supercial
non-poroid hymenium (extended denition in13). ese at basidiomata (“crusts” in further text) usually cover
wood or bark.
Albertini & Schweinitz14 described one of the wood-dwelling corticioid fungi as Hydnum bicolor, currently
named Resinicium bicolor. e epithet bicolor reected the contrast between the white surface of fruiting bodies
and the rusty brown tips of hymenial spines (the coloration frequently observed in older basidiomata). is epi-
thet additionally matches an even more distinctive colour contrast—the white (or partly translucent) fungal crust
is regularly green beneath. e green tinge is caused by an algal layer formed directly below the white fungal coat.
A detailed investigation of various corticioid fungi in European woodlands revealed living algal cells thriving
below or inside the crusts of several unrelated fungal species. We propose to name these alliances “alcobioses”
(singular “alcobiosis”), such as alga and corticioid fungus in symbiosis. Whereas some alcobioses form unstable
associations where the algal cells are few in scattered colonies, others apparently have a tight relationship where
algae form a lichen-like algal layer15. Only a few algal taxa have previously been reported in an association with
corticioid fungi, mainly unicellular members of the green-algal class Trebouxiophyceae such as Coccomyxa
glaronensis15 and undetermined species of Coccomyxa and Elliptochloris16,17. ese algae are ubiquitous in ter-
restrial habitats and exhibit a general tendency to enter lichen-like symbioses18. However, their symbiotic and
free-living members are morphologically uniform, and separation of the cryptic lineages requires molecular
data19,20 not available for alcobioses yet.
e similarity of alcobioses and crustose lichens is remarkable, but the former have received little and only
supercial attention. e question of whether alcobioses have a nutritional character, as in lichens21, has not been
addressed. Here we provide the most comprehensive morphological and taxonomical assessment so far concern-
ing both symbiotic partners. We also demonstrate that algal cells in these consortia are alive and metabolically
and photosynthetically active even when fully embedded in the fungal crust. And we also studied carbon transfer
from algal polyols (ribitol and sorbitol in our cases) into fungal mannitol that would conrm the nutritional
relationship of the fungus to the algae.
Results
Alcobioses are stratied systems with an internal algal layer. Algal cells were observed enclosed
either in the lower part of crustose basidiomata (subiculum) or in the substratum below the crusts, mostly rot-
ten wood, however they never covered the crust surface. e density of algal cells varied considerably among
infraspecic individuals and among species from an entire absence to a distinct thick continuous layer. e
thickness of the algal layer varied, but frequently exceeded 100µm. Whereas the algal layer was formed only
occasionally in some species (Exidiopsis calcea and Tubulicrinis subulatus), it was found regularly in Lyomyces
sambuci (Fig.1), Resinicium bicolor (Fig.2), Skvortzovia furfuracea (Fig.3) and some Xylodon spp. All these
fungi, however, were also recorded living separately, without an internal algal layer. In all cases, the colonies of
algal cells were enclosed in fungal tissue (although they are also found in thesubstrate below the crust Fig.3E).
Nevertheless, a truly close contact where algae are encircled by fungal hyphae was mostly not observed. e only
exception was Lyomyces sambuci which occasionally formed spherical goniocyst-like structures (20–40µm in
diameter) where a group of algal cells was tightly enclosed within the hyphal network (Fig.1F).
Diverse fungal partners are associated with several ecologically distinct algae. e fungal part-
ners appear to be more diverse in alcobioses than the associated algae. We found nine fungal species dispersed
across the phylogeny of Agaricomycetes (TableS1, Fig.S3) and only three algal partners from the class Trebouxi-
ophycae (TableS1, Fig.4, FiguresS4, S5, S6). e vast majority of fungi involved (i.e. seven species) entered
the symbiosis with a single Coccomyxa species (FiguresS4, S6A,B,C,D). All these fungi have similar ecology,
occurring in temperate forests on decaying wood, especially on fallen spruce trunks, in shaded conditions (Fig-
ures2 and 3; TableS1). e associated Coccomyxa sp. has a rbcL sequence almost identical (> 99% identity) to a
lichen photobiont of Sticta22 and to a symbiont in an allegedly lichenised Schizoxylon albescens9. Both symbioses
have quite distinct ecology: the former is a tropical lichen from Cuba and the latter is known from aspen bark
in Northern Europe. In addition, the same algal genotype labelled Coccomyxa sp. “cort06” in terms of the rbcL
sequence (~99.8% identity) was found to colonize the bark of trees with no report of association with fungi23.
e single fungus, Lyomyces sambuci, was regularly associated with another green algal species Desmococcus
olivaceus, as determined both by morphological identity (Fig.S6E,F,G,H) and high rbcL sequence similarity
(~99.5%; Fig.S5) to a typical strain of this species—SAG 1.9424. is alcobiosis diers in ecology from the
associations with Coccomyxa. It is more tolerant of drying, occurs in lighter sites and has a strong anity to
bark and wood of living or dying Sambucus shrubs. e alliance Lyomyces-Desmococcus is apparently the most
intimate among the observed alcobioses as it forms goniocyst-like structures and the algal carbon is undoubtedly
transferred to the fungus (see below). However, again, the same species of Desmococcus also occurs free-living
(Fig.1A,24) and its symbiosis with Lyomyces is clearly facultative.
Finally, a green alga matching Tritostichococcus coniocybes, both in morphology (Fig.S6H,I) and the rbcL
sequence (97–99% identity; Fig.S5), was detected in alcobiosis with a single fungus, Kneiella abieticola. It was
observed on so rotten wood of spruce snag in a microsite sheltered from rain. Tritostichococcus coniocybes is a
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
Vol.:(0123456789)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
lichen photobiont detected in Chaenothecopsis spp.24, Arthonia thoriana and Chaenotheca spp. (Fig.4), and also
was observed free-living (our data). It always occurred in rain-sheltered microhabitats.
Algae thrive in the association. Absolute uorescence intensity (e.g. Fo) is very variable and demon-
strates heterogeneity in algal chlorophyll abundance (Fig.5A,D,G). e maximal quantum yield of photosystem
II (Fv/Fm) is very homogeneous (0.55 to 0.75) for well hydrated alcobioses studied (Fig.5, Fig.S7). Moreover, Fv/
Fm for alcobioses and uncovered algal layer does not dier substantially (Fig.5C,F,I). In contrast, dry systems
have Fv/Fm close to zero and recover quickly aer rehydration (FiguresS7, S8). Minimal uorescence (Fo) was
increased to some extent when fungal crusts were removed by razor blade from algal layer. It may demonstrate a
shielding eect of the fungal partner on algae, particularly in Lyomyces-Desmococcus system (Fig.5J,K,L,M,N,O).
CO2 exchange of alcobioses was also very variable but easily detectable, and it conrms the viability and
physiological activity of the partners. Typical light response curves of CO2 assimilation for ve alcobioses,
Figure1. Association of Lyomyces sambuci and Desmococcus olivaceus (GPS: 48.9409975N, 14.5175219E;
voucher: PRA-JV25262). (A) bark of Sambucus nigra covered by a free-living Desmococcus algal crust which is
largely overgrown by Lyomyces; (B) vertical section of the Lyomyces crust with a distinct algal layer; (C) vertical
section with the red chlorophyll autouorescence; (D) algal colonies incorporated in a loose hyphal tissue, below
the cover of compact fungal tissue; (E) Desmococcus in the algal layer; (F) Desmococcus and Lyomyces form
lichen-like goniocysts. Scales: (B, C), 100µm; (D, E, F), 20µm.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
Vol:.(1234567890)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
free-living terrestrial alga and foliose lichen are in Fig.S9. e following features were common for all systems:
(1) Respiration of just rehydrated alcobioses was much higher (typically ve to tenfold) than of those hydrated for
a long time, equilibrating to steady-state in hours to days. (2) All alcobioses respond to light, and photosynthesis
exceeded overall respiration in some cases (particularly under elevated [CO2]). (3) Photosynthesis is saturated
under unusually dim light (typically 50 to 100µmol photons m−2 s−1) in systems with Coccomyxa algae, but
may be still increasing in the others, under full sun intensity (ca 1800µmol m−2 s−1) and elevated CO2, Fig.S5).
Moreover, lower temperatures are benecial for the net carbon gain (Fig.S10).
Algal carbon is absorbed by fungi in some alcobioses. Pilot (HPLC–MS, GC–MS) experiments did
not detect any signicant carbon transfer from algal polyols (sorbitol and ribitol) to fungal substances (mannitol
and ergosterol) in Resinicium bicolor-Coccomyxa, Skvortzovia furfuracea-Coccomyxa and Xylodon asper-Cocco-
myxa. In contrast, the control lichens, Hypogymnia physodes and Multiclavula mucida, expressed a clear pattern
of 13C transfer from algal ribitol to fungal mannitol aer two hours of assimilation in the 13C-enriched air. Inter-
mediate results were obtained for Lyomyces sambuci-Desmococcus where mannitol expresses some degree of 13C
enrichment, but barely signicant in our set-up (data not shown).
Figure2. Association of Resinicium bicolor and Coccomyxa (GPS: 48.6679583N, 14.7053083E; voucher:
PRA-JV25257). (A) typical habitat–vertical surface of rotten spruce trunks; (B) R. bicolor crust; (C) vertical
section with a distinct algal layer below the fungal coat; (D) the red chlorophyll autouorescence indicates
locations of Coccomyxa cells in the vertical section. Scales: (B) 5mm; (C, D) 50µm.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
Vol.:(0123456789)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
Subsequent detailed studies using isotope ratio mass spectrometry (IRMS), much more sensitive for isotope
abundances, were performed (Method S1). ese experiments delivered two strikingly dierent results: (1) e
Skvortzovia furfuracea-Coccomyxa system did not display any algal-fungal carbon transport. e principal polyols
were algal ribitol and fungal mannitol (Fig.6A). TMS-ribitol was highly 13C enriched (3.90 ± 0.26 At%, t = 24.4,
P < 0.001, N = 5) aer 18h of labelling whereas the fungal TMS-mannitol remained very close to the natural 13C
abundance (1.07 ± 0.03 At%, t = − 0.8, P = 0.45, N = 5; Fig.6D). (2) e algal-fungal carbon transfer was conrmed
in the Lyomyces sambuci-Desmococcus alliance, where sorbitol is the principal algal polyol (Fig.6C). We recorded
a substantial 13C signal in TMS-sorbitol (2.15 ± 0.07 At%, t = 28.8, P = 0.001; N = 4) and a slightly lower but still
very signicant signal in TMS-mannitol (1.48 ± 0.10 At%, t = 8.2, P = 0.004, N = 4; Fig.6F).
Two specimens of Botryobasidium-Coccomyxa, measured in parallel, having low biomass and, thus, low
polyol abundances (Fig.6B), delivered very convincing negative result. Its TMS-ribitol 13C enrichment was
high and consistent (3.70 ± 0.14 At%, t = 25.9, P = 0.024; N = 2) but mannitol invariant from natural abundance
(1.11 ± 0.04 At%, t = 1.4, P = 0.39, N = 2; Fig.6E), very similar to Skvortzovia furfuracea-Coccomyxa system. See
also chromatograms (FiguresS11, S12).
Snails rejuvenate alcobioses and produce isidia‑like diaspores. Literature says little about the per-
sistence of the corticioid fungal crusts involved in this study. According to our eld observations, crusts of
Lyomyces sambuci, Resinicium bicolor and Skvortzovia furfuracea can persist over several years. Persistence of
alcobioses is apparently assisted by snail grazing. Large areas of fungal crusts including algae are frequently
removed by snails and the algal layer is uncovered (Fig.7A). Grazed spots are quickly overgrown by rejuvenated
fungal hyphae (Fig.7B). erefore, fungal crusts, especially of Skvortzovia furfuracea, typically form a continu-
ous mosaic of younger and older (i.e. thinner and thicker) patches.
Snail excrements form distinct structures accompanying alcobioses. ey have a granular or caterpillar-like
shape (Fig.7C) and, when lying on grazed fungus, they are quickly overgrown by young hyphae and incorpo-
rated into the fungal crust (Fig.7D). Fresh excrements are green, densely lled by living algal cells (Fig.7E) in
a mixture with remnants of fungal hyphae (Fig.7F). e algal-fungal content makes excrements an analogous
structure to vegetative propagules (e.g. isidia) of lichens and may serve for vegetative reproduction when being
Figure3. Association of Skvortzovia furfuracea and Coccomyxa (GPS: 48.6679583N, 14.7053083E; voucher:
PRA-JV25255). (A) typical habitat – shaded surface of rotten spruce trunks; (B) S. furfuracea crust grazed by
snails; (C, E) vertical sections of S. furfuracea crust. Fungal tissues coloured by lactoglycerol cotton blue. A
distinct algal layer is visible below a dark blue fungal coat; (D) the red chlorophyll autouorescence indicates
locations of Coccomyxa cells in the vertical section; (F) Coccomyxa loosely integrated in the fungal tissue. Scales:
(B) 1cm; (C, D) 50µm; (E) 20µm; (F) 10µm.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
Vol:.(1234567890)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
placed/translocated outside the current fungal crust. e same function was described at mite excrements that
contained viable fungal and algal cells of the lichen Xanthoria parietina25.
Discussion
Corticioid fungi associated with algae. Corticioid fungi are generally considered saprophytic, rarely
parasitic and mycorrhizal; the frequent association with algae has largely been ignored in most monographs26.
However, Parmasto27, already in 1967, observed an alcobiosis in the newly described fungus Phlebia lichenoides,
as he refers to an internal algal layer containing Chlorococcus-like cells. Hjortstam etal.28 considered Phlebia
lichenoides as a synonym of P. subcretacea (currently Cabalodontia subcretacea) and dismissed its association
with algae.
Poelt & Jülich15 provided the most comprehensive study so far on alcobiosis revealed in Resinicium bicolor,
and refer to the presence of an algal layer formed, allegedly, of Coccomyxa glaronensis. (is algal species was
described as a symbiont in the lichen Solorina saccata29). Poelt & Jülich15 observed algae in all surveyed specimens
Figure4. Linkage between algal and fungal partners in alcobioses. Algae are arranged in rbcL phylogenetic
trees of Trebouxiophyceae; only parts with Stichococcus s.lat., Desmococcus and Coccomyxa depicted. Fungi are
arranged in the ITS tree of selected Agaricomycetes. Symbionts in alcobioses are in grey rectangles. Detailed
trees are available on FiguresS3, S4, S5.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
Vol.:(0123456789)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
and algal colonies were either restricted below fungal crusts or only slightly expanded out of the fungal spots
and formed a surrounding green rim (which corresponds with our observations; Fig.5A,B). e authors classi-
ed the algal-fungal contact as intimate, but without obvious appressoria or haustoria. Subsequent short notes
in literature13,3032 have only repeated the ndings by Poelt & Jülich. Whereas Oberwinkler speculated in 1970
that R. bicolor represents a basidiolichen30, he omitted this fungus from his later overview of basidiolichens33.
Two related recent conference contributions refer to associations of corticioid Hyphodontia s.lat. (namely
Lyomyces crustosus, Hyphodontia pallidula and Xylodon brevisetus) with algae Coccomyxa and Elliptochloris, but
Figure5. Chlorophyll uorescence imaging of fungal-algal associations. Minimal uorescence (Fo), visual
frame and maximal quantum yield of photosyntem II (Fv/Fm) for Resinicium bicolor-Coccomyxa (A, B, C, J, K),
Skvortzovia furfuracea-Coccomyxa (D, E, F, L, M) and Lyomyces sambuci-Desmococcus (G, H, I, N, O). Each
sample contains both algal patches and alcobiosis where fungal crust completely covers the algae. Bottom pairs
of frames demonstrate shielding eect of fungal crusts. Upper frames are intact alcobioses (J, L, N), whereas
fungal part of the system was carefully removed by razor blade in red frames of the bottoms (K, M, O).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
8
Vol:.(1234567890)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
without molecular sequence data16,17. It is not clear from the published meeting abstracts, if both algal genera
were observed co-occurring in an association with a single fungal crust, or if each algal-fungal system involved
only a single algal partner. We have only observed the latter case. e anatomical observations reported by
Voytsekhovich etal.17 allegedly revealed appressoria and haustoria in hyphae attached to algal cells and, on
this basis, the authors consider these fungi to be optionally lichenised. Physiological relationships between the
corticioid fungi and their algae were not studied. Our data conrmed that some species of Hyphodontia s.lat.
are frequently involved in alcobioses, but we do not consider them lichens (see below). Whereas Coccomyxa is
undoubtedly frequent in alcobioses, we did not detect Elliptochloris in specimens collected in the current study.
Associations of corticioid fungi and algae sometimes have a typically parasitic character. It was demonstrated
for Athelia epiphyla15 where fungal hyphae form haustoria penetrating algal cells. is fungus causes bleaching
of corticolous algae, i.e. forms characteristic pale-grey rounded spots on otherwise green tree bark. Although A.
epiphyla is apparently a parasite, Jülich34 and Oberwinkler33 refer to symbiotic relationships between epiphytic
algae and some species of Athelia and Athelopsis. Some of these cases may be close to our concept of alcobiosis.
Parallel research has been conducted on associations of polypore fungi and their epiphytic algae. e surface
of polypore fruiting bodies is a suitable substrate for numerous algal (and cyanobacterial) species35,36. However,
these associations do not show any specicity–a single polypore is oen covered by a community of several
Figure6. Abundance and 13C enrichment in trimethylsilyls of principal algal and fungal polyols aer 18h
assimilation of alcobioses in 13CO2 atmosphere. Algal polyols: ribitol (R) and sorbitol (S); fungal mannitol
(M). Median (central point), 25% and 75% quantiles (box) and min–max (whiskers) are shown. Dotted line
represents natural 13Cabundance (1.07 At %) and t-test was performed against it; NS: not signicant (P > 0.05),
*: 0.05 < P > 0.01, **: 0.01 < P > 0.001, ***: P < 0.001.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
9
Vol.:(0123456789)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
algal species with broader niches (not restricted to polypores). Carbon transfer from epiphytic algae to polypore
fungal tissues was repeatedly reported using 14C tracer37,38, but we are skeptical of that result as the algae usually
grow on/in dead polypore tissues. us, this association seems to be very dierent from alcobioses with algae
embedded within the viable fungal tissue.
e taxonomic composition of algae found in alcobioses is in general not surprising. Unicellular trebouxi-
ophyte algae such as Coccomyxa and Stichococcus sensu lato are ubiquitous in terrestrial habitats and have been
long known to create various types of fungal-algal consortia39. In our study we employed molecular barcoding of
the algal partners in alcobioses for the rst time, which allowed us to place all three detected taxa into narrowly
dened and strongly supported monophyletic clades of Coccomyxa sp., Desmococcus olivaceus, and Tritostichoc-
occus coniocybes. Each of these clades probably corresponds to a single algal species, judging from their uniform
morphology and high rbcL gene sequence identity (Fig.4, FiguresS3, S4, S5). e Coccomyxa sp. found in our
samples evidently represents a species new to science, but its transfer to pure culture, deposition of a holotype,
and formal description was beyond the scope of the current study. As well as living in alcobioses, we know from
previously published sequences that each of these three genospecies can live as a free-living terrestrial alga23,24.
In the case of D. olivaceus, our results are the rst reliable observation of this species in symbiotic association
with fungi, but S. coniocybes and the particular clade of Coccomyxa have previously been reported as genuine
lichen photobionts22,24.
Ecophysiological implications. We conrmed the viability of algae in alcobioses via chlorophyll a
uorescence40 and gas exchange measurement. e maximal quantum yield of photosystem II (Fv/Fm) is a well-
accepted measure of vitality and impact of stress. In the case of alcobioses, Fv/Fm conrms that the algae are not
stressed under the fungal layer, having Fv/Fm comparable to values for adjacent free-living algae (Fig.5C,F,I).
Photochemistry recovered within minutes to tens of minutes to values > 0.4 aer dry alcobioses were rehydrated
(FiguresS7, S8), being considered “physiologically active”, in accordance with studies on terrestrial algae41.
Figure7. Snail-grazed Skvortzovia furfuracea. (A) a green area of exposed algal layer formed by snail
grazing; (B) regenerated mycelium overgrowing grazed areas. Algal cells are indicated by the red chlorophyll
autouorescence. (C) fresh snail excrement; (D) excrements are quickly overgrown by regenerated mycelium;
(E) the red chlorophyll autouorescence indicates high density of Coccomyxa cells in an excrement; (F) a
mixture of Coccomyxa and remnants of fungal tissue in an excrement. Scales: (A) 2mm; (B) 0µm; (C, D) 1mm;
(E, F) 20µm.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
10
Vol:.(1234567890)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
e minimal uorescence (Fo), in turn, shows a low shielding eect of fungal crusts to incident light, particu-
larly for shade-adapted alcobioses formed by Coccomyxa and fungi with rather thin basidiomata (Fig.5J,K,L,M).
A substantial shielding eect was detected at the thicker and more light-scattering crust of Lyomyces sambuci,
growing in more-light exposed sites, which may be photoprotective for Desmococcus algae in this association
(Fig.5 N,O).
e CO2 exchange, reecting the real photosynthetic performance and production process, has not been
studied in alcobioses yet, but reference data are available for lichens4244. Here we provide gas-exchange data for
four alcobiosis systemes in the context of lichen symbiosis (represented by Parmelia sulcata) and a free-living
algal crust (Trentepohlia aurea). FiguresS9 and S10 demonstrate relationships of the CO2 exchange to an incident
light intensity, gaseous CO2 concentration and ambient temperature. Respiration of the fungal partner plus co-
occurring microbiomes was usually higher in shade-adapted Coccomyxa-based alcobioses (Fig.S9). us, net
carbon balance of the system was mostly negative and algal photosynthesis could not serve as a principal source
of carbon for the fungus. Saturating light intensity was also unusually low (50 to 100µmol photons m−2 s−1)
suggesting a strong shade adaptation of the algae involved. In contrast, free-living Trentepohlia aurea, the lichen
Hypogymnia physodes and the alcobiosis Lyomyces sambuci-Desmococcus occurring in more light-exposed con-
ditions, had higher maximal CO2 assimilation and higher light saturation intensity and the carbon gain of these
systems was frequently positive. CO2 assimilation rate rising even close to full sun intensity under elevated CO2
suggests high diusional limitation of the photobionts and their high photosynthetic capacity as well (Fig.S9).
In addition, the overall carbon balance of those systems is strongly temperature dependent. Lower temperatures
promote carbon gain of alcobioses, as demonstrated on Xylodon system (Fig.S10). is can be explained by higher
temperature dependency of dark respiration (Rdark) than gross photosynthetic capacity (Agross). Comparable
magnitudes of Rdark and Agross will lead to substantial eect of temperature on carbon gain45.
We observed a signicant alga-to-fungus 13C transfer in only one of the systems studied. We selectively meas-
ured algal (ribitol, sorbitol) and fungal (mannitol) compounds. Polyols are not suitable for gas chromatography,
therefore their trimethylsilyls (TMS-) were measured46. at is, the ve-carbon ribitol was converted to the
twenty-carbon TMS-ribitol (C20H52O5Si5). Similarly, six-carbon mannitol and sorbitol led to the formation of
24-carbon TMS-derivatives (C24H62O6Si6). us, the 13C enrichment (indicating carbon transfer) is four-times
more pronounced in mother molecules than in TMS derivatives measured here and thus the negative result for
Skvortzovia furfuracea-Coccomyxa is very convincing (Figs.7and S11). e isotopic precision of IRMS is better
than 0.01 At% of 13C for isotope ratios close to natural (< 5 At% of 13C). Mannitol 13C is clearly unchanged from
its natural value in this system, whereas TMS-ribitol is highly enriched up to 4.6 At% 13C (by 3.5 At% compared
to natural). is means that the parent ribitol should have up to 15.1 At% 13C (4 × 3.5 + 1.1). In contrast, the
Lyomyces-Desmococcus TMS-mannitol was signicantly enriched by approximately 0.4 At% (Figs.7andS12).
us, mannitol should be enriched up to 2.7 At% (1.1 + 4 × 0.4 At%). Despite the clear statistical signicance,
more work is needed to decipher the timing, environmental dependencies and, thus, ecological signicance of
those carbon transfers.
Symbiosis on the threshold of lichenisation. e lichen is dened above in the Premise 2 as a symbio-
sis of alga or cyanobacterium (photobiont) and fungus (mycobiont) with following specics: (1) e mycobiont
is nutritionally dependent on its photobiont47. (2) e mycobiont is not obviously harmful to its photobiont48.
(3) e photobiont occurs within the mycobiont thallus49. (4) Mycobionts and photobionts usually cannot per-
sist over a long period outside the symbiosis50,51.
Alcobioses do comply with points (2) and (3). ey have an internal lichen-like algal layer (Figs.1, 2, 3) and
the algae thrive in the symbiosis (Fig.5, FiguresS3, S4, S5, S6). Only Lyomyces-Desmococcus partly complies
with point (1); we conrmed carbon transfer from alga to fungus (Fig.6, FiguresS11, S12). We suggest however
that Lyomyces is not fully dependent on algal assimilates, because it has been observed occasionally without the
algal symbiont. Consequently, Lyomyces-Desmococcus does not meet point (4), because both symbionts may live
apart. Surprisingly, Desmococcus olivaceus is absent from the list of lichen photobionts39 and is mostly reported
free-living, forming extensive epiphytic or epilithic algal crusts52. Another member of the genus, Desmococcus
vulgaris, was found overgrowing fruiting bodies of the polypore Fomes fomentarius53. Lyomyces sambuci, like
other Hyphodontia s. lat., is believed to cause white rot, i.e. is saprophytic and, remarkably, the frequently present
algal layer in this common fungus was not mentioned or illustrated in the monographs26,28. Point (4) is partly met
in some alcobioses with the Coccomyxa species known from lichens9,22. At our sampling sites, the alga appears
to occur only inside and below the fungal crusts (Fig.5A,B). Contrary to our observations, an almost identical
genotype of Coccomyxa was found free-living on the bark of live pine and oak trees in the study of corticolous
algae by Kulichová etal.23. erefore, the lifestyle of this alga seems to be diverse, and investigations on the popu-
lation level are needed to elucidate the potential specic life strategies of individual strains. Fungi associated with
Coccomyxa were observed to live without algae, but the alcobiosis is almost omnipresent in Resinicium bicolor and
Skvortzovia furfuracea. e nature of this symbiosis remains enigmatic as it probably has no nutritional character.
In conclusion, the photosynthetic potential of algal partners is clearly substantial, but their direct nutritional
importance for fungi (and whole alcobioses) is still obscure, even in Lyomyces-Desmococcus, and needs more
study. Simultaneously, the insignicant carbon exchange observed in most systems implies that there must be
other ecological advantages keeping the partners in a stabile association (e.g. exchange of bioactive substances
under stress conditions such as drought).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
11
Vol.:(0123456789)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
Materials & methods
In the period 2016–2021, we recorded 58 specimens with alcobiosis (TableS1). Corticioid fungi were identi-
ed to species from their morphology and 27 specimens were barcoded by ITS nrDNA sequences. Algae were
determined to genus according to their morphology and 20 specimens were barcoded by sequencing the ribulose
bisphosphate carboxylase large subunit gene (rbcL). e NCBI accession numbers of the sequences obtained
are provided in TableS1. Field collections from 2019–2021 were employed in morphological and anatomical
observations (uorescent microscope Olympus BX 61 in bright eld and uorescent mode) and in physiological
measurements that were done within one week aer collection. Before that, samples were hydrated if necessary,
put in Petri-dishes and accommodated in LED illuminated cultivating room under 20°C, irradiation about
100µmol m−2 s−1 and photoperiod of 12h for at least two days.
DNA barcoding & phylogenetic analysis. Fresh specimens were used for DNA extraction of both fun-
gal and algal partners. DNA was extracted with a cetyltrimethylammonium bromide (CTAB)-based protocol54.
ITS nrDNA locus was found to be a useful fungal barcode sequence, easily ampliable, and moreover with
a sucient number of references in the NCBI database. In the case of algae, we sequenced nrDNA ITS and
18S regions and the rbcL gene. e latter was found as best amplied and therefore selected for barcoding and
phylogenetic analysis. Polymerase chain reactions were performed in a reaction mixture containing master mix
consisting of 2.5mmol/L MgCl2, 0.2mmol/L of each dNTP, 0.3μmol/L of each primer, 0.5 U Taq polymerase
in the manufacturer’s reaction buer (Top-Bio, Praha, Czech Republic), and milli-Q water to make up a nal
volume of 10μL. e primers used for PCR and the cycling conditions are summarized in TableS2. Successful
amplications were sent for Sanger sequencing (GATC Biotech, Konstanz, Germany). Sequences were edited
using BioEdit v.7.0.9.055 and Geneious Prime 2022.0 (https:// www. genei ous. com).
Sequences of our specimens were supplemented by relevant sequences from GenBank—NCBI database.
Sequences were aligned by MAFFT v.756; available online at http:// ma. cbrc. jp/ align ment/ server/) using the
Q-INS-i algorithm and adjusted manually. e best-t model of sequence evolution was selected using the Akaike
information criterion calculated in jModelTest v.0.1.157. Relationships were assessed using Bayesian inference as
implemented in MrBayes v.3.1.258. Two runs starting with a random tree and employing four simultaneous chains
each (one hot, three cold) were executed. e temperature of a hot chain was set empirically to 0.1, and every
100th tree was saved. e analysis was considered to be completed when the average standard deviation of split
frequencies dropped below 0.01. e rst 25% of trees were discarded as the burn-in phase, and the remaining
trees were used for construction of a 50% majority consensus tree.
Gas exchange. CO2 exchange was measured using the portable photosynthetic system LI-6400XT (Li-
Cor, Lincoln, NE, USA) connected to a custom-made peltier-conditioned gas exchange chamber described
elsewhere59,60, see also Figure. S1. e chamber allows to accommodate samples up to 64 cm2 of ground area and
1cm in thickness. It is of high sensitivity, as we have demonstrated in the past by successfully measuring the very
small gas exchange of moss capsules61. Algae-containing basidiomata were attached to the razor-thinned native
substrate to minimise microbial respiration. “White” PAR irradiation was supplied by a LI6400-18 RGB system
source controlled by LI-6400XT. Temperature was maintained at 20°C (except when measuring the response
to varying temperature) and airow at 250µmol mol−1. e light response of CO2 assimilation was measured in
continual-logging mode (each 5s), at least 3min in each light intensity. Aer a steady state was reached, data
were collected, and light intensity changed to the next value. e temperature curve relied on Peltier-cooling of
the cryptogamic chamber. At least 15min was allowed to pass before data on gas exchange was taken, to allow
equilibrium to be reached. Finally, background (empty chamber response) was measured and subtracted.
Chlorophyll a uorescence imaging. 2D uorescence measurements were performed by the FluorCam
FC800 instrument (PSI, Drásov, Czech Republic). Red LEDs (λ = 660nm) were used for both, measuring light
(10, 20 or 33µs ashes, in average < 1µmol m−2 s−1) and actinic light (continuous, photosynthesis driving, about
100µmol m−2 s−1) photon sources. White LEDs (≈ 2500µmol m−2 s−1) delivered saturation pulses of 1s in dura-
tion. e camera has 720 × 560 px resolution, 12-bit data depth and is equipped with a zoom objective imaging
frame down-to ca 10 × 7.5cm (≈ 0.13mm per pixel).
Metabolite transfer measurement. Stable carbon 13C was chosen to trace photobiont assimilated car-
bon. Substrate-attached sporocarps with active algal partner (Li-6400XT-proved, Figure. S1) were enabled to
assimilate 13CO2 enriched air in a water-sealed 3L inverse Petri dish with internal fan (See Figure. S2). e label-
ling device is described in detail elsewhere61. [13CO2] was about 1000µmol mol−1. Atmospheric CO2 was previ-
ously replaced by CO2 free synthetic air and then ca 3mL of 13CO2 (> 99 atom %, Sigma-Aldrich, Luis., USA) was
injected. Samples assimilated for two to 18h under about 200µmol m−2 s−1 of white LED light. en fungal/algal
crusts were scratched down by razor and killed in boiling methanol. Homogenised and ltered metOH extracts
were used for analysis of polyols and ergosterol (see Methods S1 for further details).
We employed three approaches to trace 13C enriched metabolites. (1) HPLC–MS to separate algal polyols
(ribitol and sorbitol) from fungal mannitol and to measure their 13C content. Although mannitol was detected
in some algal groups62,63, it is not produced by algae in our systems, i.e., Coccomyxa and Desmococcus64. (2) GC-
IRMS to separate trimethylsilyl derivatives of those polyols (TMS-ribitol, TMS-mannitol and TMS-sorbitol) and,
again, to quantify their 13C enrichment. And (3) HPLC–MS to separate fungal specic ergosterol and to measure
if it is 13C enriched. For more details see Method S1.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
12
Vol:.(1234567890)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
Data availability
e datasets generated during and/or analysed during the current study are available from the corresponding
author on reasonable request.
Received: 7 May 2022; Accepted: 3 February 2023
References
1. Wilkinson, D. At cross purposes. Nature 412, 485 (2001).
2. de Bary, H. A. Über Symbiose [On Symbiosis]. Tageblatt für die Versammlung Dtsch. Naturforscher und Aerzte (in Cassel) [Daily
J. Conf. Ger. Sci. Phys.] (in Ger. 51, 121–126 (1878).
3. Lücking, R., Leavitt, S. D. & Hawksworth, D. L. Species in lichen-forming fungi: balancing between conceptual and practical
considerations, and between phenotype and phylogenomics. Fungal Div.109, 99–154 (Springer, Netherlands, 2021).
4. de Vries, J. & Archibald, J. M. Plant evolution: Landmarks on the path to terrestrial life. New Phytol. 217, 1428–1434 (2018).
5. Ahmadjian, V. e Lichen Symbiosis (John Wiley & Sons, 1993).
6. Lücking, R., Hodkinson, B. P. & Leavitt, S. D. e 2016 classication of lichenized fungi in the Ascomycota and Basidiomycota-
approaching one thousand genera. Bryologist 119, 361–416 (2016).
7. Schneider, K., Resl, P. & Spribille, T. Escape from the cryptic species trap: lichen evolution on both sides of a cyanobacterial acquisi-
tion event. Mol. Ecol. 25, 3453–3468 (2016).
8. Wedin, M., Döring, H. & Gilenstam, G. Saprotrophy and lichenization as options for the same fungal species on dierent substrata:
Environmental plasticity and fungal lifestyles in the Stictis-Conotrema complex. New Phytol. 164, 459–465 (2004).
9. Muggia, L., Baloch, E., Stabentheiner, E., Grube, M. & Wedin, M. Photobiont association and genetic diversity of the optionally
lichenized fungus Schizoxylon albescens. FEMS Microbiol. Ecol. 75, 255–272 (2011).
10. Sanders, W. B., Moe, R. L. & Ascaso, C. 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). Eur. J. Phycol.
40, 353–361 (2005).
11. Vondrák, J. et al. From Cinderella to Princess. Preslia 94, 143–181 (2022).
12. Hawksworth, D. L. e variety of fungal-algal symbioses, their evolutionary signicance, and the nature of lichens. Bot. J. Linn.
Soc. 96, 3–20 (1988).
13. L arsson, K. H. & Ryvarden, L. Corticioid fungi of Europe 1. Acanthobasidium–Gyrodontium. Synop. Fungorum 43, 1–266 (2021).
14. Albertini, J. B., von Schweinitz, L. D. Conspectus fungorum in Lusatiae Superioris agro Niskiensi crescentium, e methodo Persooniana.
(DE: Sumtibus Kummerianis, Lipsiae 1805) https:// doi. org/ 10. 5962/ bhl. title. 3601.
15. Poelt, J. & Jülich, W. Über die Beziehungen zweier corticioider Basidiomyceten zu Algen. Österr. Bot. Zeitschri 116, 400–410
(1969).
16. Voytsekhovich, A., Ordynets, O. & Akimov, Y. Optionally lichenized fungi of Hyphodontia (Agaricomycetes, Schizoporaceae) and
their photobiont composition. Aктyaльнi Пpoблeми Бoтaнiки Ta Eкoлoгiї. Maтepiaли Miжнapoднoї Кoнфepeнцiї Moлoдиx
Учeниx 65 (2013).
17. Voytsekhovich, A., Mikhailyuk, T., Akimov, Y., Ordynets, A., Gustavs, L. Optionally lichenized fungi of Hyphodontia (Agaricomy-
cetes, Schizoporaceae). 8th Congress of the International Symbiosis Society, Lisbon, 12–18 July 2015. Lisbon, PT:, 217 (Conf. abstract)
(2015).
18. Gustavs L, Schiefelbein U, Darienko T, P. T. Symbioses of the green algal genera Coccomyxa and Elliptochloris (Trebouxiophyceae,
Chlorophyta). in Algal and Cyanobacteria Symbioses (ed. Grube M, Seckbach J) 169–208 (2017).
19. Darienko, T., Gustavs, L., Eggert, A., Wolf, W. & Pröschold, T. Evaluating the species boundaries of green microalgae (Coccomyxa,
Trebouxiophyceae, Chlorophyta) using integrative taxonomy and DNA barcoding with further implications for the species iden-
tication in environmental samples. PLoS ONE 10, 1–31 (2015).
20. Malavasi, V. et al. 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 (2016).
21. Green, T. G. A., Nash, T. H. Lichen Biology. In Lichen Biology, Second Edition 152–181 (Cambridge University Press, Cambridge,
2008) https:// doi. org/ 10. 1017/ CBO97 80511 790478.
22. Lindgren, H. et al. Cophylogenetic patterns in algal symbionts correlate with repeated symbiont switches during diversication
and geographic expansion of lichen-forming fungi in the genus Sticta (Ascomycota, Peltigeraceae). Mol. Phylogenet. Evol. 150,
106860 (2020).
23. Kulichová, J., Škaloud, P. & Neustupa, J. Molecular diversity of green corticolous microalgae from two sub-mediterranean European
localities. Eur. J. Phycol. 49, 345–355 (2014).
24. Pröschold, T. & Darienko, T. e green puzzle Stichococcus (Trebouxiophyceae, Chlorophyta): New generic and species concept
among this widely distributed genus. Phytotaxa 441, 113–142 (2020).
25. Meier, F. A., Scherrer, S. & Honegger, R. Faecal pellets of lichenivorous mites contain viable cells of the lichen-forming ascomycete
Xanthoria parietina and its green algal photobiont. Trebouxia arboricola. Biol. J. Linn. Soc. 76, 259–268 (2002).
26. Bernicchia, A. & Gorjón, S. P. Corticiaceae s.l. 1008 (2010), ISBN: 9788890105791.
27. Parmasto, E. Descriptiones taxorum novorum. Combinationes novae. Proc. Acad. Sci. Est. SSR. Biol. 16, 377–394 (1967).
28. Hjortstam, K., Larsson, K., Ryvarden, L. & Eriksson, J. e Corticiaceae of North Europe. (Oslo: Fungiora, 1988).
29. Jaag, O. Coccomyxa schmidle Monographie einer algengattung. Beitr. Kryptogamenora Schweiz 8, 1–132 (1933).
30. Ober winkler, F. Die gattungen der Basidiolichenen. Vorträge aus dem Gesamtgebiet der Botanik. Herausgegeb. v. d. Deutsch. bot.
Ges. Neue Folge 4, 139–169 (1970).
31. Poelt, J. Basidienechten, eine in den Alpen lange übersehene Panzengruppe. Jahrb. Vereins Schutze Alpenp. Tiere 40, 81–92
(1975).
32. Eriksson, J., Hjortstam, K. e Corticiaceae of North Europe. Vol. 6. (Grønlands Eskefabrikk, 1981).
33. Oberwinkler, F. Basidiolichens. In Fungal Association 211–225 (Springer, Berlin Heidelberg, Berlin, 2001). https:// doi. org/ 10. 1007/
978-3- 662- 07334-6_ 12.
34. Jülich, W. A new lichenized Athelia from Florida. Persoonia 10, 149–151 (1978).
35. Zavada, M. S. & Simoes, P. e possible demi-lichenization of the basidiocarps of Trametes Versicolor (L.:Fries) pilat (polyporaceae).
Northeast. Nat. 8, 101–112 (2001).
36. Neustroeva, N., Mukhin, V., Novakovskaya, I. & Patova, E. Biodiversity of symbiotic algae of wood decay Basidimycetes in the
Central Urals. III Russ. Natl. Conf. “Information Technol. Biodivers. Res. 1, 83–92 (2020).
37. Zavada, M. S., DiMichele, L. & Toth, C. R. e possible demi-lichenization of Trametes versicolor (L.: Fries) Pilát (Polyporaceae):
e transfer of xed 14CO2 from epiphytic algae to T. versicolor. Northeast. Nat. 11, 33–40 (2004).
38. Mukhin, V. A., Patova, E. N., Kiseleva, I. S., Neustroeva, N. V. & Novakovskaya, I. V. Mycetobiont symbiotic algae of wood-
decomposing fungi. Russ. J. Ecol. 47, 133–137 (2016).
39. Sanders, W. B. & Masumoto, H. Lichen algae: e photosynthetic partners in lichen symbioses. Lichenologist 53, 347–393 (2021).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
13
Vol.:(0123456789)
Scientic Reports | (2023) 13:2957 | https://doi.org/10.1038/s41598-023-29384-4
www.nature.com/scientificreports/
40. Krause, G. & Weis, E. Chlorophyll uorescence and photosynthesis: the basics. Annu. Rev. Plant Biol. 42(1), 313–349 (1991).
41. Lüttge, U. & Büdel, B. Resurrection kinetics of photosynthesis in desiccation-tolerant terrestrial green algae (Chlorophyta) on tree
bark. Plant Biol. 12, 437–444 (2010).
42. Lange, O. L. Moisture content and CO2 exchange of lichens: I. Inuence of temperature on moisture-dependent net photosynthesis
and dark respiration in Ramalina maciformis. Oecologia 45, 82–87 (1980).
43. Palmqvist, K. & Sundberg, B. Light use eciency of dry matter gain in ve macrolichens: Relative impact of microclimate condi-
tions and species-specic traits. Plant Cell Environ. 23, 1–14 (2000).
44. Vondrak, J. & Kubásek, J. Algal stacks and fungal stacks as adaptations to high light in lichens. Lichenol. 45(1), 115 (2013).
45. Smith, N. G. & Dukes, J. S. Plant respiration and photosynthesis in global-scale models: Incorporating acclimation to temperature
and CO2. Glob. Chang. Biol. 19, 45–63 (2013).
46. Medeiros, P. M. & Simoneit, B. R. T. Analysis of sugars in environmental samples by gas chromatography-mass spectrometry. J.
Chromatogr. A 1141, 271–278 (2007).
47. Honegger, R. Functional aspects of the lichen symbiosis. Annu. Rev. Plant Physiol. Plant Mol. Biol. 42, 553–578 (1991).
48. Honegger, R. e lichen symbiosis—What is so spectacular about it?. Lichenologist 30, 193–212 (1998).
49. Kirk, P. M. et al. (eds) Dictionary of the Fungi 10th edn. (CABI, Netherlands, 2008).
50. Ahmadjian, V. e lichen alga Trebouxia: Does it occur free-living?. Plant Syst. Evol. 158, 243–247 (1988).
51. Sanders, W. B. 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.
Am. J. Bot. 101, 1836–1848 (2014).
52. Rindi, F. & Guiry, M. Composition and spatial variability of terrestrial algal assemblages occurring at the bases of urban walls in
Europe. Phycologia 43, 225–235 (2004).
53. Stonyeva, M. P., Uzunov, B. A. & Gärtner, G. Aerophytic green algae, epimycotic on Fomes fomentarius (L. ex Fr.) Kickx. Annu.
Soa Univ “St. Kliment Ohridski”. Fac. Biol. 99, 19–25 (2015).
54. Aras, S. & Cansaran, D. Isolation of DNA for sequence analysis from herbarium material of some lichen specimens. Turk. J. Bot.
30, 449–453 (2006).
55. Hall, T. BioEdit: A userfriendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids
Symp. Ser. 41, 95–98 (1999).
56. Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment soware version 7: Improvements in performance and usability.
Mol. Biol. Evol. 30, 772–780 (2013).
57. Posada, D. jModelTest: Phylogenetic model averaging. Mol. Biol. Evol. 25, 1253–1256 (2008).
58. Huelsenbeck, J. P. & Ronquist, F. MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17, 754–755 (2001).
59. Vondrák, J. & Kubásek, J. Algal stacks and fungal stacks as adaptations to high light in lichens. Lichenol. 45, 115–124 (2013).
60. Kubásek, J., Hájek, T. & Glime, J. M. Bryophyte photosynthesis in sunecks: Greater relative induction rate than in tracheophytes.
J. Bryol. 36, 110–117 (2014).
61. Kubásek, J. et al. Moss stomata do not respond to light and CO2 concentration but facilitate carbon uptake by sporophytes: A gas
exchange, stomatal aperture, and C-13-labelling study. New Phytol. 230, 1815–1828 (2021).
62. Feige, G. & Kremer, B. Unusual carbohydrate pattern in Trentepohlia species. Phytochemistry 19, 1844–1845 (1980).
63. Tonon, T., Li, Y. & McQueen-Mason, S. Mannitol biosynthesis in algae: More widespread and diverse than previously thought.
New Phytol. 213, 1573–1579 (2017).
64. Gustavs, L., Görs, M. & Karsten, U. Polyol patterns in biolm-forming aeroterrestrial green algae (Trebouxiophyceae, Chlorophyta).
J. Phycol. 47, 533–537 (2011).
Acknowledgements
Linda in Arcadia kindly revised the manuscript. Ondřej Peksa and Zdeněk Palice kindly consulted some points.
Our research received support by a long-term research development grant RVO 67985939andby the Technology
Agency of the Czech Republic, grant TH03030469.
Author contributions
J.V. designed the research, J.V. and S.S. conducted eldwork and sample studies, J.K. performed physiological
experiments, L.Z., V.P., L.S. and J.M. provided algological and mycological expertise. J.K. analysed DNA sequence
data. All authors contributed to the writing of the manuscript.
Competing interests
e authors declare no competing interests.
Additional information
Supplementary Information e online version contains supplementary material available at https:// doi. org/
10. 1038/ s41598- 023- 29384-4.
Correspondence and requests for materials should be addressed to J.K.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access is article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons licence, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this licence, visit http:// creat iveco mmons. org/ licen ses/ by/4. 0/.
© e Author(s) 2023
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Notably, Asterochloris asteroidea can form mutualism with Mortierella elongata, where algal cells become internalized within the fungal hyphae, resulting in a photosynthetic hypha capable of fixing CO 2 (Du et al. 2019). This symbiosis, termed "alcobiosis," is common in corticioid fungi and may be linked to fungi's light sensing and responses (Yu and Fischer 2019;Vondrák et al. 2023). This find might highlight the role of soil biofilm for CO 2 fixing through symbiotic structure. ...
Article
Full-text available
Despite the major influence of soils on climate change, carbon sequestration, pollution remediation, and food security, soil remains a largely unexplored media with an extreme complexity of microbes, minerals, and dead organic matter, most of them being actually poorly known. In particular, soil biofilms have recently attracted attention because they strongly influence biogeochemical reactions and processes. Here we review biofilms with focus on their behavior, proliferation, distribution, characterization methods, and applications. Characterization methods include optical, electron, scanning probe, and X-ray microscopy; metagenomics, metatranscriptomics, metaproteomics, metabolomics; and tracking approaches. Applications comprise pollution remediation by metal immobilization or organics degradation; and methane oxidation, carbon dioxide reduction, and carbon sequestration. Advanced methods such as DNA-stable isotope probing and meta-omics have uncovered the multiple functions of soil biofilms and their underlying molecular mechanisms. Investigations have improved our understanding of inter- and intra-kingdom interactions, and of gene transfer. Extracellular materials such as polysaccharides enhance the transport of substances and electrons flow among microorganisms.
... The surface of polypore sporocarps can be a substratum for many organisms such as epiphytic algae (Zavada & Simoes 2001, Stonyeva et al. 2015, Vondrák et al. 2023, non-lichenized (e.g., Hutchison 1987, Sun et al. 2019, Maurice et al. 2021) and lichenized fungi (Hawksworth et al. 2014), forming so-called epimycotic or fungicolous communities (Stonyeva et al. 2015, Maurice et al. 2021. For example, a globally distributed genus of poroid white-rotting fungi, Trichaptum (Hymenochaetales, Agaricomycetes, Basidiomycota; Larsson et al. 2006), may host more than 20 green algal (Chlorophyta) species (Mukhin et al. 2018) and is one of the richest among polypores in terms of associated fungicolous fungi (Maurice et al. 2021). ...
Article
The globally distributed genus Trichaptum is one of the most species-rich among polypores in terms of hosting other fungi. Among Trichaptum -associates, there is a group of mazaediate lichenized fungi ( Coniocybomycetes , Ascomycota ) that previously had an uncertain phylogenetic position. DNA sequences – mitochondrial small subunit (mtSSU), nuclear large subunit rDNA (nuLSU), and internal transcribed spacer (ITS) – were obtained from 29 specimens collected from Europe and North America. Maximum likelihood and Bayesian inference analyses of these three gene loci were used to infer phylogenetic position and relationships among lineages. Statistical tests were used to find which phenotypical characteristics distinguish species. The molecular sequence data provide evidence that the fungicolous specimens form a distinct lineage within Coniocybomycetes sister to the combined clade of Chaenotheca s. lat. and Sclerophora . Considering its phylogenetic placement and strict specialization, we describe a new genus – Chaenotricha . This fungicolous lineage contains three species based on molecular characteristics. Morphological characters mostly overlap except for spore size and stalk length of apothecia. We provide a new combination, Chaenotricha obscura , for the only previously described species for which we designate an epitype, and introduce a new species – Chaenotricha cilians . The third lineage remains undescribed because of a small sample size, which did not allow us to clearly delineate species boundaries.
... While these practices sometimes harbour anthropocentric biases, they hint at the potential for balanced, mutualistic interspecies interactions. The recent discovery of alcobiosis [75] adds more interesting insights into symbiotic associations. This relationship between select algae and fungi challenges conventional symbiotic paradigms. ...
Thesis
Climate change, water and food supply, epidemics and extreme events represent growing concerns for our society. While one approach is to rely solely on technological strategies, an alternative is to rethink the socioeconomic system holistically. Our research pursues this latter approach by looking for ecological insights that could better inform decision-makers in facing this transformation. We aim at deriving management solutions for natural resources within planetary boundaries, using a hybrid approach (denominated "Ecomirroring") that combines nature-based solutions and biomimicry-based problem-solving. By implementing ecosystems' adaptation and resilience strategies, we seek to reimagine a future society that shifts away from economic growth and embraces the complex interdependence between human well-being and ecosystem health.
Article
The first findings of recently described polypore fungus Sidera tibetica are revealed in the Caspian lowland forests of the Samursky National Park in the Republic of Dagestan, Russia. Basidiomata collected on Pinus brutia var. eldarica were studied by both light microscopy and molecular methods. Detailed information on new localities and habitats of Sidera tibetica as well as the collection numbers of specimens deposited in the Mycological Herbarium of the Komarov Botanical Institute of RAS (LE) are provided. Two complete sequences of ITS1-5.8S-ITS2 nuclear ribosomal DNA for Caucasian specimens of S. tibetica have been obtained and submitted to the GenBank database. Newly generated sequences are formed a separate well-supported clade in the Maximum Likelihood phylogenetic analysis together with all available ITS nrDNA sequences of S. tibetica originated from Belarussian and Chinese collections including the reference sequence from a holotype.
Article
Full-text available
Biodiversity is a key criterion in nature protection and often indicates habitats and localities rich in endangered species. Our research, using 48 one-man one-day field trips, located an exceptional lichen diversity hotspot and refugium for rare species, the Týřov National Nature Reserve (Czech Republic, central Bohemia). Within its 410 hectares, we detected 787 species of lichens and related taxa (675 lichens, 35 semilichens, 58 lichenicolous fungi and 19 bark microfungi). This is more species of these organisms than has ever been recorded from such a small area, up to 10 km2 , anywhere in Europe (and probably anywhere in the world). The species richness is positively correlated with the habitat heterogeneity within Týřov, which is very far from uniform. In most of the reserve, the species richness is fairly typi�cal for the broader region, and only three sites, with an overall area of a mere 80 hectares, have distinctly higher species richnesses. The most species-rich site, with 502 species, is only about 25 hectares and is distinctly more diverse in habitats than other sites. The enormous importance of Týřov for biodiversity protection is emphasized by the nine species described as new to sci�ence: Acarospora fissa, Bacidia hyalina, Buellia microcarpa, Micarea substipitata, Micro�calicium minutum, Rufoplaca griseomarginata, Verrucaria substerilis, V. tenuispora and V. teyrzowensis. Three species are new to Europe, 55 to the Czech Republic and 191 species are included in the national Red-list.
Article
Full-text available
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).
Article
Full-text available
Lichens are symbiotic associations resulting from interactions among fungi (primary and secondary mycobionts), algae and/or cyanobacteria (primary and secondary photobionts), and specific elements of the bacterial microbiome associated with the lichen thallus. The question of what is a species, both concerning the lichen as a whole and its main fungal component, the primary mycobiont, has faced many challenges throughout history and has reached new dimensions with the advent of molecular phylogenetics and phylogenomics. In this paper, we briefly revise the definition of lichens and the scientific and vernacular naming conventions, concluding that the scientific, Latinized name usually associated with lichens invariably refers to the primary mycobiont, whereas the vernacular name encompasses the entire lichen. Although the same lichen mycobiont may produce different phenotypes when associating with different photobionts or growing in axenic culture, this discrete variation does not warrant the application of different scientific names, but must follow the principle "one fungus = one name". Instead, broadly agreed informal designations should be used for such discrete morphologies, such as chloromorph and cyanomorph for lichens formed by the same mycobiont but with either green algae or cyanobacteria. The taxonomic recognition of species in lichen-forming fungi is not different from other fungi and conceptual and nomenclatural approaches follow the same principles. We identify a number of current challenges and provide recommendations to address these. Species delimitation in lichen-forming fungi should not be tailored to particular species concepts but instead be derived from empirical evidence, applying one or several of the following principles in what we call the LPR approach: lineage (L) coherence vs. divergence (phylogenetic component), phenotype (P) coherence vs. divergence (morphological component), and/or reproductive (R) compatibility vs. isolation (biological component). Species hypotheses can be established based on either L or P, then using either P or L (plus R) to corroborate them. The reliability of species hypotheses depends not only on the nature and number of characters but also on the context: the closer the relationship and/or similarity between species, the higher the number of characters and/or specimens that should be analyzed to provide reliable delimitations. Alpha taxonomy should follow scientific evidence and an evolutionary framework but should also offer alternative practical solutions, as long as these are scientifically defendable. Taxa that are delimited phylogenetically but not readily identifiable in the field, or are genuinely cryptic, should not be rejected due to the inaccessibility of proper tools. Instead, they can be provisionally treated as undifferentiated complexes for purposes that do not require precise determinations. The application of infraspecific (gamma) taxonomy should be restricted to cases where there is a biological rationale, i.e., lineages of a species complex that show limited phylogenetic divergence but no evidence of reproductive isolation. Gamma taxonomy should not be used to denote discrete phenotypical variation or ecotypes not warranting the distinction at species level. We revise the species pair concept in lichen-forming fungi, which recognizes sexually and asexually reproducing morphs with the same underlying phenotype as different species. We conclude that in most cases this concept does not hold, but the actual situation is complex and not necessarily correlated with reproductive strategy. In cases where no molecular data are available or where single or multi-marker approaches do not provide resolution, we recommend maintaining species pairs until molecular or phylogenomic data are available. This recommendation is based on the example of the species pair Usnea aurantiacoatra vs. U. antarctica, which can only be resolved with phylogenomic approaches, such as microsatellites or RADseq. Overall, we consider that species delimitation in lichen-forming fungi has advanced dramatically over the past three decades, resulting in a solid framework, but that empirical evidence is still missing for many taxa. Therefore, while phylogenomic approaches focusing on particular examples will be increasingly employed to resolve difficult species complexes, broad screening using single barcoding markers will aid in placing as many taxa as possible into a molecular matrix. We provide a practical protocol how to assess and formally treat taxonomic novelties. While this paper focuses on lichen fungi, many of the aspects discussed herein apply generally to fungal taxonomy. The new combination Arthonia minor (Lücking) Lücking comb. et stat. nov. (Bas.: Arthonia cyanea f. minor Lücking) is proposed.
Article
Full-text available
Stomata exert control on fluxes of CO2 and water (H2O) in the majority of vascular plants and thus are pivotal for planetary fluxes of carbon and H2O. However, in mosses, the significance and possible function of the sporophytic stomata are not well understood, hindering understanding of the ancestral function and evolution of these key structures of land plants. Infrared gas analysis and ¹³CO2 labelling, with supporting data from gravimetry and optical and scanning electron microscopy, were used to measure CO2 assimilation and water exchange on young, green, ± fully expanded capsules of 11 moss species with a range of stomatal numbers, distributions, and aperture sizes. Moss sporophytes are effectively homoiohydric. In line with their open fixed apertures, moss stomata, contrary to those in tracheophytes, do not respond to light and CO2 concentration. Whereas the sporophyte cuticle is highly impermeable to gases, stomata are the predominant sites of ¹³CO2 entry and H2O loss in moss sporophytes, and CO2 assimilation is closely linked to total stomatal surface areas. Higher photosynthetic autonomy of moss sporophytes, consequent on the presence of numerous stomata, may have been the key to our understanding of evolution of large, gametophyte‐independent sporophytes at the onset of plant terrestrialization.
Article
Full-text available
Phylogenetic analyses have revealed that the traditional order Prasiolales, which contains filamentous and pseudoparenchymatous genera Prasiola and Rosenvingiella with complex life cycle, also contains taxa of more simple morphology such as coccoids like Pseudochlorella and Edaphochlorella or rod-like organisms like Stichococcus and Pseudostichococcus (called Prasiola clade of the Trebouxiophyceae). Recent studies have shown a high biodiversity among these organisms and questioned the traditional generic and species concept. We studied 34 strains assigned as Stichococcus, Pseudostichococcus, Diplosphaera and Desmocococcus. Phylogenetic analyses using a multigene approach revealed that these strains belong to eight independent lineages within the Prasiola clade of the Trebouxiophyceae. For testing if these lineages represent genera, we studied the secondary structures of SSU and ITS rDNA sequences to find genetic synapomorphies. The secondary structure of the V9 region of SSU is diagnostic to support the proposal for separation of eight genera. The complex taxonomic history was summarized and revised. The ITS-2/CBC approach was used for species delimitation. Considering all these results, we revised the genera Stichococcus, Pseudostichococcus, Diplosphaera and Desmococcus and proposed four new genera and four new species for the science community. The usage of the V9 region and the ITS-2 barcodes discovered potential new species among the Stichococcus-like organisms in culture-independent studies.
Article
Full-text available
Conclusions: a streptophyte algal perspective on land plant trait evolution 1432 Acknowledgements 1432 ORCID 1433 References 1433 SUMMARY: Photosynthetic eukaryotes thrive anywhere there is sunlight and water. But while such organisms are exceptionally diverse in form and function, only one phototrophic lineage succeeded in rising above its substrate: the land plants (embryophytes). Molecular phylogenetic data show that land plants evolved from streptophyte algae most closely related to extant Zygnematophyceae, and one of the principal aims of plant evolutionary biology is to uncover the key features of such algae that enabled this important transition. At the present time, however, mosaic and reductive evolution blur our picture of the closest algal ancestors of plants. Here we discuss recent progress and problems in inferring the biology of the algal progenitor of the terrestrial photosynthetic macrobiome.