Jump to content

Teloschistaceae: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
→‎Conservation: ce & links
→‎Description: copyedit per FAC suggestions
(One intermediate revision by the same user not shown)
Line 301: Line 301:


==Description==
==Description==
In general, Teloschistaceae members are known for their vibrant colours, spanning a spectrum of yellow, orange, and red hues, attributed to anthraquinone pigments.<ref name="Cannon & Kirk 2007"/> This group of lichens demonstrates a broad range of physical forms — from the thin, encrusting ([[crustose lichen|crustose]]) to leaf-like ([[foliose lichen|foliose]]), or even bushy ([[fruticose lichen|fruticose]]) formations.<ref name="Cannon & Kirk 2007"/><ref name="Frey 2016"/> Although it is an atypical growth form for the Teloschistaceae, members of genus ''[[Ioplaca]]'' are somewhat {{lichengloss|umbilicate}}, meaning they have a somewhat circular, leafy thallus attached to the substrate at a single point.<ref name="Kondratyuk et al. 2018b"/>
In general, Teloschistaceae members are known for their vibrant colours, spanning a spectrum of yellow, orange, and red hues, attributed to [[anthraquinone]] pigments.<ref name="Cannon & Kirk 2007"/> This group of lichens demonstrates a broad range of physical forms — from thin, encrusting ([[crustose lichen|crustose]]) to leaf-like ([[foliose lichen|foliose]]) or even bushy ([[fruticose lichen|fruticose]]) formations.<ref name="Cannon & Kirk 2007"/><ref name="Frey 2016"/> Although it is an atypical growth form for the Teloschistaceae, members of genus ''[[Ioplaca]]'' are somewhat {{lichengloss|umbilicate}}, meaning they have a somewhat circular, leafy thallus attached to the substrate at a single point.<ref name="Kondratyuk et al. 2018b"/>


Teloschistaceae lichens have a symbiotic relationship with a {{lichengloss|photobiont}}, generally a member of the [[green alga]]l genus ''[[Trebouxia]]''.<ref name="Frey 2016"/> Their reproductive structures, or [[ascomata]], are usually brightly coloured. In most species, {{lichengloss|apothecia}}te ascomata have a {{lichengloss|lecanorine}} form, in which the apothecial {{lichengloss|disc}} is surrounded by a pale rim of tissue known as a {{lichengloss|thalline margin}}. Fewer Teloschistaceae species have {{lichengloss|biatorine}} or {{lichengloss|lecideine}} forms, in which the apothecial disc lacks a thalline margin.<ref name="Frey 2016"/><ref name="Cannon & Kirk 2007"/> Reproductive [[propagule]]s, such as [[isidia]] and [[soredia]], can be found in select species.<ref name="Cannon & Kirk 2007"/>
Teloschistaceae lichens have a symbiotic relationship with a {{lichengloss|photobiont}}, generally a member of the [[green alga]]l genus ''[[Trebouxia]]''.<ref name="Frey 2016"/> The lichen's reproductive structures, or [[ascomata]], are usually brightly coloured, and typically in the form of an apothecium–a wide, open, saucer-shaped or cup-shaped fruit body. In most species, these {{lichengloss|apotheciate}} ascomata have a {{lichengloss|lecanorine}} form, in which the apothecial {{lichengloss|disc}} is surrounded by a pale rim of tissue known as a {{lichengloss|thalline margin}}. Fewer Teloschistaceae species have {{lichengloss|biatorine}} or {{lichengloss|lecideine}} forms, in which the apothecial disc lacks a thalline margin.<ref name="Frey 2016"/><ref name="Cannon & Kirk 2007"/> Reproductive [[propagule]]s, such as [[isidia]] and [[soredia]], can be found in select species.<ref name="Cannon & Kirk 2007"/>


The ascomata encase {{lichengloss|asci}}, cylindrical formations that commonly contain between four to sixteen {{lichengloss|ascospores}}, with eight being the most prevalent count. These asci are characterised by a well-developed J+ layer [[amyloid (mycology)|amyloid]] cap and a rudimentary internal apical apparatus.<ref name="Frey 2016"/> (The term "J+" refers to the positive 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 translucent ([[hyaline]]) ascospores typically feature between one and three internal partitions called [[septum|septa]], marked by a robust central septum and a canal connecting the internal cavities, or {{lichengloss|lumina}}.<ref name="Frey 2016"/><ref name="Cannon & Kirk 2007"/> Despite the polarilocular nature of ascospores suggesting Teloschistaceae lineage, these spores are often not overtly distinctive.<ref name="CRC HoL"/> Although polarilocular ascospores were formerly considered to be a distinguishing feature of the Teloschistaceae, the addition of genera like ''Apatoplaca'', ''Cephalophysis'', ''Fulgensia'', and ''Xanthopeltis'', all of which have simple or septate spores, prompted a reevaluation of the primary characteristics defining this family.<ref name="Gaya et al. 2008"/>
The ascomata encase {{lichengloss|asci}}, cylindrical formations that commonly contain between four to sixteen {{lichengloss|ascospores}}, with eight being the most prevalent count. These asci are characterised by a well-developed J+ layer [[amyloid (mycology)|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 considered rudimentary, relative to the more complex apical structures that occurs in other related families.<ref name="Frey 2016"/>


Teloschistaceae ascospores, typically [[hyaline|translucent]], usually have one to three septa – internal partitions – highlighted by a prominent central septum connected by a canal to the {{lichengloss|lumina}}, the spores' internal cavities.<ref name="Frey 2016"/><ref name="Cannon & Kirk 2007"/> While 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.<ref name="CRC HoL"/> Historically, polarilocular ascospores were regarded as a hallmark of the Teloschistaceae. However, 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 characterizes this group.<ref name="Gaya et al. 2008"/>
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.<ref name="Cannon & Kirk 2007"/> Within this family, [[asexual reproduction]] leads to the creation of {{lichengloss|pycnidia}}-type [[conidiomata]], producing clear {{lichengloss|conidia}} that can be either {{lichengloss|bacillar}} (rod-shaped) or {{lichengloss|bifusiform}} (double-spindle shaped).<ref name="Cannon & Kirk 2007"/><ref name="Frey 2016"/> The tissue composition of the thallus and apothecia is characterised by a loosely {{lichengloss|paraplectenchymatous}} structure, wherein the constituent fungal hyphae are oriented in various directions.<ref name="Ahmadjian & Hale 1973"/>

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.<ref name="Cannon & Kirk 2007"/> Within this family, [[asexual reproduction]] leads to the creation of {{lichengloss|pycnidia}}-type [[conidiomata]] (i.e., small, flask-shaped fruiting bodies), producing clear asexual spores ({{lichengloss|conidia}}) that are either {{lichengloss|bacillar}} (rod-shaped) or {{lichengloss|bifusiform}} (double-spindle shaped).<ref name="Cannon & Kirk 2007"/><ref name="Frey 2016"/> The tissue composition of the thallus and apothecia is characterised by a loosely {{lichengloss|paraplectenchymatous}} structure, wherein the constituent fungal hyphae are oriented in various directions.<ref name="Ahmadjian & Hale 1973"/>


{| class="center toccolours"
{| class="center toccolours"

Revision as of 00:31, 9 November 2023

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 family of mostly lichen-forming fungi belonging to the class Lecanoromycetes in the division Ascomycota. Many members of the Teloschistaceae are readily identifiable by their vibrant orange to yellow hue, a result of their frequent anthraquinone content. The presence of these anthraquinone pigments, which confer protection from ultraviolet light, enabled this group to expand from shaded forest habitats to harsher environmental conditions of sunny and arid ecosystems during the Late Cretaceous. Collectively, the family has a cosmopolitan distribution, although members occur predominantly in subtropical and temperate regions. Although most members are lichens that either live on rock or on bark, about 40 species are lichenicolous fungi–meaning they live on other lichens.

Teloschistaceae lichens typically have one of a few physical growth forms. Depending on the species, the thallus (the main body of the lichen) is either leaf-like (foliose), bushy or shrub-like (fruticose) or crust-like (crustose). These lichens typically partner with a photosynthetic companion (a photobiont) from the green algal genus Trebouxia or similar genera. 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, contributing to the lichen's overall structure and appearance. In Teloschistaceae, the tip of the ascus, the structure responsible for producing spores, characteristically turns blue when treated with iodine. The ascospores are released through a longitudinal slit in the tip of the ascus, a unique trait common to this group of lichens.

The family, first formally proposed in 1898, was extensively revised in 2013, including the recognition of three distinct subfamilies (Caloplacoideae, Teloschistoideae, and Xanthorioideae), and the creation or resurrection of 31 genera. 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. Current estimates suggest that the family contains about 840 species in up to 117 genera. Three species from the Teloschistaceae have been globally assessed for conservation status, with others appearing on regional lists, such as the rare New Zealand species Caloplaca allanii. The full diversity of this family remains underexplored in vast regions like South America and China.

Systematics

Historical taxonomy

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, and the latter in his 1771 work Mantissa Plantarum II.[2]

In his 1852 work Synopsis Lichenum Blasteniosporum ("Synopsis of Lichen Blasteniospores"),[3] Italian 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 perforation. Although Massalongo's efforts to arrange these taxa into more natural genera were largely ignored by later subsequent researchers, several of his proposed genera were resurrected for use 16 decades later, such as Blastenia, Gyalolechia, Pyrenodesmia, and Xanthocarpia.[4]

An ascus of the crustose lichen Pyrenodesmia variabilis, containing eight polarilocular ascospores.

The family Teloschistaceae was formally circumscribed by lichenologist Alexander Zahlbruckner in 1898. In his initial version, he grouped together foliose and fruticose taxa having polarilocular (i.e. two-locule) or four-locule ascospores, including the genera Xanthoria, Teloschistes, and Lethariopsis.[5] At that time, the growth form of the lichen thalli was often used in classical lichen taxonomy to segregate groups of species into families,[6] 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.[7] However, the distinctness of the family Caloplacaceae was largely rejected by other authors,[8] and it is now a historical synonym of Teloschistaceae.[1] In another older classification, crustose genera were grouped together in the family Blasteniaceae[9] or the Placodiaceae.[10][note 1] In 1971, Carroll William Dodge proposed the family Xanthoriaceae to contain Xanthodactylon, Xanthopeltis and Xanthoria,[13] but it was not validly published.[8]

In the 20th century, particularly with the widespread use of electron microscopy, the details of ascus structure became quite important considerations in the taxonomy of lichen-forming fungi.[14] Several studies on various Teloschistaceae species have noted the consistent presence of a cap-like zone at the tip of the ascus that shows a strong reaction to iodine, characteristic of amyloid substances.[15][16][17] Using advanced transmission electron microscopy, Rosmarie Honegger confirmed a unique type of ascus in Teloschistaceae, later named the Teloschistes-type. This ascus is distinguished by a special outer layer that reacts to certain stains and lacks the typical structures seen at the tip, opening in an unusual pattern during spore release.[18] The presence of this ascus type was later used as a diagnostic character for the family Teloschistaceae following an ultrastructural study that corroborated her work.[19] In 1989, Ingvar Kärnefelt revised the family, accepting ten genera,[20] and this served as the main taxonomic classification for the family until the molecular era.[21] In one of the last classifications of the family prior to the widespread use and implementation of molecular techniques, Ove Eriksson, in his popular Outline of the Ascomycota series, accepted 12 genera in Teloschistaceae in 2006: Caloplaca, Cephalophysis, Fulgensia, Huea, Ioplaca, Josefpoeltia, Seirophora, Teloschistes, Xanthodactylon, Xanthomendoza, Xanthopeltis, and Xanthoria.[22] The family continues to undergo significant changes. For example, in 2020, of all fungal families, Teloschistaceae had the fourth-highest number of new fungal names (a total of 128), including 8 genera, 48 new species and infraspecific[note 2] taxa, and 72 new combinations.[23]

Etymology

As is standard practice in botanical nomenclature,[24] 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 Norwegian botanist Johannes M. Norman in 1852,[25] comprises two 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.[26]

Subfamilial and ordinal classification

Teloschistales
Teloschistaceae
Brigantiaeaceae
Cladogram showing the 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).[27] Species names have been updated to reflect current taxonomy. Single quote marks around a genus name suggest that the species is likely misclassified.

Teloschistaceae is divided into three recognised subfamilies: Xanthorioideae, Caloplacoideae, and Teloschistoideae.[28] In 2015, researchers proposed a fourth subfamily, Brownlielloideae,[29] which was later shown by genetic studies to be a grouping based on mixed or misinterpreted data rather than a distinct lineage.[30][31] 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.[32] DNA evidence also dispersed members of the informally introduced subfamily Ikaerioideae across the three acknowledged subfamilies, primarily within Teloschistoideae.[33] Despite this, Sergey Kondratyuk and his colleagues continue to use Brownlielloideae and Ikaerioideae in their publications, assigning nine genera to the former and two to the latter.[34] 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.[35] These groups are genetically distinct, with each subfamily showing unique patterns in their nuclear large ribosomal subunit RNA sequences.[36]

  • 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.[37]
  • 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.[38]
  • 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.[28]

The order Teloschistales was first proposed by David Hawksworth and Eriksson in 1986, with a single family (Teloschistaceae); other families were added later.[39] In the 1990s, several authors recognised the Teloschistales as a suborder within the Lecanorales;[40] as a suborder it was named Teloschistineae.[41] Following the appearance of preliminary molecular studies,[42] the Teloschistaceae was classified by some within the order Lecanorales, although others maintained the Teloschistales as a valid order.[43] The large-scale, multigene phylogenetic study of the class Lecanoromycetes by Jolanta Miądlikowska and colleagues 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).[44] The suborder Teloschistineae was formally proposed by Ester Gaya and François Lutzoni in 2015.[45]

Molecular phylogenetics

Molecular phylogenetics has revolutionised our understanding of the Teloschistaceae. Historically, classification 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. Molecular studies, however, have shown that these features can be inconsistent, leading to unreliable evolutionary interpretations.[46]

The development of advanced DNA analysis techniques has allowed scientists to identify and differentiate cryptic species—those that are visually similar but genetically diverse. This technological progression has led to the recognition of distinct species within groups formerly considered homogeneous, such as within the genus Caloplaca, by revealing their unique genetic signatures.[47]

While Teloschistaceae is 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.[34][48][49] Over time, as more sequences became available, researchers like Ester Gaya and colleagues in 2012 began to develop a clearer understanding of the family's phylogeny.[21]

One significant finding from molecular data is that the traditional morphological methods had mistakenly grouped different species together. For example, the genus Caloplaca was once thought to be descended from a single lineage (i.e. monophyletic), but is now understood to have been composed of multiple, unrelated groups (polyphyletic). This insight prompted suggestions to redefine the genus into smaller, monophyletic groups,[50] but such changes have met with resistance due to the vast number of species reclassifications they would entail.[51]

Moreover, molecular evidence has helped to map the family's relationship within the class Lecanoromycetes. A 2018 study, for instance, identified the Megalosporaceae as the Teloschistaceae's closest relative.[52]

In response to these discoveries, experts like Robert Lücking recommend extensive analysis using multiple genetic markers to accurately determine the family's evolutionary lineage. Such comprehensive studies are important for revising the taxonomic classification system of this diverse and widespread group of lichens.[45]

Description

In general, Teloschistaceae members are known for their vibrant colours, spanning a spectrum of yellow, orange, and red hues, attributed to anthraquinone pigments.[1] This group of lichens demonstrates a broad range of physical forms — from thin, encrusting (crustose) to leaf-like (foliose) or even bushy (fruticose) formations.[1][53] 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.[54]

Teloschistaceae lichens have a symbiotic relationship with a photobiont, generally a member of the green algal genus Trebouxia.[53] The lichen's reproductive structures, or ascomata, are usually brightly coloured, and typically in the form of an apothecium–a wide, open, saucer-shaped or cup-shaped fruit body. In most species, these apotheciate ascomata have a lecanorine form, in which the apothecial disc is surrounded by a pale rim of tissue known as a thalline margin. Fewer Teloschistaceae species have biatorine or lecideine forms, in which the apothecial disc lacks a thalline margin.[53][1] Reproductive propagules, such as isidia and soredia, can be found in select species.[1]

The ascomata encase asci, cylindrical formations that commonly contain between four to 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 considered rudimentary, relative to the more complex apical structures that occurs in other related families.[53]

Teloschistaceae ascospores, typically translucent, usually have one to three septa – internal partitions – highlighted by a prominent central septum connected by a canal to the lumina, the spores' internal cavities.[53][1] While 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.[55] Historically, polarilocular ascospores were regarded as a hallmark of the Teloschistaceae. However, 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 characterizes this group.[8]

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 clear asexual spores (conidia) that are either bacillar (rod-shaped) or bifusiform (double-spindle shaped).[1][53] The tissue composition of the thallus and apothecia is characterised by a loosely paraplectenchymatous structure, wherein the constituent fungal hyphae are oriented in various directions.[14]

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 primarily associate with trebouxioid green algal photobionts. An early study investigating the ultrastructure of the interaction between the fungus and alga in various 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.[56] The widespread Xanthoria parietina species complex has been identified to be in association with various Trebouxia species, including T. arboricola, T. decolorans, and T. italiana.[57] 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.[44]

Research on Teloschistaceae photobionts has shown that all studied foliose (Xanthoria, Xanthomendoza) and fruticose (Teloschistes) types were affiliated with specific Trebouxia clades. This indicates a degree of specificity at the genus level, where only certain subclades of the Trebouxia clade are seen as suitable partners. This specificity, however, can vary based on the habitat; in extreme climates, lichens might be associated with a broader range of photobionts.[57]

Chemistry

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

The main group of lichen products that occur in the Teloschistaceae are chemical pigments called anthraquinones. These substances, which are deposited in the upper cortical layer of the lichen,[58] have photoprotective properties,[59] as they can absorb ultraviolet (UV) and blue light.[58] Evolutionary innovations in secondary metabolite production allowed the family to broaden its geographical range and transition from shaded, plant-based habitats to sun-exposed, arid environments. The production of protective chemicals is thought to be a direct contributor to the evolutionary success of the familial lineage. A 2023 study reported using comparative genomics to identify a metabolic gene cluster involved in anthraquinone metabolism and shared uniquely across the Teloschistales. Phylogenetic analyses of fungal polyketide synthases (PKSs) reveal a consistent grouping, hinting at a shared ancestral trait for anthraquinone biosynthesis within the subphylum Pezizomycotina. While the genetic machinery (like the PKSs) involved in anthraquinone biosynthesis in Teloschistales and some non-lichenised fungi is conserved and shows similarities, the specific arrangement of the involved enzymes seems to be a distinguishing feature in the Teloschistales' approach to anthraquinone biosynthesis. The discovery of an ABC transporter gene within the pigment gene cluster provides clues as to how the lichens are able to accumulate large amounts of potentially toxic anthraquinone crystals in their thallus and reproductive structures.[58]

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.[60] Parietin is an antioxidant molecule that is produced in greater amounts (upregulated) in lichen thalli that are exposed to excess nitrogen.[61] 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".[60] 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, teloschistin, fallacinal, and parietinic acid as the main substances.[62] Parietin acts as a UV-light filter to provide optimal light intensities for the photobionts that are resident in the internal algal layer. Investigations into the parietin concentration in Xanthoria parietina across a light gradient demonstrate a direct relationship between light intensity and concentration. In the Teloschistaceae, parietin may serve an additional defensive role. In the Negev desert, the parietin-containing Teloschistaceae species Elenkiniana ehrenbergii and Seirophora lacunosa are avoided by grazing snails, while they frequently consume species like Diploicia canescens and Buellia subalbula (both in family Caliciaceae), which lack parietin.[63]

In their large-scale phylogenetic analysis of the Teloschistaceae, Arup and colleagues analysed about 4000 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).[28] For example, the genus Catenaria, which contains three South American species, is characterised by the presence of 7-chlorocatenarin, a secondary metabolite previously unknown in lichens.[64] Similarly, the substance usnic acid characterises the genus Usnochroma, while 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.[28]

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.[65]

Although most Teloschistaceae lichens produce anthraquinone pigments in shades ranging from yellow to orange to red, the genera Apatoplaca and Cephalophysis lack these anthraquinones. Similarly, the genus Pyrenodesmia encompasses species where anthraquinones are absent and replaced by substances such as Cinereorufa-green or Sedifolia-grey; these insoluble lichen pigments may confer UV-protective ability similar to anthraquinones. Taxa of the closely related genera Kuettlingeria and Sanguineodiscus have anthraquinones in their apothecia and Sedifolia-grey in their thalli.[66] The species Kuettlingeria neotaurica features apothecia of two colour variants: orange-red (with anthraquinones) and grey (with Sedifolia-grey). The absence of anthraquinones is not a synapomorphic character, but appears independently in unrelated lineages of Teloschistaceae; as such, it is a phylogenetically unreliable character.[67]

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 shifted from shaded, bark-dwelling habitats to sunlit, rocky areas during their diversification.[68]

The analysis of phenotypic traits and diversification rates reveals 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.[68] 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.[58]

Genera

This section presents a compilation of the genera in the Teloschistaceae, based largely on a 2021 fungal classification review and new reports published since then.[69] Each genus is paired with its taxonomic authority, denoting the first describers using standardised author abbreviations, the publication year, and the number of species.

Contemporary estimates of the number of Teloschistaceae taxa include: 10 genera and 47 species (2001),[43] 12 genera and 644 species (2008);[70] 51–53 genera and about 700 species (2016);[53] 65 genera and 755 species (2017);[45] and 71 genera and about 840 species (2022).[69] Also in 2022, Kondratyuk and colleagues enumerated all members of the Teloschistaceae with publicly available DNA sequences, and confirmed 590 species in 115 genera.[34] As of November 2023, 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.[71][note 3]

In terms of species diversity, Teloschistaceae stood as the sixth-largest lichen-forming fungal family by 2017, following the Parmeliaceae, Graphidaceae, Verrucariaceae, Ramalinaceae, and the Lecanoraceae.[45] Genera are organised here by subfamily:

Caloplacoideae

Examples from subfamily Caloplacoideae (clockwise from upper left): Blastenia ferruginea; Caloplaca maculata; 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): Brownliella cinnabarina; Niorma hosseusiana; Wetmoreana brouardii; and Fulgogasparrea appressa

Xanthorioideae

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

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

Habitat, distribution, and ecology

The scrambled egg lichen, Gyalolechia fulgens, is a terricolous species and a component of some biological soil crusts.
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.

Collectively, the family has a cosmopolitan distribution, although members occur predominantly in subtropical and temperate regions. Most members either grow on rock (saxicolous) or on bark (corticolous).[53] As an exception to this general ecological preference, genus Bryoplaca contains species that only grow on mosses and detritus.[28] Some Fulgensia species grow on soils (terricolous), particularly those rich in lime.[163] 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 South America.[164]

In general, the family is moderately to strongly nitrophilous. This suggests a preference of many of its species for habitats that are rich in nitrogen, particularly in the form of nitrate.[14] Sun-adapted lichens, such as the Teloschistaceae, have an enhanced ability to upregulate the levels at which they fix carbon from the atmosphere and absorb excess nitrogen. Small foliose and crustose lichens are in general more tolerant to higher levels of nitrogen.[61] Xanthoria parietina is one example of a widespread lichen that appears to be experiencing an increase in its range due to its ability to tolerate nitrogenous pollutants, and its potential ability to displace native lichen species as a result.[165] Caloplaca, Fulgensia, Teloschistes, and Xanthoria are genera that are characteristic of sun-exposed habitats; in some extreme desert environments, Caloplaca (in the broad sense) may be the only genus present, while Caloplaca and Xanthoria dominate harsh coastal environments.[63]

There are several Teloschistaceae genera that contain lichenicolous (lichen-dwelling) species. These 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.).[166] 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.[167]

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.[168] Examples include Xanthomendoza borealis, Austroplaca soropelta, and Caloplaca phlogina.[169] Conversely, there is a relatively low diversity of crustose Teloschistaceae in Central Europe. Localised exceptions occur in 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.[170] Some Teloschistaceae genera have a strong geographic centre of species richness; examples include Elixjohnia (Australasia),[105] Orientophila (east Asia), Shackletonia (Antarctic and subantarctic), Stellarangia (south-western Africa) and Xanthoria (Mediterranean area).[28]

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

Species interactions

Teloschistaceae species are known to be host to many lichenicolous fungi, with certain fungi like Cercidospora caudata and Stigmidium cerinae displaying a broad range of hosts within this family. However, most fungi show a preference for specific Teloschistaceae species or genus. The relationship between Teloschistaceae lichens and the fungus Tremella caloplacae is particularly notable. 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 officially described in 2023, each uniquely adapted to a single host species or genus within the Teloschistaceae.[177][178]

Human interactions and uses

Economic significance

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

There are no species in the Teloschistaceae that have any major economic significance.[1] The ability of some members to grow on rock surfaces, however, has led to several recorded instances where Teloschistaceae species have damaged marble surfaces. 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.[179] Caloplaca pseudopoliotera and C. cupulifera are two crustose species implicated in the slow degradation of the Konark Sun Temple in India.[180]

Traditional medicine

Some Teloschistaceae species have been utilised in traditional medicine practices across various cultures. Xanthoria parietina is particularly prominent; in Spain, it has been used in wine decoctions for menstrual issues, in water for kidney and tooth ailments, as an analgesic, and as an ingredient in a cold medicine. In Europe during the early modern era, it was boiled in milk to alleviate jaundice—a treatment shared with Polycauliona candelaria—and employed for diarrhea, dysentery, stopping bleeding, as a malaria remedy in lieu of quinine, and for treating hepatitis. In Traditional Chinese medicine the lichen has been used as an antibacterial.[181]

In regional practices, Rusavskia elegans is used for treating wounds; in Afghanistan, it is applied directly, while in Kyrgyzstan, the lichen is mixed with butter and used as a remedy for diarrhea 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.[181]

In science

Xanthoria parietina has been suggested for use as a potential reliable pollutant tolerance biomonitor in urban ecosystems due to its widespread presence and ability to adapt to high pollution levels.[182] 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 showed the lichen to have an impressive ability to survive under these conditions.[183]

Conservation

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.[184] Seirophora aurantiaca (Endangered, 2020) is particularly vulnerable to climate change impacts in the Canadian Arctic, leading to eroding coasts, increased sea ice melt, saline wash from storm surges, permafrost melting, and potential invasive species intrusions.[185] 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 presence of invasive species like goats and cows altering the habitat.[186]

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" using the New Zealand Threat Classification System.[187]

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,[188] 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.[189]

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.[11][12]
  2. ^ The term "infraspecific" refers to a taxonomic rank below that of species, including subspecies, variety, and form.
  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.
  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.[81]
  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.[81]
  6. ^ Kondratyuk and colleagues suggest that Oceanoplaca is a synonym of Loekoeslaszloa.[93]
  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.[109]
  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.[109]
  9. ^ Wilk and colleagues suggest that Raesaeneniana is a putative synonym of Villophora, a genus with which its single species fits well morphologically.[123]
  10. ^ Wilk and colleagues proposed to reduce Tarasginia to synonymy with Sirenophila.[128]
  11. ^ Wilk and colleagues proposed to reduce Tayloriellina to synonymy with Villophora.[130]
  12. ^ Wilk and colleagues suggest that Thelliana is a putative synonym of Filsoniana.[123]
  13. ^ Wilk and colleagues suggest that Dijigiella is a putative synonym of Teuvoahtiana.[123]
  14. ^ a b c d e Kondratyuk and colleagues in 2014 suggested that the genus Dufourea was polyphyletic, and divided it into four genera: Dufourea, Jackelixia, Langeottia, Ovealmbornia, and Xanthokarrooa; Wilk and colleagues prefer to retain a more broadly defined Dufourea pending additional research.[109]
  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.[109]
  16. ^ Bungartz and colleagues synonymise Huriella with Squamulea.[149]

Citations

  1. ^ a b c d e f g h i j Cannon, Paul F.; Kirk, Paul M. (2007). Fungal Families of the World. Wallingford: CAB International. pp. 345–346, 440. ISBN 978-0-85199-827-5. OCLC 60741230.
  2. ^ Jørgensen, Per M. (1994). "Linnaean lichen names and their typification". Botanical Journal of the Linnean Society. 115 (4): 261–405. doi:10.1111/j.1095-8339.1994.tb01784.x.
  3. ^ 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.
  4. ^ a b Nimis, Pier Luigi (2016). The Lichens of Italy. A Second Annotated Catalogue. Trieste: Edizioni Università di Trieste. pp. 7, 24. ISBN 978-88-8303-755-9.
  5. ^ 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.
  6. ^ Döring, Heidi; Lumbsch, H. Thorsten (1998). "Ascoma ontogeny: is this character set of any use in the systematics of lichenized ascomycetes?". The Lichenologist. 30 (4–5): 489–500. doi:10.1006/lich.1998.0144. S2CID 86114296.
  7. ^ 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].
  8. ^ a b c Gaya, Ester; Navarro-Rosinés, Pere; Llimona, Xavier; Hladun, Néstor; Lutzoni, François (2008). "Phylogenetic reassessment of the Teloschistaceae (lichen-forming Ascomycota, Lecanoromycetes)". Mycological Research. 112 (5): 528–546. doi:10.1016/j.mycres.2007.11.005. PMID 18406120.
  9. ^ Dodge, C.W.; Baker, G.E. (1938). "The Second Byrd Antarctic Expedition-Botany. II. Lichens and Lichen parasites". Annals of the Missouri Botanical Garden. 25 (2): 515–727 [604]. doi:10.2307/2394232. JSTOR 2394232.
  10. ^ Räsänen, V. (1943). "Das System der Flechten" [The system of lichens]. Acta Botanica Fennica (in German). 33: 1–82.
  11. ^ "Record Details: Blasteniaceae C.W. Dodge & G.E. Baker, Ann. Mo. bot. Gdn 25(2): 604 (1938)". Index Fungorum. Retrieved 23 August 2023.
  12. ^ "Record Details: Placodiaceae A. Fisch. [as 'Placodinae'], Mitt. berin. naturf. Ges.: 17 (1871)". Index Fungorum. Retrieved 23 August 2023.
  13. ^ Dodge, C. (1971). Some lichens of tropical Africa. V. Lecanoraceae to Physciaceae. Beihefte zur Nova Hedwigia. Vol. 38. p. 115. ISBN 978-3-7682-5438-0.
  14. ^ a b c Ahmadjian, Vernon; Hale, Mason E. (1973). The Lichens. New York: Academic Press. p. 625. ISBN 978-0-12-044950-7.
  15. ^ Richardson, D.H.S (1970). "Ascus and ascocarp structure in lichens". The Lichenologist. 4 (4): 350–361. doi:10.1017/s0024282970000440. S2CID 85251464.
  16. ^ Letrouit-Galinou, Marie-Agnès (1973). "Les asques des lichens et le type archaeascé" [The ascus of lichens and the archaeascid type]. The Bryologist (in French). 76 (1): 30–47. doi:10.2307/3241230. JSTOR 3241230.
  17. ^ Henssen, Aino; Jahns, Hans Martin (1974). Lichenes. Eine Einführung in die Flechtenkunde [Lichens: An Introduction to Lichenology] (in German). Stuttgart: George Thieme Verlag. ISBN 978-3-13-496601-5.
  18. ^ Honegger, R. (1978). "The ascus apex in lichenized fungi I. The Lecanora-, Peltigera- and Teloschistes-types" (PDF). The Lichenologist. 10 (1): 47–67. doi:10.1017/s0024282978000079. S2CID 84629945.
  19. ^ 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, Lichenologie (in French). 7 (3): 189–211.
  20. ^ Kärnefelt, I. (1989). "Morphology and phylogeny in the Teloschistales". Cryptogamic Botany. 1: 147–203.
  21. ^ a b Gaya, Ester; Högnabba, Filip; Holguin, Ángela; Molnar, Katalin; Fernández-Brime, Samantha; Stenroos, Soili; Arup, Ulf; Søchting, Ulrik; Boom, Pieter Van den; Lücking, Robert; Sipman, Harrie J.M.; Lutzoni, François (2012). "Implementing a cumulative supermatrix approach for a comprehensive phylogenetic study of the Teloschistales (Pezizomycotina, Ascomycota)". Molecular Phylogenetics and Evolution. 63 (2): 374–387. doi:10.1016/j.ympev.2012.01.012. PMID 22306043.
  22. ^ Eriksson, O.E. (2006). "Outline of Ascomycota – 2006". Myconet. 12: 1–82.
  23. ^ Wang, Ke; Cai, Lei; Yao, Yijian (2021). "Overview of nomenclature novelties of fungi in the world and China (2020)". Biodiversity Science. 29 (8): 1064–1072. doi:10.17520/biods.2021202. S2CID 240568551.
  24. ^ Hawksworth, D.L. (1974). Mycologist's Handbook. Kew: Commonwealth Mycological Institute. p. 39. ISBN 978-0-85198-300-4.
  25. ^ 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].
  26. ^ Ulloa, Miguel; Aguirre-Acosta, Elvira (2020). Illustrated Generic Names of Fungi. St. Paul, Minnesota: APS press. pp. 362–363. ISBN 978-0-89054-618-5.
  27. ^ Wilk et al. 2021, p. 282.
  28. ^ a b c d e f g Arup, Søchting & Frödén 2013.
  29. ^ Kondratyuk et al. 2015b.
  30. ^ Vondrák, Jan; Shahidin, Hurnisa; Moniri, Mahroo Haji; Halıcı, Gökhan; Košnar, Jiří (2018). "Taxonomic and functional diversity in Calogaya (lichenised Ascomycota) in dry continental Asia". Mycological Progress. 17 (8): 897–916. doi:10.1007/s11557-018-1402-9. S2CID 255312095.
  31. ^ Wilk et al. 2021.
  32. ^ Wilk et al. 2021, p. 278.
  33. ^ Wilk et al. 2021, p. 283.
  34. ^ a b c Kondratyuk, S.Y.; Popova, L.P.; Kondratiuk, A.S.; Lőkös, L. (2022). "The first enumeration of members of the Teloschistaceae (lichen-forming Ascomycetes) status of which confirmed by three gene phylogeny" (PDF). Studia Botanica Hungarica. 53 (2): 137–234. doi:10.17110/studbot.2022.53.2.137. S2CID 256569290.
  35. ^ Arup, Søchting & Frödén 2013, pp. 285–287.
  36. ^ Bungartz, Søchting & Arup 2020, pp. 528, 563.
  37. ^ Bungartz, Søchting & Arup 2020, p. 528.
  38. ^ Arup, Ulf; Søchting, Ulrik; Frödén, Patrik (2013b). "Addendum to 'A new taxonomy of the family Teloschistaceae'". Nordic Journal of Botany. 31 (2): 256. doi:10.1111/j.1756-1051.2013.00295.x.
  39. ^ Hawksworth, D.L.; Eriksson, O.E. (1986). "The names of accepted orders of ascomycetes". Systema Ascomycetum. 5 (2): 175–184.
  40. ^ Hafellner, J.; Hertel, H.; Rambold, G.; Timdal, E. (1994). "Discussion 4: Lecanorales". In Hawksworth, D.L. (ed.). Ascomycete Systematics: problems and perspectives in the nineties. NATO Advanced Science Institutes Series A269. New York: Plenum Press. pp. 379–387. ISBN 978-0-306-44882-9.
  41. ^ Tehler, A.; Wedin, Mats (2008). "Systematics of lichenized fungi". In Nash III, Thomas H. (ed.). Lichen Biology (2nd ed.). Cambridge, UK: Cambridge University Press. p. 349. ISBN 978-0-521-69216-8.
  42. ^ Grube, Martin; Winka, Katarina (2002). "Progress in understanding the evolution and classification of lichenized ascomycetes". Mycologist. 16 (2): 67–76. doi:10.1017/S0269-915X(02)00206-9 (inactive 2 November 2023).{{cite journal}}: CS1 maint: DOI inactive as of November 2023 (link)
  43. ^ a b Kirk, P.M.; Cannon, P.F.; David, J.C.; Stalpers, J.A. (2001). "Teloschistaceae". Ainsworth & Bisby's Dictionary of the Fungi (9th ed.). Oxon, UK: CABI Bioscience. pp. 515–516. ISBN 978-0-85199-377-5.
  44. ^ a b Miadlikowska, Jolanta; Kauff, Frank; Högnabba, Filip; Oliver, Jeffrey C.; Molnár, Katalin; Fraker, Emily; et al. (2014). "A multigene phylogenetic synthesis for the class Lecanoromycetes (Ascomycota): 1307 fungi representing 1139 infrageneric taxa, 317 genera and 66 families". Molecular Phylogenetics and Evolution. 79: 132–168. doi:10.1016/j.ympev.2014.04.003. PMC 4185256. PMID 24747130.
  45. ^ a b c d Lücking, Robert; Hodkinson, Brendan P.; Leavitt, Steven D. (2017). "The 2016 classification of lichenized fungi in the Ascomycota and Basidiomycota–Approaching one thousand genera". The Bryologist. 119 (4): 361–416. doi:10.1639/0007-2745-119.4.361. S2CID 90258634.
  46. ^ Arup, Søchting & Frödén 2013, p. 17.
  47. ^ Frolov, Ivan; Vondrák, Jan; Fernández-Mendoza, Fernando; Wilk, Karina; Khodosovtsev, Alexander; Halıcı, Mehmet Gökhan (2016). "Three new, seemingly-cryptic species in the lichen genus Caloplaca (Teloschistaceae) distinguished in two-phase phenotype evaluation". Annales Botanici Fennici. 53 (3–4): 243–262. doi:10.5735/085.053.0413. hdl:10995/117780. S2CID 88825723.
  48. ^ Lumbsch, H.Thorsten; Schmitt, Imke; Palice, Zdenek; Wiklund, Elisabeth; Ekman, Stefan; Wedin, Mats (2004). "Supraordinal phylogenetic relationships of Lecanoromycetes based on a Bayesian analysis of combined nuclear and mitochondrial sequences". Molecular Phylogenetics and Evolution. 31 (3): 822–832. doi:10.1016/j.ympev.2003.11.001. PMID 15120381.
  49. ^ Peršoh, Derek; Beck, Andreas; Rambold, Gerhard (2004). "The distribution of ascus types and photobiontal selection in Lecanoromycetes (Ascomycota) against the background of a revised SSU nrDNA phylogeny". Mycological Progress. 3 (2): 103–121. doi:10.1007/s11557-006-0081-0. S2CID 495074.
  50. ^ 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.
  51. ^ 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.
  52. ^ Kraichak, Ekaphan; Huang, Jen-Pan; Nelsen, Matthew; Leavitt, Steven D.; Lumbsch, H. Thorsten (2018). "A revised classification of orders and families in the two major subclasses of Lecanoromycetes (Ascomycota) based on a temporal approach". Botanical Journal of the Linnean Society. 188 (3): 233–249. doi:10.1093/botlinnean/boy060. S2CID 92280920.
  53. ^ a b c d e f g h Jaklitsch, Walter; Baral, Hans-Otto; Lücking, Robert; Lumbsch, H. Thorsten (2016). Frey, Wolfgang (ed.). Syllabus of Plant Families: Adolf Engler's Syllabus der Pflanzenfamilien. Vol. 1/2 (13 ed.). Berlin Stuttgart: Gebr. Borntraeger Verlagsbuchhandlung, Borntraeger Science Publishers. pp. 136–137. ISBN 978-3-443-01089-8. OCLC 429208213.
  54. ^ a b Kondratyuk, S.Y.; Persson, P.E.; Hansson, M.; Mishra, G.K.; Nayaka, S.; Liu, D.; Hur, J.S.; Thell, A. (2018). "Upretia, a new caloplacoid lichen genus (Teloschistaceae, lichen-forming Ascomycota) from India". Cryptogam Biodiversity and Assessment. S2018: 22–31. doi:10.21756/CAB.ESP5. S2CID 89792633.
  55. ^ Hafellner, Josef (1988). "Principles of classification and main taxonomic groups". In Galun, Margalith (ed.). CRC Handbook of Lichenology. Vol. III. Boca Raton: CRC Press. p. 42. ISBN 978-0-8493-3583-9.
  56. ^ Galun, Margalith; Ben-Shaul, Y.; Paran, Navah (1971). "Fungus-alga association in lichens of the Teloschistaceae: an ultrastructural study". New Phytologist. 70 (5): 837–839. doi:10.1111/j.1469-8137.1971.tb02584.x.
  57. ^ a b Nyati, Shyam; Scherrer, Sandra; Werth, Silke; Honegger, Rosmarie (2014). "Green-algal photobiont diversity (Trebouxia spp.) in representatives of Teloschistaceae (Lecanoromycetes, lichen-forming ascomycetes)" (PDF). The Lichenologist. 46 (2): 189–212. doi:10.1017/s0024282913000819. S2CID 46909050.
  58. ^ a b c d Llewellyn, Theo; Nowell, Reuben W.; Aptroot, Andre; Temina, Marina; Prescott, Thomas A.K.; Barraclough, Timothy G.; Gaya, Ester (2023). "Metagenomics shines light on the evolution of "sunscreen" pigment metabolism in the Teloschistales (lichen-forming Ascomycota)". Genome Biology and Evolution. 15 (2): evad002. doi:10.1093/gbe/evad002. PMC 9907504. PMID 36634008.
  59. ^ Nguyen, Khanh-Hung; Chollet-Krugler, Marylène; Gouault, Nicolas; Tomasi, Sophie (2013). "UV-protectant metabolites from lichens and their symbiotic partners". Natural Product Reports. 30 (12): 1490–1508. doi:10.1039/c3np70064j. PMID 24170172.
  60. ^ a b Santesson, Johan (1970). "Anthraquinones in Caloplaca". Phytochemistry. 9 (10): 2149–2166. Bibcode:1970PChem...9.2149S. doi:10.1016/s0031-9422(00)85380-7.
  61. ^ a b Delves, Jay; Lewis, Jason E.J.; Ali, Niaz; Asad, Saeed A.; Chatterjee, Sudipto; Crittenden, Peter D.; et al. (2023). "Lichens as spatially transferable bioindicators for monitoring nitrogen pollution". Environmental Pollution. 328: 121575. doi:10.1016/j.envpol.2023.121575. PMID 37028790.
  62. ^ Søchting, Ulrik (1997). "Two major anthraquinone chemosyndromes in Teloschistaceae". In Türk, Roman; Zorer, Roberto (eds.). Progress and Problems in Lichenology in the Nineties – IAL 3. Bibliotheca Lichenologica. Vol. 68. Berlin/Stuttgart: J. Cramer. pp. 135–144. ISBN 978-3-443-58047-6.
  63. ^ a b Rundel, Philip (1978). "The ecological role of secondary lichen substances". Biochemical Systematics and Ecology. 6 (3): 157–170. doi:10.1016/0305-1978(78)90002-9.
  64. ^ a b Søchting, Ulrik; Søgaard, Majbrit Zeuthen; Elix, John A.; Arup, Ulf; Elvebakk, Arve; Sancho, Leopoldo G. (2014). "Catenarina (Teloschistaceae, Ascomycota), a new Southern Hemisphere genus with 7-chlorocatenarin". The Lichenologist. 46 (2): 175–187. doi:10.1017/s002428291300087x. S2CID 83906534.
  65. ^ Sipman, Harrie; Aptroot, André (2020). "Ikaeria serusiauxii, a new Caloplaca-like lichen from Macaronesia and mainland Portugal, with a lichen checklist for Porto Santo". Plant and Fungal Systematics. 65: 120–130. doi:10.35535/pfsyst-2020-0006.
  66. ^ a b Frolov, Ivan; Vondrák, Jan; Košnar, Jiří; Arup, Ulf (2021). "Phylogenetic relationships within Pyrenodesmia sensu lato and the role of pigments in its taxonomic interpretation". Journal of Systematics and Evolution. 59 (3): 454–474. doi:10.1111/jse.12717. S2CID 234535735.
  67. ^ Vondrák, Jan; Šoun, Jaroslav; Vondrákov, Olga; Fryday, Alan M.; Khodosovtsev, Alexander; Davydov, Evgeny A. (2012). "Absence of anthraquinone pigments is paraphyletic and a phylogenetically unreliable character in the Teloschistaceae". The Lichenologist. 44 (3): 401–418. doi:10.1017/s0024282911000843. S2CID 54693944.
  68. ^ a b Gaya, Ester; Fernández-Brime, Samantha; Vargas, Reinaldo; Lachlan, Robert F.; Gueidan, Cécile; Ramírez-Mejía, Martín; Lutzoni, François (2015). "The adaptive radiation of lichen-forming Teloschistaceae is associated with sunscreening pigments and a bark-to-rock substrate shift". Proceedings of the National Academy of Sciences. 112 (37): 11600–11605. Bibcode:2015PNAS..11211600G. doi:10.1073/pnas.1507072112. PMC 4577145. PMID 26324894.
  69. ^ a b Wijayawardene, N.N.; Hyde, K.D.; Dai, D.Q.; Sánchez-García, M.; Goto, B.T.; Saxena, R.K.; et al. (2022). "Outline of Fungi and fungus-like taxa – 2021". Mycosphere. 13 (1): 53–453 [156–158]. doi:10.5943/mycosphere/13/1/2. S2CID 249054641.
  70. ^ Kirk, P.M.; Cannon, P.F.; Minter, D.W.; Stalpers, J.A., eds. (2008). "Teloschistaceae". Dictionary of the Fungi (10th ed.). Wallingford, UK: CAB Europe. p. 680. ISBN 978-0-85199-826-8.
  71. ^ Source dataset. Species Fungorum Plus: Species Fungorum for CoL+. "Teloschistaceae". Catalogue of Life Version: 2023-07-18. Retrieved 23 August 2023.
  72. ^ 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.
  73. ^ Arup, Søchting & Frödén 2013, p. 68.
  74. ^ 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.
  75. ^ Kilias, H. (1985). "Cephalophysis (Hertel) Kilias gen. nov., eine weitere Gattung der Teloschistaceae mit einzelligen Sporen" [Cephalophysis (Hertel) Kilias gen. nov., another genus of the Teloschistaceae family with single-celled spores]. Herzogia. 7 (1–2): 181–190. doi:10.1127/herzogia/7/1985/181. S2CID 249722636.
  76. ^ a b c d e f g Kondratyuk, S.Y.; Jeong, M.-H.; Yu, N.-N.; Kärnefelt, I.; Thell, A.; Elix, J.A.; Kim, J.; Kondratiuk, A.S.; Hur, J.-S. (2014). "A revised taxonomy for the subfamily Caloplacoideae (Teloschistaceae, Ascomycota) based on molecular phylogeny". Acta Botanica Hungarica. 56 (1–2): 93–123. doi:10.1556/abot.56.2014.1-2.10.
  77. ^ a b c Kondratyuk, S.Y.; Lőkös, L.; Kim, J.A.; Kondratiuk, A.S.; Jeong, M.-H.; Jang, S.H.; Oh, S.-O.; Wang, X.Y.; Hur, J.-S. (2016). "Fauriea, a new genus of the lecanoroid caloplacoid lichens (Teloschistaceae, lichen-forming ascomycetes)" (PDF). Acta Botanica Hungarica. 58 (3–4): 303–318. doi:10.1556/abot.58.2016.3-4.6.
  78. ^ 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.
  79. ^ Kondratyuk et al. 2017, p. 90.
  80. ^ 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.
  81. ^ a b Wilk et al. 2021, p. 285.
  82. ^ Kondratyuk et al. 2017, p. 91.
  83. ^ Poelt, J. (1977). "Ioplaca gen. nov. Teloschistacearum (Flechten des Himalaya 16)". Khumbu Himal, Ergebnisse des Forschungsunternehemens Nepal Himalaya (in German). 6: 443–446.
  84. ^ a b c Kondratyuk, S.Y.; Lőkös, L.; J.A., Kim; Kondratiuk, A.S.; Jeong, M.H.; Jang, S.H.; Oh, S.O.; Hur, J.S. (2015). "Three new monotypic genera of the caloplacoid lichens (Teloschistaceae, lichen-forming ascomycetes)". Mycobiology. 43 (3): 195–202. doi:10.5941/MYCO.2015.43.3.195. PMC 4630424. PMID 26539034.
  85. ^ Kondratyuk et al. 2017, p. 107.
  86. ^ 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.
  87. ^ Bungartz, Søchting & Arup 2020, p. 534.
  88. ^ Kondratyuk et al. 2017, p. 108.
  89. ^ a b c d Kondratyuk, S.Y.; Lőkös, L.; Farkas, E.; Kärnefelt, I.; Thell, A.; Yamamoto, Y.; Hur, J.-S. (2020). "Three new genera of the Teloschistaceae proved by three gene phylogeny" (PDF). Acta Botanica Hungarica. 62 (1–2): 109–136. doi:10.1556/034.62.2020.1-2.7. S2CID 226056287.
  90. ^ 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].
  91. ^ a b Bungartz, Frank; Søchting, Ulrik; Arup, Ulf (2021). "Obscuroplaca gen. nov. – a replacement name for Phaeoplaca; Teloschistaceae (lichenized Ascomycota) from the Galapagos Islands". Plant and Fungal Systematics. 66 (2): 240–241. doi:10.35535/pfsyst-2021-0022. S2CID 245619926.
  92. ^ Bungartz, Søchting & Arup 2020, p. 540.
  93. ^ a b Kondratyuk, S.Y.; Persson, P.-E.; Hansson, M.; Lőkös, L.; Kondratiuk, A.S.; Fayyaz, I; Kouser, R.; Afshan, N.S.; Niazi, A.R.; Zulfiqar, R.; Khalid, A.N.; Kärnefelt, I.; Farkas, E.; Hur, J.-S.; Thell, A. (2022). "Contributions to Molecular Phylogeny of Lichens 4. New names in the Teloschistaceae". Acta Botanica Hungarica. 64 (3–4): 313–336. doi:10.1556/034.64.2022.3-4.7. S2CID 253796366.
  94. ^ Kondratyuk et al. 2017, p. 112.
  95. ^ Kondratyuk et al. 2017, p. 113.
  96. ^ Massalongo, A.B. (1852). "Monografia dei licheni blasteniospori" [Monograph of blasteniospore lichens]. Atti dell'Istituto Veneto Scienze. 2 (in Italian) (3): 5–131 [119].
  97. ^ Arup, Søchting & Frödén 2013, p. 74.
  98. ^ Poelt, J. (1983). "Musterbeispiele analoger Lagerdifferenzierung bei Flechten: Almbornia, Speerschneidera, Seirophora gen. nov" [Exemplary cases of analog differential storage in lichens: Almbornia, Speerschneidera, Seirophora gen nov.]. Flora (in German). 174 (5/6): 439–445. doi:10.1016/S0367-2530(17)31400-7.
  99. ^ Bungartz, Søchting & Arup 2020, p. 548.
  100. ^ a b Arup, Søchting & Frödén 2013, p. 75.
  101. ^ Kondratyuk et al. 2017, p. 118.
  102. ^ Wilk et al. 2021, pp. 292–293.
  103. ^ Kondratyuk et al. 2013, p. 265.
  104. ^ Wilk et al. 2021, pp. 293–294.
  105. ^ a b Kondratyuk et al. 2017, p. 86.
  106. ^ Kondratyuk et al. 2013, p. 267.
  107. ^ Dodge, C.W. (1966). "New lichens from Chile". Nova Hedwigia. 12: 307–352.
  108. ^ Kondratyuk et al. 2013, p. 268.
  109. ^ a b c d Wilk et al. 2021, p. 284.
  110. ^ Arup, Søchting & Frödén 2013, p. 61.
  111. ^ Kondratyuk et al. 2017, p. 96.
  112. ^ a b Kondratyuk, S.Y.; Persson, P.-E.; Hansson, M.; Lőkös, L.; Liu, D.; Hur, J.-S.; Kärnefelt, I.; Thell, A. (2018). "Hosseusiella and Rehmanniella, two new genera in the Teloschistaceae" (PDF). Acta Botanica Hungarica. 60 (1–2): 89–113. doi:10.1556/034.60.2018.1-2.7.
  113. ^ Kondratyuk et al. 2017, p. 105.
  114. ^ Fayyaz, Iram; Kouser, Rubina; Afshan, Najam-ul-Sehar; Niazi, Abdul Rehman; Zulfiqar, Rizwana; Khalid, Abdul Nasir; Kondratyuk, Sergey Yakovych (2022). "Iqbalia kashmirensis gen. et sp. nov. from Pakistan (Teloschistaceae, lichenized ascomycetes) based on multigene phylogeny". Mycological Progress. 21 (8): 1–15. doi:10.1007/s11557-022-01823-y. S2CID 251031210.
  115. ^ a b Kondratyuk, S.; Kärnefelt, I. (1997). "Josefpoeltia and Xanthomendosa, two new genera in the Teloschistaceae (lichenized Ascomycotina)". In Türk, Roman; Zorer, Roberto (eds.). Progress and Problems in Lichenology in the Nineties. Bibliotheca Lichenologica. Vol. 68. Berlin/Stuttgart: J. Cramer in der Gebrüder Borntraeger Verlagsbuchhandlung. pp. 19–44. ISBN 978-3-443-58047-6.
  116. ^ Kondratyuk et al. 2013, p. 269.
  117. ^ Kondratyuk et al. 2017, p. 109.
  118. ^ Kondratyuk, S.Y.; Lőkös, L.; Farkas, E.; Jang, S.-H.; Liu, D.; Halda, J.; Persson, P.-E.; Hansson, M.; Kärnefelt, I.; Thell, A.; Fačkovcová, Z.; Yamamoto, Y.; Hur, J.-S. (2019). "New and noteworthy lichen-forming and lichenicolous fungi 9". Acta Botanica Hungarica. 61 (3–4): 325–367. doi:10.1556/034.61.2019.3-4.6. S2CID 229283962.
  119. ^ Kondratyuk et al. 2015, p. 330.
  120. ^ Kondratyuk et al. 2017, p. 111.
  121. ^ 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.
  122. ^ Kondratyuk et al. 2015b, p. 331.
  123. ^ a b c Wilk et al. 2021, p. 287.
  124. ^ Arup, Søchting & Frödén 2013, p. 62.
  125. ^ Arup, Søchting & Frödén 2013, p. 64.
  126. ^ Kondratyuk et al. 2015b, p. 334.
  127. ^ Kondratyuk et al. 2015b, p. 335.
  128. ^ Wilk et al. 2021, pp. 279, 284, 287, 294.
  129. ^ 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.
  130. ^ Wilk et al. 2021, pp. 279, 287, 294.
  131. ^ Arup, Søchting & Frödén 2013, p. 65.
  132. ^ Kondratyuk et al. 2015b, p. 337.
  133. ^ a b Arup, Søchting & Frödén 2013, p. 66.
  134. ^ a b Kondratyuk, S.Y.; Mosyakin, S.L. (2022). "Wilketalia S.Y.Kondr., a new name for Andina Wilk, Pabijan & Lücking, nom. illeg. (Teloschistaceae, lichenized Ascomycota)". Ukrainian Botanical Journal. 79 (1): 3–5. doi:10.15407/ukrbotj79.01.003. S2CID 247535880.
  135. ^ a b Søchting, Ulrik; Garrido-Benavent, Isaac; Seppelt, Rod; Castello, Miris; Pérez-Ortega, Sergio; De Los Ríos Murillo, Asunción; Sancho, Leopoldo Garcia; Frödén, Patrik; Arup, Ulf (2014). "Charcotiana and Amundsenia, two new genera in Teloschistaceae (lichenized Ascomycota, subfamily Xanthorioideae) hosting two new species from continental Antarctica, and Austroplaca frigida, a new name for a continental Antarctic species". The Lichenologist. 46 (6): 763–782. doi:10.1017/S0024282914000395. S2CID 86840341.
  136. ^ Arup, Søchting & Frödén 2013, p. 34.
  137. ^ Arup, Søchting & Frödén 2013, p. 37.
  138. ^ Arup, Søchting & Frödén 2013, p. 38.
  139. ^ Arup, Søchting & Frödén 2013, p. 40.
  140. ^ a b Kondratyuk, S.Y.; Kärnefelt, I.; Lőkös, L.; Hur, J.S.; Thell, A. (2018). "Coppinsiella and Seawardiella – two new genera of the Xanthorioideae (Teloschistaceae, lichen-forming Ascomycota)" (PDF). Acta Botanica Hungarica. 60 (3–4): 369–386. doi:10.1556/034.60.2018.3-4.8. S2CID 91496354.
  141. ^ Kondratyuk et al. 2017, p. 79.
  142. ^ Luyken, J.A. (1809). Tentamen historiae lichenum [A Study of Lichen History] (in Latin). Gottingen. p. 93.
  143. ^ Arup, Søchting & Frödén 2013, p. 44.
  144. ^ Kondratyuk et al. 2017, p. 87.
  145. ^ a b c d e Fedorenko, Natalya M.; Stenroos, Soili; Thell, Arne; Kärnefelt, Ingvar; Elix, John A.; Hur, Jae-Seoun; Kondratyuk, Sergij Y. (2012). "Molecular phylogeny of xanthorioid lichens (Teloschistaceae, Ascomycota), with notes on their morphology". In Kärnefelt, Ingvar; Seaward, Mark R.D.; Thell, Arne (eds.). Systematics, Biodiversity and Ecology of Lichens. Bibliotheca Lichenologica. Vol. 108. pp. 45–64. ISBN 978-3-443-58087-2.
  146. ^ a b c d e f Kondratyuk, S.Y.; Kärnefelt, I.; Thell, A.; Elix, J.A.; Kim, J.; Jeong, M.H.; Yu, N.N.; Kondratiuk, A.S.; Hur, J.S. (2014). "A revised taxonomy of the subfamily Xanthorioideae (Teloschistaceae, Ascomycota) based on molecular phylogeny". Acta Botanica Hungarica. 56 (1–2): 141–178. doi:10.1556/abot.56.2014.1-2.12.
  147. ^ Arup, Søchting & Frödén 2013, p. 31.
  148. ^ Kondratyuk et al. 2017, p. 101.
  149. ^ Bungartz, Søchting & Arup 2020, p. 527.
  150. ^ Kondratyuk et al. 2015b, p. 329.
  151. ^ Arup, Søchting & Frödén 2013, p. 47.
  152. ^ a b Fedorenko, Natalya M.; Stenroos, Soili; Thell, Arne; Kärnefelt, Ingvar; Kondratyuk, Sergey Y. (2009). "A phylogenetic analysis of xanthorioid lichens (Teloschistaceae, Ascomycota) based on ITS and mtSSU sequences". In Thell, Arne; Seaward, Mark; Feuerer, Tassilo (eds.). Diversity of Lichenology – Anniversary Volume. Bibliotheca Lichenologica. Vol. 100. J. Cramer in der Gebrüder Borntraeger Verlagsbuchhandlung. pp. 49–84. ISBN 978-3-443-58079-7.
  153. ^ 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.
  154. ^ Arup, Søchting & Frödén 2013, p. 48.
  155. ^ Arup, Søchting & Frödén 2013, p. 49.
  156. ^ Hue, A. (1908). "Quator lichenum exoticorum" [Four exotic lichens]. Bulletin de la Société linnéenne de Normandie. 6 (in Latin). 1: 75.
  157. ^ a b c Arup, Søchting & Frödén 2013, p. 55.
  158. ^ Kondratyuk et al. 2017, p. 115.
  159. ^ Kondratyuk et al. 2017, p. 117.
  160. ^ Duvigneaud, P.A. (1941). "Xanthodactylon Duvign., genre nouveau de lichen de l'Afrique du Sud" [Xanthodactylon Duvign., a new genus of lichen from South Africa, related to the lichens of the genera Teloschistes and Apatoplaca (Teloschistaceae)]. Bulletin du Jardin Botanique de l'État à Bruxelles (in French). 16 (2–3): 259–265. doi:10.2307/3666548. JSTOR 3666548.
  161. ^ Santesson, R. (1949). "Dolichocarpus and Xanthopeltis, two new lichen genera from Chile". Svensk Botanisk Tidskrift. 43 (2–3): 547–567.
  162. ^ Kondratyuk, S.Y.; Kim, J.A.; Yu, N.-H.; Jeong, M.-H.; Jang, S.H.; Kondratiuk, A.S.; Zarei-Darki, B.; Hur, J.-S. (2015). "Zeroviella, a new genus of xanthorioid lichens (Teloschistaceae, Ascomycetes) proved by three gene phylogeny". Ukrainian Botanical Journal. 72 (6): 574–584. doi:10.15407/ukrbotj72.06.574.
  163. ^ Brodo, Irwin M.; Sharnoff, Sylvia Duran; Sharnoff, Stephen (2001). Lichens of North America. Yale University Press. p. 319. ISBN 978-0-300-08249-4.
  164. ^ García, Renato; Magnin, Lucia; Miotti, Laura; Barrientos, Gustavo (2020). "Lichens growing on human bone remains: A case study from continental Patagonia (Deseado Massif, Santa Cruz, Argentina)". Journal of King Saud University – Science. 32 (3): 2219–2221. doi:10.1016/j.jksus.2020.02.029. S2CID 216163263.
  165. ^ 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.
  166. ^ Diederich, Paul; Lawrey, James D.; Ertz, Damien (2018). "The 2018 classification and checklist of lichenicolous fungi, with 2000 non-lichenized, obligately lichenicolous taxa". The Bryologist. 121 (3): 340–425 [364–366]. doi:10.1639/0007-2745-121.3.340. S2CID 92396850.
  167. ^ Kantvilas, Gintaras; Suija, Ave; Motiejūnaitė, Jurga (2021). "Caloplaca tephromelae (Teloschistaceae), a new lichenicolous species from Tasmania". The Lichenologist. 53 (4): 317–325. doi:10.1017/s0024282921000207. S2CID 236502837.
  168. ^ Søchting, Ulrik; Olech, Maria (1995). "The lichen genus Caloplaca in polar regions". The Lichenologist. 27 (6): 463–471. doi:10.1016/s0024-2829(95)80006-9. S2CID 83696222.
  169. ^ Garrido-Benavent, Isaac; Pérez-Ortega, Sergio (2017). "Past, present, and future research in bipolar lichen-forming fungi and their photobionts". American Journal of Botany. 104 (11): 1660–1674. doi:10.3732/ajb.1700182.
  170. ^ Vondrák, Jan; Frolov, Ivan; Říha, Pavel; Hrouzek, Pavel; Palice, Zdeněk; Nadyeina, Olga; Halici, Gökhan; Khodosovtsev, Alexander; Roux, Claude (2013). "New crustose Teloschistaceae in Central Europe". The Lichenologist. 45 (6): 701–722. doi:10.1017/s0024282913000455. hdl:10995/27377. S2CID 85942896.
  171. ^ Vondrák, Jan; Frolov, Ivan; Davydov, Evgeny A.; Yakovchenko, Lidia; Malíček, Jiří; Svoboda, Stanislav; Kubásek, Jiří (2019). "The lichen family Teloschistaceae in the Altai-Sayan region (Central Asia)". Phytotaxa. 396 (1): 1. doi:10.11646/phytotaxa.396.1.1. S2CID 92315392.
  172. ^ a b c d Vondrák, Jan; Ismailov, Aziz; Urbanavichus, Gennadii (2017). "Lichens of the family Teloschistaceae in Dagestan, an eastern part of the Caucasian biodiversity hot-spot". Nova Hedwigia. 104 (4): 483–498. doi:10.1127/nova_hedwigia/2016/0387.
  173. ^ Mishra, G.K.; Upreti, D.K.; Nayaka, S.; Thell, A.; Kärnefelt, I.; Lőkös, L.; Hur, J.-S.; Sinha, G.P.; Kondratyuk, S.Y. (2020). "Current taxonomy of the lichen family Teloschistaceae from India with descriptions of new species". Acta Botanica Hungaricae. 62 (3–4): 309–391. doi:10.1556/034.62.2020.3-4.5. S2CID 228982203.
  174. ^ Bungartz, Søchting & Arup 2020.
  175. ^ Herrera-Campos, Ma. de los Ángeles; Lücking, Robert; Pérez-Pérez, Rosa Emilia; Miranda-González, Ricardo; Sánchez, Norberto; Barcenas-Peña, Alejandrina; Carrizosa, Abraham; Zambrano, Angel; Ryan, Bruce D.; Nash, Thomas H. (2014). "Biodiversidad de líquenes en México" [Biodiversity of lichens in Mexico]. Revista Mexicana de Biodiversidad (in Spanish). 85: 82–99. doi:10.7550/rmb.37003.
  176. ^ 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.
  177. ^ Freire-Rallo, Sandra; Wedin, Mats; Diederich, Paul; Millanes, Ana M. (2023). "To explore strange new worlds – The diversification in Tremella caloplacae was linked to the adaptive radiation of the Teloschistaceae". Molecular Phylogenetics and Evolution. 180: 107680. doi:10.1016/j.ympev.2022.107680. PMID 36572164.
  178. ^ Freire-Rallo, Sandra; Diederich, Paul; Millanes, Ana M.; Wedin, Mats (2023). "Five new species in the Tremella caloplacae complex". The Lichenologist. 55 (5): 223–239. doi:10.1017/s0024282923000154.
  179. ^ Tudor, Phoebe B.; Matero, Frank G.; Koestler, Robert J. (1990). "A Case Study of the Compatibility of Biocidal Cleaning and Consolidation in the Restoration of a Marble Statue". In Llewellyn, G.C.; O’Rear, C.E. (eds.). Biodeterioration Research. Vol. 3. Boston, MA: Springer US. pp. 525–533. doi:10.1007/978-1-4757-9453-3_42. ISBN 978-1-4757-9455-7.
  180. ^ Nayak, Sandeep Kumar; Behera1, Prashant Kumar; Bajpai, Rajesh; Upreti, Dalip Kumar; Satapathy, Kunja Bihari (2017). "Lichens growth on Sun Temple of Konark in Odisha, India- A curse or blessing". Cryptogam Biodiversity and Assessment. 2 (02): 48–52. doi:10.21756/cab.v2i02.11119.{{cite journal}}: CS1 maint: numeric names: authors list (link)
  181. ^ a b Crawford, Stuart (2019). "Lichens Used in Traditional Medicine". In Ranković, Branislav (ed.). Lichen Secondary Metabolites. Bioactive Properties and Pharmaceutical Potential (2 ed.). Springer Nature Switzerland AG. p. 76. ISBN 978-3-030-16813-1.
  182. ^ Belguidoum, Amina; Haichour, Rima; Lograda, Takia; Ramdani, Messaoud (2022). "Biomonitoring of air pollution by lichen diversity in the urban area of Setif, Algeria". Biodiversitas Journal of Biological Diversity. 23 (2): 970–981. doi:10.13057/biodiv/d230240. S2CID 247442207.
  183. ^ Brandt, Annette; de Vera, Jean-Pierre; Onofri, Silvano; Ott, Sieglinde (2014). "Viability of the lichen Xanthoria elegans and its symbionts after 18 months of space exposure and simulated Mars conditions on the ISS". International Journal of Astrobiology. 14 (3): 411–425. doi:10.1017/s1473550414000214. S2CID 53421340.
  184. ^ Ravera, S. (2017). "Caloplaca rinodinae-albae". IUCN Red List of Threatened Species. 2017. Retrieved 23 August 2023.
  185. ^ Sokoloff, P.; McMullin, T. (2020). "Seirophora aurantiaca". IUCN Red List of Threatened Species. 2020. Retrieved 23 August 2023.
  186. ^ 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.
  187. ^ Sparkes, J.H.; de Lange, P.J.; Blanchon, D.J. (2014). "Notes on Caloplaca allanii Zahlbr. (Teloschistaceae) a poorly known West Auckland, North Island, New Zealand endemic". New Zealand Journal of Botany. 52 (3): 304–309. doi:10.1080/0028825x.2014.891240. S2CID 85673860.
  188. ^ Bungartz, Søchting & Arup 2020, p. 516.
  189. ^ Xinli, Wei; Hong, Deng; Jiangchun, Wei (2020). "Threatened categories assessment of lichens in China". Biodiversity Science. 28 (1): 54–65 [58]. doi:10.17520/biods.2019154. S2CID 218807267.

Cited literature

  • Arup, Ulf; Søchting, Ulrik; Frödén, Patrik (2013). "A new taxonomy of the family Teloschistaceae". Nordic Journal of Botany. 31 (1): 16–83. doi:10.1111/j.1756-1051.2013.00062.x.
  • Bungartz, Frank; Søchting, Ulrik; Arup, Ulf (2020). "Teloschistaceae (lichenized Ascomycota) from the Galapagos Islands: a phylogenetic revision based on morphological, anatomical, chemical, and molecular data". Plant and Fungal Systematics. 65 (2): 515–576. doi:10.35535/pfsyst-2020-0030. S2CID 234385561.
  • Kondratyuk, S.; Jeong, M.-H.; Yu, N.-H.; Kärnefelt, I.; Thell, A.; Elix, J.; Kim, J.; Kondratyuk, A.; Hur, J.-S. (2013). "Four new genera of teloschistoid lichens (Teloschistaceae, Ascomycota) based on molecular phylogeny". Acta Botanica Hungarica. 55 (3–4): 251–274. doi:10.1556/abot.55.2013.3-4.8.
  • Kondratyuk, S.Y.; Kärnefelt, I.; Thell, A.; Elix, J.A.; Kim, J.; Kondratiuk, A.S.; Hur, J.-S. (2015b). "Brownlielloideae, a new subfamily in the Teloschistaceae (Lecanoromycetes, Ascomycota)" (PDF). Acta Botanica Hungarica. 57 (3–4): 321–343. doi:10.1556/034.57.2015.3-4.6.
  • Kondratyuk, S.Y.; Lőkös, L.; Upreti, D.K.; Nayaka, S.; Mishra, G.K.; Ravera, S.; Jeong, M.-H.; Jang, S.-H.; Park, J.S.; Hur, J.S. (2017). "New monophyletic branches of the Teloschistaceae (lichen-forming Ascomycota) proved by three gene phylogeny". Acta Botanica Hungarica. 59 (1–2): 71–136. doi:10.1556/034.59.2017.1-2.6. hdl:10447/414429.
  • Wilk, Karina; Pabijan, Maciej; Saługa, Marta; Gaya, Ester; Lücking, Robert (2021). "Phylogenetic revision of South American Teloschistaceae (lichenized Ascomycota, Teloschistales) reveals three new genera and species". Mycologia. 113 (2): 278–299. doi:10.1080/00275514.2020.1830672. PMID 33428561. S2CID 231586897.