Teloschistaceae

Source: Wikipedia, the free encyclopedia.

Teloschistaceae
bushy dark orange lichen growing on a tree branch
Teloschistes flavicans is the type species of the type genus of the family Teloschistaceae.
Scientific classification Edit this classification
Domain: Eukaryota
Kingdom: Fungi
Division: Ascomycota
Class: Lecanoromycetes
Order: Teloschistales
Family: Teloschistaceae
Zahlbr. (1898)
Type genus
Teloschistes
Norman (1853)
Subfamilies

Caloplacoideae – 37 genera Teloschistoideae – 33 genera Xanthorioideae – 45 genera

Synonyms[1]
  • Caloplacaceae Zahlbr. (1907)

The Teloschistaceae are a large

arid ecosystems during the Late Cretaceous
.

Teloschistaceae lichens typically have one of a few physical

green algal genus Trebouxia. Teloschistaceae members are also characterised by their apothecia (the fruiting bodies where sexual reproduction occurs), which generally have a well-defined encircling rim of tissue. In the Teloschistaceae, the tip of the ascus, the structure that produces spores, characteristically turns blue when stained with iodine. The ascospores
are released through a longitudinal slit in the ascus tip, a unique trait common to this family of lichens.

The family, first formally proposed in 1898, was extensively revised in 2013, including the creation or resurrection of 31

genera. Three subfamilies – Caloplacoideae, Teloschistoideae, and Xanthorioideae – are recognised. Since 2013, several dozen new genera have been added to the family, but there has been some debate about these additions. Ongoing DNA studies are helping to provide clearer insights into how the different groups within this family are related. The family contains more than 800 species in around 120 genera. Three species from the Teloschistaceae have been globally assessed for conservation status and others, such as the rare New Zealand species Caloplaca allanii, appear on regional lists. The full diversity of this family remains underexplored in vast regions like South America and China. Regarding human interactions and applications, although lacking any major economic impact, several rock-dwelling Teloschistaceae species are known to damage marble surfaces, and others are used in some traditional medicines. One member, Rusavskia elegans, is used in research as a model organism to investigate resilience against the harsh conditions of outer space
.

Systematics

Historical taxonomy

Vibrant yellow lichen with orange centers clustered on a dark background.
Yellow fruticose lichen with orange apothecia growing on a tree branch.
The common and widespread Teloschistaceae species Xanthoria parietina (left) and Teloschistes chrysophthalmus (right) were two of the earliest lichens to be formally described.

The first members of the present-day Teloschistaceae to be formally described were the common sunburst lichen (Xanthoria parietina) and the gold-eye lichen (Teloschistes chrysophthalmus). These were two of several dozen lichen species described by the Swedish taxonomist Carl Linnaeus, the former in his influential 1753 treatise Species Plantarum,[2] and the latter in his 1771 work Mantissa Plantarum Altera.[3]

In his 1852 work Synopsis Lichenum Blasteniosporum ('Synopsis of Lichen Blasteniospores'),[4] the lichenologist Abramo Bartolommeo Massalongo attempted to classify what he called "blasteniospore lichens". This term referenced species, diverse in growth forms and appearance, united by the distinct polarilocular spores now attributed to the family Teloschistaceae. These are spores that are divided into two compartments (locules) separated by a central septum with a small hole. Although Massalongo's efforts to arrange these taxa into more natural genera were largely ignored by subsequent researchers, several of his proposed genera were resurrected for use 16 decades later, such as Blastenia, Gyalolechia, Pyrenodesmia, and Xanthocarpia.[5]

Microscopic view of an oval-shaped lichen ascocarp with a cluster of circular, brownish-orange spores, each displaying intricate, radiating patterns.
An ascus of the crustose lichen Pyrenodesmia variabilis, containing eight polarilocular ascospores

The

fruticose taxa having polarilocular (i.e. two-locule) or four-locule ascospores, including the genera Xanthoria, Teloschistes, and Lethariopsis.[6] At that time, the growth form of the lichen thalli was often used in classical lichen taxonomy to segregate groups of species into families,[7] and so in a subsequent (1926) publication, Zahlbruckner introduced the family Caloplacaceae to contain crustose lichens with polarilocular ascospores; this family included the genera Caloplaca, Blastenia, Bombyliospora, and Protoblastenia.[8] The distinctness of the family Caloplacaceae was largely rejected by other authors,[9] and it is now a historical synonym of Teloschistaceae.[1] In another older classification, crustose genera were grouped together in the family Blasteniaceae[10] or the Placodiaceae.[11][note 1] In 1971, Carroll William Dodge proposed the family Xanthoriaceae to contain Xanthodactylon, Xanthopeltis, and Xanthoria,[14] but it was not validly published.[9]

In the 20th century, particularly with the widespread use of

Etymology

As is standard practice in botanical nomenclature,[26] the name Teloschistaceae is based on the name of the type genus, Teloschistes, with the ending -aceae indicating the rank of family. The genus name, assigned by the Norwegian botanist Johannes M. Norman in 1852,[27] comprises two Ancient Greek words: τέλος (télos), meaning 'end', 'final', or 'term'; and σχιστός (-schistós), meaning 'divided into', 'split', or 'separated'. It refers to the split ends of the thallus branches that are characteristic of that genus.[28]

Subfamilial and ordinal classification

Teloschistales
Teloschistaceae
Brigantiaeaceae
phylogeny of some species and genera in family Teloschistaceae and in the order Teloschistales; based on analysis by Wilk et al. in 2021 (simplified from original).[29]
Species names have been updated to reflect current taxonomy. Single quote marks around a genus name means that the species is likely misclassified.

Teloschistaceae is divided into three recognised subfamilies: Xanthorioideae, Caloplacoideae, and Teloschistoideae.[30] In 2015, researchers proposed a fourth subfamily, Brownlielloideae,[31] which was later shown by genetic studies to be a grouping based on mixed or misinterpreted data rather than a distinct lineage.[32][33] Further analysis placed what was thought to be Brownlielloideae within the already established Teloschistoideae, suggesting the proposed subfamily was not a separate branch of the family tree.[34] DNA evidence also dispersed members of the informally introduced subfamily Ikaerioideae across the three acknowledged subfamilies, primarily within Teloschistoideae.[35] Despite this, Sergey Kondratyuk and colleagues continue to use Brownlielloideae and Ikaerioideae in their publications, assigning nine genera to the former and two to the latter.[36] The well-supported subfamilies (Xanthorioideae, Caloplacoideae, and Teloschistoideae) encompass a range of growth forms – crustose, foliose, and fruticose – demonstrating the diverse evolutionary paths within the family.[37] These groups are genetically distinct, each subfamily showing unique patterns in their nuclear large ribosomal subunit RNA sequences.[38]

  • Caloplacoideae Arup, Søchting & Frödén (2020)
    • Type genus: Caloplaca. Proposed by Ester Gaya and colleagues in 2012 and validly published in 2020, Caloplacoideae consists mostly of crustose lichens with a wide geographical spread and produces a range of unique chemical compounds.[39]
  • Teloschistoideae Arup, Søchting & Frödén (2020)
    • Type genus: Teloschistes. Initially proposed in 2013 and validly published with a full diagnosis in 2020, this subfamily is predominantly found in the Southern Hemisphere.[40]
  • Xanthorioideae Arup, Søchting & Frödén (2020)
    • Type genus: Xanthoria. Named by Gaya and colleagues in 2012 and formally validated in 2020, Xanthorioideae species are primarily distributed in the Northern Hemisphere.[41]

The order Teloschistales was first proposed by David Hawksworth and Eriksson in 1986, with a single family (Teloschistaceae); other families were added later.[42] In the 1990s, several authors recognised the Teloschistales as a suborder within the Lecanorales;[43] as a suborder it was named Teloschistineae.[44] Following the appearance of preliminary molecular studies,[45] the Teloschistaceae was classified by some within the order Lecanorales, although others maintained the Teloschistales as a valid order.[46] A large-scale, multigene phylogenetic study of the class Lecanoromycetes published in 2014 corroborated the ordinal status of the Teloschistales, and showed it comprises two clades: Letrouitineae (containing Brigantiaeaceae and Letrouitiaceae) and its sister clade, Teloschistineae (containing Teloschistaceae and Megalosporaceae).[47] The suborder Teloschistineae was formally proposed by Ester Gaya and François Lutzoni in 2016.[48]

Molecular phylogenetics

Historically, classification of taxa within the family relied on physical characteristics such as growth form, the nature of the outer layer of the lichen (the cortex), and spore type. Studies using modern molecular phylogenetics have shown that phenotypic characteristics (characters) are not always reliable markers of phylogenetic relationships, and classifying species based on these characters has occasionally led to inaccurate interpretations of their evolutionary history.[49] Advanced DNA analysis techniques have allowed scientists to identify and differentiate cryptic species, which, though visually indistinguishable, are genetically distinct. This approach has unveiled distinct species within previously thought homogeneous groups, like the genus Caloplaca, by uncovering their unique genetic markers.[50]

Although Teloschistaceae is now well represented in GenBank, with thousands of DNA sequences, the early molecular studies were limited by having too few examples of each species to draw definitive conclusions.[36][51][52] With the increasing availability of genetic sequences, researchers began to gain a better understanding of the family's phylogeny.[22]

One significant finding from molecular data is that the traditional

monophyletic), but is now understood to have been composed of multiple, unrelated groups (polyphyletic). This insight has prompted numerous proposals to redefine the genus into smaller, monophyletic groups;[53] but such taxonomic changes have sometimes met with resistance due to the vast number of species reclassifications they would entail.[54]

According to the lichenologist Robert Lücking, families like Teloschistaceae, which have undergone several changes in genus classification through various studies, require phylogenetic consolidation through extensive multi-locus analysis, incorporating all available data and employing rigorous analytical methods. This strategy, akin to approaches taken with families such as Collemataceae, Graphidaceae, Pannariaceae, and Parmeliaceae, is essential for accurately revising the taxonomic classification of this diverse and widespread group of lichens.[55]

Molecular evidence has also helped to map the family's relationships within the class Lecanoromycetes. A 2018 study identified the Megalosporaceae as the Teloschistaceae's closest relative.[56]

Description

In general, Teloschistaceae members are known for their vibrant colours, spanning a spectrum of yellow, orange, and red hues, attributed to

fruticose) formations.[1][57] Although it is an atypical growth form for the Teloschistaceae, members of genus Ioplaca are somewhat umbilicate, meaning they have a somewhat circular, leafy thallus attached to the substrate at a single point.[58]

A group of orange, disc-shaped lichen on a grey stone surface.
In contrast to its barely visible thallus, the apothecia of Athallia holocarpa are prominent. Like most Teloschistaceae, these apothecia are lecanorine in form: rounded with thick, pale, thalline margins.

Teloschistaceae lichens have a symbiotic relationship with a

soredia, can be found in select species.[1]

The ascomata encase asci, cylindrical formations that commonly contain between four and sixteen ascospores, with eight being the most prevalent count. These asci are characterised by a well-developed J+ layer amyloid cap; the term "J+" refers to the positive staining reaction of the ascus tip to iodine, specifically when it turns blue or dark blue in the presence of iodine-based solutions like Melzer's reagent or Lugol's iodine. The internal apical structure of the ascus is rudimentary, relative to the more complex apical structures in other related families.[57]

Teloschistaceae ascospores, typically translucent, usually have one to three septa (internal partitions) with a prominent central septum connected by a canal to the spores' internal cavities, or lumina.[57][1] Although the presence of a two-chambered (polarilocular) structure in these ascospores is generally indicative of the Teloschistaceae, the spores lack other distinctive features that could be useful as defining taxonomic characteristics.[59] Historically, polarilocular ascospores were regarded as a hallmark of the Teloschistaceae. The incorporation of genera such as Apatoplaca, Cephalophysis, Fulgensia, and Xanthopeltis, which have either non-septate or simply septate spores, has necessitated a reassessment of what fundamentally characterises this group.[9]

A distinctive feature of Teloschistaceae is the presence of the gelatinous paraphyses (filament-like support structures in the reproductive apparatus), with either unbranched or slightly branched structures culminating in bulbous ends.[1] Within this family, asexual reproduction leads to the creation of pycnidia-type conidiomata (i.e., small, flask-shaped fruiting bodies), producing translucent asexual spores (conidia) that are either bacillar (rod-shaped) or bifusiform (double-spindle shaped).[1][57] The tissue composition of the thallus and apothecia is characterised by a loosely paraplectenchymatous structure, meaning the constituent fungal hyphae are oriented in different directions.[15]

Morphological variety of Teloschistaceae

Photobionts

In lichens, photobionts are the photosynthetic organisms that collaborate with fungal partners to enable the unique lichen symbiosis. Members of the Teloschistaceae associate with trebouxioid green algal photobionts (i.e. resembling or belonging to the green algal genus Trebouxia). An early study investigating the ultrastructure of the interaction between the fungus and alga in several Teloschistaceae species showed that, in most cases, the cells were merely in close proximity to one another, with only a few instances of fungal cells invading the algal cells.[60] The widespread Xanthoria parietina species complex has been identified to associate with various trebouxioid species, including Asterochloris italiana, Trebouxia arboricola, and T. decolorans.[61] Within the order Teloschistales, unlike the Teloschistaceae, species in the families Letrouitiaceae and Megalosporaceae primarily partner with the green algal genus Dictyochloropsis. Due to their resilience to desiccation, Trebouxia species serve as the main photobionts for lichen-forming fungi found in extreme environments such as the Antarctic, Arctic, alpine regions, and deserts, where lichens face continual exposure to intense dryness and temperature shifts.[47]

Studies of photobionts in the Teloschistaceae, including foliose genera (Xanthoria, Xanthomendoza) and a fruticose genus (Teloschistes), demonstrate a consistent association with specific Trebouxia clades. This finding suggests a genus-level specificity, with only select Trebouxia subclades forming symbiotic relationships. This specificity is not absolute and may vary with habitat: lichens in extreme climates have been observed to associate with a broader range of photobionts.[61]

Chemistry

chemical structure
chemical structure
Parietin (top) and the structurally similar emodin (bottom) are anthraquinone pigments common in the Teloschistaceae.

The predominant chemical compounds found in the Teloschistaceae are pigments known as

ABC transporter gene in the pigment gene cluster suggests a mechanism for how lichens accumulate substantial amounts of potentially toxic anthraquinone crystals in their thallus and reproductive structures.[62]

Between 1897 and 1906, the mycologist Friedrich Wilhelm Zopf and the chemist Oswald Hesse conducted a series of early chemical studies on members of the Teloschistaceae, leading to the extraction of the reddish pigment parietin from selected species.[64] Parietin is an antioxidant molecule that is produced in greater amounts in lichen thalli that are exposed to excess nitrogen.[65] In a 1970 publication, Johan Santesson surveyed 230 Caloplaca species for anthraquinones as part of a phytochemical study of the Teloschistaceae, and concluded that the studied species could be arranged according to their anthraquinone content in thirteen chemical groups.[64] In 1997, Ulrik Søchting analysed secondary metabolites from species of Caloplaca, Teloschistes, and Xanthoria to look for chemical patterns of consistent combinations and proportions of lichen products. He identified two chemosyndromes (characteristic sets of chemical compounds) with parietin, emodin, fallacinol, fallacinal, and parietinic acid as the main substances.[66] Parietin acts as a UV-light filter to provide optimal light intensities for the photobionts that are resident in the internal algal layer. Studies show that in Xanthoria parietina, the more light the lichen is exposed to, the higher the concentration of parietin. In the Teloschistaceae, parietin may also serve a defensive role. In the Negev desert, the parietin-containing Teloschistaceae species Elenkiniana ehrenbergii and Seirophora lacunosa are avoided by grazing snails, but they frequently consume lichens like Diploicia canescens and Buellia subalbula (both in family Caliciaceae), which lack parietin.[67]

In their large-scale phylogenetic analysis of the Teloschistaceae, the lichenologists Ulf Arup, Ulrik Søchting, and Patrik Frödén analysed about 4000 specimens of members of the family using high-performance liquid chromatography, and identified more than 100 secondary metabolites, mostly anthraquinones. They noted that in the large majority of cases, the distribution of lichen products was more or less constant within species. In some instances, the secondary chemistry is important at higher taxonomic levels (i.e., ranks higher than species).[68] For example, the genus Catenaria is characterised by the presence of 7-chlorocatenarin, a secondary metabolite previously unknown in lichens.[69] The substance usnic acid characterises the genus Usnochroma, and 5-chloroemodin occurs in all but one species of Shackletonia. The secondary chemistry of the Caloplacoideae is the most diverse amongst the three Teloschistaceae subfamilies, as it contains both chlorinated anthraquinones and depsidones.[68]

Several irregularly disc-shaped, orange structures with a black rim on a whitish, rough-textured surface
Apothecia of Ikaeria serusiauxii grown in full light develop completely black margins resulting from the accumulation of the pigment Cinereorufa-green.[70]

Although most Teloschistaceae lichens produce anthraquinone pigments in shades ranging from yellow to orange to red, the genera

synapomorphic character, but appears independently in unrelated lineages of Teloschistaceae; as such, it is a phylogenetically unreliable character.[72]

Adaptive radiation

Adaptive radiation in the Teloschistaceae has been studied to understand the key phenotypic changes leading to their diversification. This diversification is believed to be connected to the spread of anthraquinone pigments in their thallus. Initially, these pigments were thought to have appeared during the Teloschistaceae's first divergence, with a more widespread occurrence developing later. The distribution of anthraquinones in Teloschistaceae lichens varies, from being dispersed across the organism's surface to localised regions. Analysis suggests that the family's lineage witnessed a loss and subsequent return of these pigments over time, considering their presence in the thallus and apothecia as the ancestral state. Ecologically, these organisms transitioned from inhabiting shaded bark environments to colonising sunlit, rocky areas during their diversification.[73]

The analysis of phenotypic traits and diversification rates shows that anthraquinones in the thallus and greater sun exposure have contributed to an acceleration of diversification. On the contrary, living in shaded environments or having a crustose-continuous (smooth, non-scaly) growth form hindered diversification. The choice of substrate, be it rock or bark, did not have a pronounced impact on diversification rates. This adaptive radiation within the Teloschistaceae is estimated to have initiated around 100 million years ago, specifically during the Late Cretaceous period. Factors like climatic shifts, continental separations, and the emergence of flowering plants are theorised to have influenced the adaptive landscape. Such factors might have promoted the development of light-protective anthraquinones, enabling Teloschistaceae to colonise exposed environments.[73] The diversification of anthraquinone genes in their evolution is primarily due to gene reshuffling, which has given rise to novel biosynthetic enzyme pathways and gene clusters.[62]

Genera

The classification and number of species within the Teloschistaceae have evolved significantly over time. Historical estimates have varied, with 10 genera and 47 species recognised in 2001,

Species Fungorum (in the Catalogue of Life), accepts 117 genera and 805 species in the Teloschistaceae. The largest genus is Caloplaca, at 173 accepted species.[78][note 3]

In terms of species diversity, Teloschistaceae stood as the sixth-largest lichen-forming fungal family by 2017, following Parmeliaceae, Graphidaceae, Verrucariaceae, Ramalinaceae, and Lecanoraceae.[76] Each genus is paired with its taxonomic authority, denoting the first describers using standardised author abbreviations, the publication year, and the number of species. Genera are organised by subfamily:

Caloplacoideae

Igneoplaca ignea; and Kuettlingeria erythrocarpa

Teloschistoideae

bright orange crust-like growth on a whitish rock
shrubby yellow-orange lichen with many small orange circular structures
brownish crust-like growth on a rock
crust-like orangish growth on quartz-like rock
Examples from subfamily Teloschistoideae (clockwise from upper left): Neobrownliella cinnabarina; Niorma hosseusiana; Wetmoreana brouardii; and Fulgogasparrea appressa

Xanthorioideae

Examples from subfamily Xanthorioideae (clockwise from upper left): Athallia scopularis; Calogaya saxicola; Dufourea ligulata; and Xanthocarpia crenulatella

Invalid names

Some of the genera proposed during the restructuring of the family have since been shown to be nomenclaturally illegitimate or unavailable for use. For example:

  • Andina Wilk, Pabijan & Lücking (2021) has been replaced with Wilketalia.[150]
  • Phaeoplaca Søchting, Arup & Bungartz (2020) has been replaced with Obscuroplaca.[104]
  • Tayloriella S.Y.Kondr., Kärnefelt, A.Thell, Elix & Hur (2015) has been replaced with Tayloriellina.[86]

Habitat, distribution, and ecology

Cluster of bright orange lichen interspersed with yellow and white parts on a green mossy background.
The scrambled egg lichen, Gyalolechia fulgens, is a ground-dwelling species and a component of some biological soil crusts.
Red backpack resting on snow in front of a large orange-stained rock under a clear blue sky.
This rock in Gaspereau Lake is frequented by great black-backed gulls, creating the conditions for a localised nitrogen-rich environment conducive to the growth of the orange lichen Rusavskia elegans.

The family has a

grow on mosses and detritus.[185] Some Fulgensia species grow on soils, particularly those rich in lime.[186] Several crustose Teloschistaceae species, typically saxicolous in nature, have been recorded growing on human bone remains recovered at a looted Late Holocene aboriginal cairn burial site in Argentina.[187]

Many species of the family are moderately to strongly

broad sense – including lichens historically classified in this genus) may be the only genus present; Caloplaca and Xanthoria dominate harsh coastal environments.[67]

There are several Teloschistaceae genera that contain lichenicolous (lichen-dwelling) species. These non-lichenised fungi originate from subfamily Caloplacoideae: Caloplaca (26 spp.), Gyalolechia (1 sp.), Variospora (1 sp.); from subfamily Teloschistoideae: Catenarina (1 sp.), Sirenophila 1; and from subfamily Xanthorioideae: Flavoplaca (4 spp.), Pachypeltis (1 sp.), and Shackletonia (3 spp.).[189] Lichenicolous species within the Teloschistaceae generally have a broad range of hosts. Their geographical distribution seems to be influenced not just by the classification of their host lichen, but also by the substrate they grow on.[190]

Teloschistaceae has a high diversity in polar regions and a substantial number of bipolar species, i.e., species occurring in both northern and southern hemispheres but largely absent from intermediate, tropical latitudes.[191] Examples include Gallowayella borealis, Austroplaca soropelta, and Scythioria phlogina.[192] There is a relatively low diversity of crustose Teloschistaceae in Central Europe. Localised exceptions occur primarily in sunlit locations with either calcareous or nutrient-rich siliceous rock formations; these habitats are predominant in the alpine regions of the Alps and the Carpathian Mountains, as well as in the arid, warm rocky steppes.[193] Some Teloschistaceae genera have a strong geographic centre of species richness; examples include Elixjohnia (Australasia),[120] Orientophila (east Asia), Shackletonia (Antarctic and subantarctic), Stellarangia (south-western Africa), and Xanthoria (Mediterranean area).[41]

Several studies published since 2014 have enumerated the Teloschistaceae taxa occurring in certain defined geographical areas. These include:

Species interactions

Series of six close-up images showing variations of orange lichen with circular, raised fruiting bodies on rocky surfaces.
The lichenicolous Tremella caloplacae species complex on Teloschistaceae hosts: (A) on Xanthoria parietina, (B) on Rusavskia elegans, (C) on Variospora flavescens, (D) on Rusavskia sorediata, (E) on Calogaya pusilla, (F) on Xanthocarpia sp.; white circles enclose Tremella-induced galls. Scale bars = 1 mm.

Teloschistaceae species host a variety of lichenicolous fungi. Some, such as Cercidospora caudata and Stigmidium cerinae, infect a wide range of hosts within the family. More generally, these parasitic fungi have a preference for specific species or genera of Teloschistaceae. An example is the relationship between Teloschistaceae lichens and the fungus Tremella caloplacae. Integrative studies combining molecular data and ecological approaches revealed at least six distinct lineages of T. caloplacae, each specialised to a particular host, indicating a complex of closely related species. This diversification of T. caloplacae appears to have occurred in tandem with the rapid diversification of the Teloschistaceae since the Late Cretaceous period, implying coevolution. Further molecular studies have delineated the T. caloplacae group into a complex of at least nine distinct species. Out of these, five new species were formally described in 2023, each adapted to a single host species or genus within the Teloschistaceae.[200][201]

Human interactions and uses

Igneoplaca ignea covers walls of the Fortaleza de Santa Teresa, a 260-year-old military fortification in Uruguay
.

Although Teloschistaceae species are not economically significant,[1] their tendency to grow on rock surfaces has been documented to cause damage to marble structures. In some cases, the lichens, the major contributor of which was Xanthocarpia feracissima, penetrated up to 10 mm (38 in) into the stone along larger cracks and 0.05 mm (1500 in) beneath loose surface crystals, leading to crumbling of the marble surface.[202] Caloplaca pseudopoliotera and C. cupulifera are two crustose species responsible for the slow degradation of the Konark Sun Temple in India.[203]

Traditional medicine

Some Teloschistaceae species have been used in

antibacterial.[204]

In regional practices, Rusavskia elegans is used for treating wounds; in Afghanistan, it is applied directly. In Kyrgyzstan, the lichen is mixed with butter and used as a remedy for diarrhoea in livestock. Teloschistes flavicans is used in China for its purported properties of "clearing heat" in the lung and liver, and removing toxins. Oxneria fallax has been incorporated into traditional Tibetan medicinal treatments.[204]

In science

The presence of Xanthoria parietina in urban ecosystems, even in high pollution areas, offers potential insights for biomonitoring. Although its tolerance to air pollution allows it to persist where other species may not, its presence can be used in conjunction with more pollution-sensitive species to gauge overall air quality. In polluted environments, the health and abundance of X. parietina, relative to other less tolerant species, can provide data on the level and impact of urban pollution.[205] Rusavskia elegans has been studied in experiments where specimens were exposed to outer space conditions, including extreme temperatures, ultraviolet radiation, and ultra-high vacuum. The results demonstrated the lichen's substantial capability to endure these conditions.[206]

Conservation

Close-up of orange cup-shaped lichen on a rough, grainy surface.
Caloplaca allanii is a poorly known New Zealand endemic.

The conservation status of three Teloschistaceae species has been assessed for the global IUCN Red List. Caloplaca rinodinae-albae (vulnerable, 2017) is at risk from tourism development and increased erosion on Sardinia's coasts.[207] Xanthaptychia aurantiaca (Endangered, 2020) faces multiple threats due to climate change in the Canadian Arctic. These include habitat loss from rapidly eroding coasts, increased sea ice melt, saline wash from storm surges, and permafrost melting. Further, the changing climate may allow the advancement of southern vegetation communities and the introduction of invasive species, potentially exacerbating the impact on this lichen by altering its native habitat.[208] Teloschistes peruensis (Critically Endangered, 2021) is at risk due to multiple threats in Peru and Chile, including potential development, habitat fragmentation, 4x4 races like the Dakar Rally, air pollution, and the impact of livestock like goats and cows, which alter the habitat through grazing and trampling.[209]

Other Teloschistaceae members, some with limited geographic distributions, make appearances on regional red lists. For example, the crustose New Zealand endemic Caloplaca allanii, first documented in 1932, was not collected again until 81 years later. Because of its sparsity and small total area of occupancy, it has been assessed as "Threatened/Nationally Critical" in the New Zealand Threat Classification System.[210]

In some large geographical areas, the full extent of the diversity of Teloschistaceae taxa is not well known. Examples include South America, where the family has not historically received much attention,[211] and China, where of 2,164 lichen species evaluated for inclusion on its red list, only 49 were members of the Teloschistaceae; 13 of those were listed as least-concern species, and the other 36 as data deficient.[212]

Notes

  1. ^ Neither Blasteniaceae C.W.Dodge & G.E.Baker (1938) nor Placodiaceae A.Fisch. (1871) are synonyms of Teloschistaceae, because neither of these families were published validly.[12][13]
  2. ^ The term "infraspecific" refers to a taxonomic rank below that of species, including subspecies, variety, and form.[24]
  3. Species Fungorum's recognition of Teloschistaceae species is based on their taxonomic evaluation, possibly not encapsulating the entirety of the family's species diversity.[79]
  4. ^ a b c Kondratyuk and colleagues suggested that Gyalolechia was polyphyletic, and split it into Elenkiniana, Gyalolechia, and Mikhtomia; Wilk and colleagues maintain Arup et al.'s 2013 classification pending further research.[90]
  5. ^ a b Kondratyuk and colleagues suggested that Variospora was polyphyletic, and split it into Klauderuiella and Variospora; Wilk and colleagues maintain Arup et al.'s 2013 classification pending further research.[90]
  6. ^ Kondratyuk and colleagues suggest that Oceanoplaca is a synonym of Loekoeslaszloa.[106]
  7. ^ Fulgogasparrea was proposed by Kondratyuk et al. in 2013 to resolve putative polyphyly in genus Wetmoreana; Wilk and colleagues instead prefer to use a more broadly defined genus Wetmoreana pending further studies.[124]
  8. ^ Niorma was proposed by Kondratyuk et al. in 2013 for the species group centred around Teloschistes hypoglaucus; Wilk and colleagues instead use the more inclusive Teloschistes as per Arup et al. (2013) pending further studies.[124]
  9. ^ Wilk and colleagues suggest that Raesaeneniana is a putative synonym of Villophora, a genus with which its single species fits well morphologically.[138]
  10. ^ Wilk and colleagues proposed to reduce Tarasginia to synonymy with Sirenophila.[144]
  11. ^ Wilk and colleagues proposed to reduce Tayloriellina to synonymy with Villophora.[146]
  12. ^ Wilk and colleagues suggest that Thelliana is a putative synonym of Filsoniana.[138]
  13. ^ Wilk and colleagues suggest that Dijigiella is a putative synonym of Teuvoahtiana.[138]
  14. ^
    Xanthokarrooa; Wilk and colleagues prefer to retain a more broadly defined Dufourea pending further research.[124]
  15. ^ a b c d e f Kondratyuk and colleagues suggested that Xanthomendoza was polyphyletic, and split it into Gallowayella, Golubkovia, Oxneria, Honeggeria, Jesmurraya, and Xanthomendoza; Wilk and colleagues agree that Xanthomendoza is polyphyletic, but prefer a conservative approach (maintaining Arup et al.'s 2013 classification) pending further research.[124]
  16. ^ Bungartz and colleagues synonymise Huriella with Squamulea.[165]
  17. Xanthodactylon in 2008;[181] other sources place this genus in synonymy with Dufourea.[182][77][154]

References

Citations

  1. ^ .
  2. ^ Jørgensen, James & Jarvis 1994, pp. 334–335.
  3. ^ Jørgensen, James & Jarvis 1994, p. 291.
  4. ^ a b Massalongo, A.B. (1852). "Synopsis Lichenum Blasteniospororum" [Summary of Blasteniosporous Lichens]. Flora or Botanical Newspaper: Which Contains Reviews, Treatises, Essays, News and News Concerning Botany (in Latin). 35: 573.
  5. ^ .
  6. ^ Zahlbruckner, A. (1898). "Flechten (Lichenes). B. Spezieller Teil". In Engler, Adolf (ed.). Syllabus der Pflanzenfamilien 2 [Syllabus of Plant Families 2.] (in German). Berlin: Gebrüder Borntraeger Verlagsbuchhandlung. p. 45.
  7. .
  8. ^ Zahlbruckner, A. (1926). "Lichenes (Flechten). B. Spezieller Teil". In Engler, A.; Prantl, K. (eds.). Die natürlichen Pflanzenfamilien 8, Aufl. 2 [Lichens. B. Special Part]. Leipzig: Engelmann. pp. 61–270 [226].
  9. ^ a b c Gaya et al. 2008, p. 530.
  10. JSTOR 2394232
    . Retrieved 30 October 2023.
  11. ^ Räsänen, Veli (1943). Das System der Flechten [The System of Lichens] (PDF). Acta Botanica Fennica (in German). Vol. 33. Helsinki: A.-G. F. Tilgmann. pp. 1–82 [8, 11, 40].
  12. ^ "Record Details: Blasteniaceae C.W. Dodge & G.E. Baker, Ann. Mo. bot. Gdn 25(2): 604 (1938)". Index Fungorum. Archived from the original on 8 September 2023. Retrieved 23 August 2023.
  13. ^ "Record Details: Placodiaceae A. Fisch. [as 'Placodinae'], Mitt. berin. naturf. Ges.: 17 (1871)". Index Fungorum. Archived from the original on 8 September 2023. Retrieved 23 August 2023.
  14. .
  15. ^ .
  16. .
  17. .
  18. .
  19. . Retrieved 23 August 2023.
  20. ^ Bellemère, A.; Hafellner, J.; Letrouit-Galinou, M.-A. (1986). "Ultrastructure et mode de déhiscence des asques chez les lichens des genres Teloschistes et Apatoplaca (Teloschistaceae)" [Ultrastructure and dehiscence mode of asci in lichens of the genera Teloschistes and Apatoplaca (Teloschistaceae)]. Cryptogamie, Bryologie, Lichénologie (in French). 7 (3): 189–211. Retrieved 23 August 2023.
  21. ^ Kärnefelt, Ingvar (1989). "Morphology and phylogeny in the Teloschistales". Cryptogamic Botany. 1: 147–203.
  22. ^
    PMID 22306043
    .
  23. ^ Eriksson, O.E. (2006). "Outline of Ascomycota – 2006". Myconet. 12: 1–82. Archived from the original on 29 January 2016. Retrieved 13 December 2023.
  24. ^ Kirk et al. 2008, p. 340.
  25. .
  26. .
  27. ^ a b Norman, J.M. (1852). "Conatus praemissus redactionis novae generum nonnullorum Lichenum in organis fructificationes vel sporis fundatae" [The preliminary effort of the new revision of certain genera of lichens based on the fructifications or spores within their structures]. Nytt Magazin for Naturvidenskapene (in Latin). 7: 213–252 [228].
  28. .
  29. ^ Wilk et al. 2021, p. 282.
  30. ^ Arup, Søchting & Frödén 2013, passim.
  31. ^ Kondratyuk et al. 2015b, passim.
  32. .
  33. ^ Wilk et al. 2021.
  34. ^ Wilk et al. 2021, p. 278.
  35. ^ Wilk et al. 2021, p. 283.
  36. ^ .
  37. ^ Arup, Søchting & Frödén 2013, pp. 285–287.
  38. ^ Bungartz, Søchting & Arup 2020, pp. 528, 563.
  39. ^ Bungartz, Søchting & Arup 2020, p. 528.
  40. .
  41. ^ a b c Arup, Søchting & Frödén 2013, p. 34.
  42. ^ Hawksworth, D.L.; Eriksson, O.E. (1986). "The names of accepted orders of ascomycetes". Systema Ascomycetum. 5 (2): 175–184.
  43. .
  44. .
  45. doi:10.1017/S0269-915X(02)00206-9 (inactive 31 January 2024).{{cite journal}}: CS1 maint: DOI inactive as of January 2024 (link
    )
  46. ^ .
  47. ^ .
  48. ^ Lücking, Hodkinson & Leavitt 2017, p. 403.
  49. ^ Arup, Søchting & Frödén 2013, p. 17.
  50. .
  51. .
  52. .
  53. ^ Kondratyuk, S.Y.; Jeong, M.-H.; Kärnefelt, I.; Hur, J.-S.; Thell, Arne (2012). "Філогенія і таксономія родини Teloschistaceae (Ascomycota): важливість моноФілетичних груп" [Phylogeny and taxonomy of the Teloschistaceae (Ascomycota): Importance of monophyletic groups] (PDF). Modern Phytomorphology (in Ukrainian). 1: 53–57.
  54. ^ Kantvilas, Gintaras (2016). "A synopsis and key for the lichen genus Caloplaca (Teloschistaceae) on Kangaroo Island, with the description of two new species". Journal of the Adelaide Botanic Gardens. 29: 53–69.
  55. ^ Lücking, Hodkinson & Leavitt 2017, p. 369.
  56. .
  57. ^ .
  58. ^ .
  59. .
  60. .
  61. ^ .
  62. ^ .
  63. .
  64. ^ .
  65. ^ .
  66. .
  67. ^ .
  68. ^ a b Arup, Søchting & Frödén 2013, p. 29.
  69. ^ .
  70. .
  71. ^ .
  72. .
  73. ^ .
  74. ^ Kirk et al. 2008, p. 680.
  75. ^ Arup, Søchting & Frödén 2013, p. 16.
  76. ^ a b Lücking, Hodkinson & Leavitt 2017, p. 371.
  77. ^ .
  78. Species Fungorum Plus: Species Fungorum for CoL+. "Teloschistaceae". Catalogue of Life
    Version: 2023-07-18. Retrieved 20 January 2024.
  79. ^ "Searching the Index Fungorum database". Index Fungorum. Retrieved 21 December 2023.
  80. ^ Poelt, J.; Hafellner, J. (1980). "Apatoplaca - Genus novum Teloschistacearum (Lichenes)" [Apatoplaca – A new genus of Teloschistaceae (lichens)]. Mitteilungen aus der Botanischen Staatssammlung München (in German). 16: 503–528 [506].
  81. ^ Arup, Søchting & Frödén 2013, p. 68.
  82. ^ a b Fries, Theodore Magnus (1860). Lichenes Arctoi Europae Groenlandiaeque hactenus cogniti [Lichens of Arctic Europe and Greenland hitherto known] (in Latin). Uppsala: Leffler. pp. 166, 218. Retrieved 8 September 2023.
  83. .
  84. ^ Kondratyuk et al. 2014, p. 96.
  85. ^ a b Kondratyuk et al. 2014, p. 97.
  86. ^ .
  87. ^ a b Massalongo, A. (1853). Alcuni generi di licheni nuovamente limitati e descritti [Some genera of lichens newly delimited and described] (in Latin). Verona: Antonelli. pp. 10–11.
  88. ^ Kondratyuk et al. 2017, p. 90.
  89. ^ Massalongo, A.B. (1852). Ricerche sull'autonomia dei licheni crostosi [Research on the autonomy of crustose lichens]. Verona: Dalla tipografia di A. Frizierio. p. 17.
  90. ^ a b Wilk et al. 2021, p. 285.
  91. ^ Kondratyuk et al. 2017, p. 91.
  92. ^ Kondratyuk et al. 2014, p. 102.
  93. ^ Poelt, J. (1977). "Ioplaca gen. nov. Teloschistacearum (Flechten des Himalaya 16)" [Ioplaca, a new genus of the Teloschistaceae (Lichens of the Himalayas 16)]. Khumbu Himal, Ergebnisse des Forschungsunternehemens Nepal Himalaya (in German). 6: 443–446.
  94. ^
    PMID 26539034
    .
  95. ^ Kondratyuk et al. 2017, p. 107.
  96. ^ Trevisan, Vittore (1857). "Nuovi studi sui licheni spettanti alle tribù delle Patellariee, Baeomycee e Lecideinee" [New studies on lichens pertaining to the tribes Patellariaceae, Baeomycetaceae, and Lecideaceae]. Rivista Periodica dei Lavori della Imperiale Regia Accademia di Padova (in Italian). 5: 63–79 [72].
  97. ^ Bungartz, Søchting & Arup 2020, p. 534.
  98. ^ Kondratyuk et al. 2017, p. 108.
  99. ^ Kondratyuk et al. 2020, p. 116.
  100. ^ Hue, A. (1887). "Addenda nova ad Lichenographiam europaeam. Exposuit in Flora Ratisbonensi Dr. W. Nylander, in ordine vero systematico deposuit. Pars II" [New additions to the European lichenography. Presented in the Regensburg Flora by Dr. W. Nylander, arranged in systematic order. Part II]. Revue de Botanique (in Latin). 6: 5–192 [148]. Retrieved 23 August 2023.
  101. ^ Kondratyuk et al. 2014, p. 103.
  102. ^ Kondratyuk et al. 2014, p. 104.
  103. .
  104. ^ .
  105. ^ Bungartz, Søchting & Arup 2020, p. 540.
  106. ^ .
  107. ^ Kondratyuk et al. 2017, p. 112.
  108. ^ Kondratyuk et al. 2017, p. 113.
  109. ^ Kondratyuk et al. 2020, p. 118.
  110. ^ Massalongo, A.B. (1852). "Monografia dei licheni blasteniospori" [Monograph of blasteniospore lichens]. Atti dell'Istituto Veneto Scienze. 2 (in Italian) (3): 5–131 [119].
  111. ^ Arup, Søchting & Frödén 2013, p. 74.
  112. .
  113. ^ Bungartz, Søchting & Arup 2020, p. 548.
  114. ^ a b Arup, Søchting & Frödén 2013, p. 75.
  115. ^ Kondratyuk et al. 2017, p. 118.
  116. ^ Kondratyuk et al. 2014, p. 105.
  117. ^ Wilk et al. 2021, pp. 292–293.
  118. ^ Kondratyuk et al. 2013, p. 265.
  119. ^ Wilk et al. 2021, pp. 293–294.
  120. ^ a b Kondratyuk et al. 2017, p. 86.
  121. ^ Kondratyuk et al. 2013, p. 267.
  122. ^ Dodge, C.W. (1966). "New lichens from Chile". Nova Hedwigia. 12: 307–352 [334].
  123. ^ Kondratyuk et al. 2013, p. 268.
  124. ^ a b c d Wilk et al. 2021, p. 284.
  125. ^ Arup, Søchting & Frödén 2013, p. 61.
  126. ^ Kondratyuk et al. 2017, p. 96.
  127. ^ .
  128. ^ Kondratyuk et al. 2017, p. 105.
  129. .
  130. ^ .
  131. ^ Kondratyuk et al. 2013, p. 269.
  132. ^ Kondratyuk et al. 2017, p. 109.
  133. .
  134. ^ Kondratyuk et al. 2015, p. 330.
  135. ^ Kondratyuk et al. 2017, p. 111.
  136. ^ Massalongo, A. (1861). "Lichenes capenses quos collegit in itinere 1853–1856 Dr H. Wavra, a Dott. A. Massalongo delineati ac descripti" [Lichens of the Cape that were collected during the journey of 1853–1856 by Dr. H. Wavra, depicted and described by Dr. A. Massalongo]. Memorie del Reale Istituto Veneto di Scienze, Lettere ed Arti. 10: 84.
  137. ^ Kondratyuk et al. 2015b, p. 331.
  138. ^ a b c Wilk et al. 2021, p. 287.
  139. ^ Arup, Søchting & Frödén 2013, p. 62.
  140. ^ Arup, Søchting & Frödén 2013, p. 63.
  141. ^ Arup, Søchting & Frödén 2013, p. 64.
  142. ^ Kondratyuk et al. 2015b, p. 334.
  143. ^ Kondratyuk et al. 2015b, p. 335.
  144. ^ Wilk et al. 2021, pp. 279, 284, 287, 294.
  145. ^ Kondratyuk, Sergey Y.; Kärnefelt, Ingvar; Thell, Arne; Elix, John A.; Kim, Jung; Kondratiuk, Anna S.; Hur, Jae-Seoun (2015). "Tassiloa, a new genus in the Teloschistaceae (lichenized ascomycetes)". Graphis Scripta. 27 (1–2): 22–26.
  146. ^ Wilk et al. 2021, pp. 279, 287, 294.
  147. ^ Arup, Søchting & Frödén 2013, p. 65.
  148. ^ Kondratyuk et al. 2015b, p. 337.
  149. ^ a b Arup, Søchting & Frödén 2013, p. 66.
  150. ^ .
  151. ^ .
  152. ^ Arup, Søchting & Frödén 2013, p. 37.
  153. ^ Arup, Søchting & Frödén 2013, p. 38.
  154. ^ a b Arup, Søchting & Frödén 2013, p. 40.
  155. ^ .
  156. ^ Kondratyuk et al. 2017, p. 79.
  157. ^ Luyken, J.A. (1809). Tentamen historiae lichenum [A Study of Lichen History] (in Latin). Gottingen. p. 93.
  158. ^ Kondratyuk et al. 2020, p. 112.
  159. ^ Arup, Søchting & Frödén 2013, p. 44.
  160. ^ Kondratyuk et al. 2017, p. 87.
  161. ^ .
  162. ^ a b Kondratyuk et al. 2014, p. 154.
  163. ^ Arup, Søchting & Frödén 2013, p. 31.
  164. ^ Kondratyuk et al. 2017, p. 101.
  165. ^ Bungartz, Søchting & Arup 2020, p. 527.
  166. ^ .
  167. ^ Kondratyuk et al. 2014, p. 155.
  168. ^ Kondratyuk et al. 2015b, p. 329.
  169. ^ Arup, Søchting & Frödén 2013, p. 47.
  170. ^ a b Kondratyuk, S.Y.; Kärnefelt, I. (2003). "Revision of three natural groups of xanthorioid lichens (Teloschistaceae, Ascomycota)". Ukrainskiy Botanichnyi Zhurnal. 60 (4): 427–437.
  171. ^ Arup, Søchting & Frödén 2013, p. 48.
  172. ^ Arup, Søchting & Frödén 2013, p. 49.
  173. ^ Hue, A. (1908). "Quator lichenum exoticorum" [Four exotic lichens]. Bulletin de la Société linnéenne de Normandie. 6 (in Latin). 1: 75. Retrieved 23 August 2023.
  174. ^ Kondratyuk et al. 2014, p. 156.
  175. ^ a b c Arup, Søchting & Frödén 2013, p. 55.
  176. ^ Kondratyuk et al. 2017, p. 115.
  177. ^ Kondratyuk et al. 2017, p. 117.
  178. .
  179. ^ Kondratyuk et al. 2014, p. 157.
  180. JSTOR 3666548
    .
  181. ^ Kondratyuk, Sergij Y.; Kärnefelt, Ingvar; Elix, John A.; Thell, Arne (2008). "A new circumscription of the genus Xanthodactylon (Teloschistaceae, lichenized ascomycetes)". Sauteria. 15: 265–282.
  182. ^ Lücking, Hodkinson & Leavitt 2017, p. 392.
  183. ^ Santesson, Rolf (1949). "Dolichocarpus and Xanthopeltis, two new lichen genera from Chile". Svensk Botanisk Tidskrift. 43 (2–3): 547–567.
  184. .
  185. ^ Arup, Søchting & Frödén 2013, p. 35.
  186. .
  187. .
  188. ^ Fraser, Shannon J.; Bowman, E.A.; Gianopulos, Nikolas G.; Newcombe, George (2016). "Xanthoria parietina in the Inland Pacific Northwest". North American Fungi. 11 (2): 1–12.
  189. .
  190. .
  191. .
  192. .
  193. .
  194. .
  195. ^ .
  196. .
  197. ^ Bungartz, Søchting & Arup 2020.
  198. .
  199. ^ de Lange, Peter; Dan, Blanchon; Knight, Allison; Elix, John; Lücking, Robert; Frogley, Kelly; Harris, Anna; Cooper, Jerry; Rolfe, Jeremy (2018). Conservation status of New Zealand indigenous lichens and lichenicolous fungi, 2018 (PDF) (Report). New Zealand Threat Classification Series 27. Wellington: Department of Conservation. passim. Archived (PDF) from the original on 8 September 2023. Retrieved 23 August 2023.
  200. PMID 36572164
    .
  201. .
  202. .
  203. .
  204. ^ .
  205. .
  206. .
  207. ^ Ravera, S. (2017). "Caloplaca rinodinae-albae". IUCN Red List of Threatened Species. 2017. Retrieved 23 August 2023.
  208. ^ Sokoloff, P.; McMullin, T. (2020). "Seirophora aurantiaca". IUCN Red List of Threatened Species. 2020. Retrieved 23 August 2023.
  209. ^ Ramos, D.; Vargas, R.; Herrera-Campos, M.; Moat, J.; Whaley, O.; Parrinello, C.; Bungartz, F. (2021). "Teloschistes peruensis". IUCN Red List of Threatened Species. 2021. Retrieved 23 August 2023.
  210. .
  211. ^ Bungartz, Søchting & Arup 2020, p. 516.
  212. .

Cited literature