Academia.eduAcademia.edu
Fungal Diversity DOI 10.1007/s13225-014-0319-0 REVIEW Overview of Stachybotrys (Memnoniella) and current species status Yong Wang & Kevin D. Hyde & Eric H. C. McKenzie & Yu-Lan Jiang & De-Wei Li & De-Gang Zhao Received: 3 September 2014 / Accepted: 21 December 2014 # School of Science 2015 Abstract Stachybotrys (asexual Hypocreales) has a worldwide distribution. This genus inhabits substrates rich in cellulose and is closely related to Memnoniella. Classification of species has previously been based on morphology, with conidial characters being considered as important. This study reevaluates Stachybotrys and Memnoniella, which is shown to include at least seven species-groups; while Memnoniella is a synonym of Stachybotrys. The sexual genera Ornatispora and Melanopsamma are also synonyms of Stachybotrys. With the exception of Stachybotrys subsimplex, species formed a wellsupported monophyletic group in LSU data analysis belonging to Stachybotriaceae. Seventy-four accepted Stachybotrys species are discussed, while eight species are considered to belong to other genera or are doubtful in this paper and a key to these species provided. Twelve new combinations and 1 nomina nova is proposed. The status of Stachybotrys species on health, as Electronic supplementary material The online version of this article (doi:10.1007/s13225-014-0319-0) contains supplementary material, which is available to authorized users. Y. Wang : D.<G. Zhao Guizhou Key Laboratory Agro-Bioengineering, Guizhou University, Guiyang, Guizhou 550025, People’s Republic of China Y. Wang : Y.<L. Jiang Department of Plant Pathology, College of Agriculture, Guizhou University, Guiyang, Guizhou 550025, China K. D. Hyde Institute of Excellence in Fungal Research, School of Science, Mae Fah Luang University, Chiang Rai, Thailand E. H. C. McKenzie Landcare Research, Private Bag 92170, Auckland, New Zealand D.<W. Li (*) Valley Laboratory, The Connecticut Agricultural Experiment Station, 153 Cook Hill Road, Windsor, CT 06095, USA e-mail: Dewei.Li@ct.gov human or animal pathogens, in indoor environments, and use as biocontrol agents and compound discovery are also discussed. Keywords Current species . Nomen excludendum . Phylogeny . Synonyms . Taxonomic history Introduction The importance of Stachybotrys/Memnoniella Stachybotrys and Memnoniella are long-standing genera comprising more than 100 names. The classification of species in these genera is of particular importance as some species have medical/public health significance. For example, Stachybotrys chartarum has become notorious in the last 15 years. It is known as ‘black mold’ or ’toxic mold’ due to its ability to produce some rather potent mycotoxins and its association with long-term water damage. The presence of this species and potential health effects led to a large number of litigations in the USA during the same period. The taxonomy of species in these genera is unclear and the confusion that exists is a major hindrance to research development concerning public health aspects of these ‘toxic indoor fungi’. Health effects of Stachybotrys/Memnoniella Stachybotrys chartarum is a soilborne fungus and one of the major species occurring on cellulose-based building materials experiencing long-term water damage indoors. Stachybotrys chartarum, also called the “toxic black mold” by the public and media, has gained its notorious fame as a mycotoxin producer that can cause mycotoxicosis (stachybotrytoxicosis) in animals and humans and has been suspected of causing respiratory symptoms in humans, such as acute infant pulmonary hemorrhage, asthma, adult nasal and tracheal bleeding, allergies, as well as asthma-like symptoms, inflammation, and Fungal Diversity lung injury (Etzel et al. 1998; Vesper et al. 2001; Vesper and Vesper 2002; Al-Ahmad et al. 2010; Piecková et al. 2009; Yike and Dearborn 2011; Bhan et al. 2011). Although the causal relationship between pulmonary hemorrhage and S. chartarum and the role this species plays in public health is still highly controversial (Pestka et al. 2008; Yike and Dearborn 2011), infant pulmonary hemorrhage incidents in Cleveland brought this toxigenic fungus and other indoor molds to the immediate attention of the medical community, media and the public. Cases of infant pulmonary hemorrhage associated with the presence of S. chartarum in Cleveland continue to occur, increasing from nine cases in the original study to 52 cases at present. Among the cases investigated, 91 % of patients were living in residences in which S. chartarum were found (Yike and Dearborn 2011). More than ten papers from four different laboratories have reported pulmonary hemorrhage in acute animal models of instillation of S. chartarum conidia into rodent airways (Yike and Dearborn 2011). Chung et al. (2010) compared the allergenicity of S. chartarum to house dust mite extracts in a mouse model and established a suggested threshold dose (10 μg) for S. chartarum allergy induction. They concluded that exposure to S. chartarum might be easily over the sensitization threshold for a susceptible population in damp water-damaged environments. Rakkestad et al. (2010) reported that heat-treated conidia of S. chartarum induced cell death (apoptosis) within 3–6 h due to DNA damage. Nagayoshi et al. (2011) reported for the first time that the remodeling of pulmonary arteries in mice was a result of inhalation exposure to the conidia of S. chartarum. Yike and Dearborn (2011) considered this study to have significantly advanced our knowledge of the pathologic effects of S. chartarum. Recent studies found that repeated inhalation of S. chartarum conidia caused pulmonary hypertension and evoked pulmonary arterial remodeling in mice (Nagayoshi et al. 2011). Bhan et al. (2011) found that S. chartarum induced hypersensitivity pneumonitis. Characteristic symptoms and immunological tests for antibodies (IgE and IgG) specific to S. chartarum and some other fungi strongly suggest exposure to the indoor fungi. An epidemiological study reported a high prevalence of pulmonary diseases among office workers of Florida court buildings following prolonged indoor exposure to S. chartarum and Aspergillus versicolor (Hodgson et al. 1998). Stachybotrys atra was isolated from bronchoalveolar lavage fluid of a child with pulmonary hemorrhage (Elidemir et al. 1999) and S. chartarum exposure was found in an infant that developed laryngeal spasm and hemorrhage during general anesthesia (Tripi et al. 2000). Nielsen et al. (2002) showed that 35 % of the isolates of S. chartarum produced extremely cytotoxic satratoxins. Similar results showed that 39 % of S. chartarum isolates produced macrocyclic trichothecenes (Andersen et al. 2002). The toxicity of the isolates producing macrocyclic trichothecenes is 1000 times that of other isolates, which produce atronones (Jarvis 2003, pers. com.). Macrocyclic trichothecenes were detected from urine and tissue samples of patients (Straus 2011). Exposure to S. chartarum or the mycotoxins originating from this fungus is considered a potentially serious public health threat (Yike and Dearborn 2011). Stachybotrys chartarum is able to produce some of the most potent mycotoxins, macrocyclic trichothecenes and related trichoverroids: roridin E and L-2; satratoxins F, G, and H; isosatratoxins F, G, and H; verrucarins B and J; and the trichoverroids, trichoverrols A and B and trichoverrins A and B. The mycotoxins occur in all parts of the fungus (Sorenson et al. 1987; Jarvis 2002). Stachybotrys (Memnoniella) echinata is mycotoxigenic producing trichodermol, trichodermin, dechlorogrisseofulvins, memnobotrins A and B, memenoconol, memnoconone (Jarvis 2002). Stachybotrys as potential human or animal pathogens Stachybotrys species were first identified as pathogens in the Ukraine in the early 1930s, a unique disease of horses was characterized by lip edema, stomatitis, oral necrosis, rhinitis, and conjunctivitis (Forgacs 1972). Several outbreaks of livestock disease had been reported all over the world and was identified as stachybotryotoxicosis (Forgacs et al. 1958; Ye et al. 1998). The possible association of Stachybotrys species with human disease became apparent coincident with the equine epidemics. Primary disease appeared on the skin, lesions progressed from hyperemia to crusting exudates to necrosis, with subsequent resolution (Linnik 1949). The lesions were due to aerosolization of the offending substances, with primary effects in dermal areas with abundant moisture and skin-to-skin contact. Some patients suffered erosions on the oral and gingival mucosa. Respiratory symptoms were described, including catarrhal angina, bloody rhinitis, cough, throat pain, chest tightness, and occasional fever. Some patients experienced transient leukocytopenia. It was found that inhalation of conidia of S. chartarum may cause serious damage to the human lung, particularly when recurring (Ochiai et al. 2005). Stachybotrys species are quite resistant to adverse environmental conditions. The results disclosed that the conidia of S. chartarum were resistant to the antifungal activities of alveolar macrophages in terms of phagocytosis, killing and inhibition of germination. These illnesses could be attributed in part to mycotoxins released by S. chartarum. Hemolysin released by this mold was found to be hemolytic in vitro and in vivo. In addition, allergenic proteins have been characterized from S. chartarum (Ochiai et al. 2005). The exact mechanism of S. chartarum pathogenesis has not been defined (Hossain et al. 2004; Ochiai et al. 2005). Studies found that inhalation of Stachybotrys chartarum causes pulmonary arterial hypertension in mice, however, there is no direct evidence on pathological mechanism (CDC 2012; Ochiai et al. 2005). An endemic outbreak of fungal meningitis, spinal infections and other serious health Fungal Diversity complications in the USA in 2012 was caused by injecting directly into the spinal fluid or joints of patients of a steroid medication for pain control, methylprednisolone acetate, from a compounding pharmacy that reportedly was contaminated by several fungi. However, among the fungi, S. chartarum was isolated from a patient’s infected tissue (CDC 2012). Hypersensitivity pneumonitis (HP) is an inflammatory lung disease that develops after repeated exposure to inhaled particulate like S. chartarum (Bhan et al. 2011). This study suggests that TLR9 (Toll –Like-Receptor 9) is critical for the development of Th17-mediated granulomatous inflammation in the lung in response to S. chartarum. A recent study, conducted to understand toxin-regulated gene expression of S. chartarum and Aspergillus versicolor, suggested that there was no general correlation between gene expression and fragment sizes; however, all submicron fragments may contribute to inflammatory response (Pei and Gunsch 2013). Data suggests that acute exposure to trichothecene mycotoxin and Stachybotrys in the indoor air of water-damaged buildings have potential adverse health effects on neurotoxicity and inflammation within the nose and brain. Satratoxin G (SG) is a macrocyclic trichothecene mycotoxin produced by S. chartarum specifically induced apoptosis of olfactory sensory neurons (OSNs) in the olfactory epithelium (Islam et al. 2006). Further results demonstrate that the macrocyclic SG was neurotoxic in vitro and in vivo, while its biosynthetic precursor, roridin L2 was nontoxic (Islam et al. 2009). Pathogenesis studies depicts that SG was readily absorbed from the nose, distributed to tissues involved in respiratory, immune, and neuronal function, and subsequently cleared. However, a significant amount of the toxin was retained in the nasal turbinate, sufficient to evoke olfactory sensory neurons (OSNs) death (Amuzie et al. 2010). Moreover, in vivo studies induced rhinitis and apoptosis of (OSNs) in the nasal airways of rhesus monkeys whose nasal airways more closely resemble those of humans. These results provide new insight into the potential risk of nasal airway injury and neurotoxicity caused by exposure to Stachybotrys toxins in water-damaged buildings (Carey et al. 2012). Stachybotrys spp. in indoor environments Stachybotrys chartarum and its mycotoxins have been linked to damp building-associated illnesses (Frazer et al. 2012). Shelton et al. (2002) evaluated 9619 indoor samples and 2407 outdoor samples collected from 1717 buildings located across the United States. Stachybotrys chartarum was identified in the air in 6 % of the buildings studied and in 1 % of the outdoor samples. Stachybotrys chartarum DNA was detected in 28 % of samples collected from hotel rooms and odour was a predictor of S. chartarum DNA (Norback and Cai 2011). Foarde and Menetrez (2002) showed that conidia of S. chartarum released from gypsum boards at low air flow rate were positively related to air flow rate, but negatively related to relative humidity. Data showed that S. chartarum trichothecene mycotoxins can become airborne in association with highly respirable smaller particles (diameter <1 μm) specifically in water-damaged buildings (Brasel et al. 2005). Growth of S. chartarum was optimal at 25 to 30 °C at 0.995 aw, but this was modified to 0.98 aw at 30 °C for a macrocyclic trichothecene-producing strain (IBT 7711) and a nonproducing strain (IBT 1495) (1.4–1.6 mm/day, respectively) (Frazer et al. 2012). The ELISA method revealed that, in contrast to growth, satratoxin G production was maximal at 20 °C with its highest production at 0.98 aw (approximately 250 μg/g mycelia). When water was freely available (0.995 aw), satratoxin G was maximally produced at 15 °C and decreased as temperature was increased (Frazer et al. 2012). It was concluded that, textile seats are much more contaminated fungal DNA than leather seats (Fu et al. 2013). Stachybotrys (Memnoniella) echinata and S. chlorohalonata are also the hydrophilic species associated with long-term damage. Stachybotrys chartarum and S. echinata sometimes grow together indoors. These two species often share the same habitats with Chaetomium globosum Kunze or Ulocladium spp. Li and Yang (2004a, 2004b) observed occurrence of S. chartarum, S. yunnanensis, S. nephrospora, S. microspora, S. elegans, and S. chlorohalonata in indoor environments. The exact number of species of Stachybotrys present in indoor environments is not yet clear. Stachybotrys in biocontrol during mycoparasitism of hyphae and sclerotia The biocontrol of fungal plant pathogens using microorganisms has been identified as a realistic alternative to chemical methods and Stachybotrys elegans was shown to have a strong antagonistic activity against Rhizoctonia solani, the causative agent of Rhizoctonia disease of potatoes. Studies demonstrated that microparasitism is the main mechanism involved in the interaction between Stachybotrys elegans and R. solani (Benyagoub and Jabaji-Hare 1992). The mechanism basically involved the partial degradation of R. solani mycelial and sclerotial cell walls using the production of hydrolytic enzymes by Stachybotrys elegans. Protein electrophoresis revealed that, different isoforms of chitinases and β1,3-glucanases produced by S. elegans were capable of degrading Rhizoctonia solani mycelium (Tweddell et al. 1994). Results indicate that the two b-1,3-glucanases as 75kDa 1,3-beta-glucanase, 94-kDa 1,3-beta-glucanase, one b-Nacetylhexosaminidase (Taylor et al. 2002), and one b-1,4glucanase (Tweddell et al. 1996) are involved in Stachybotrys elegans mycoparasitism (Archambault et al. 1998). Cloning and characterization of an endochitinase gene (sechi44 gene (cDNA) ) from S. elegans using real-time Fungal Diversity quantitative RT-PCR, demonstrated that the sechi44 gene levels are differentially expressed when S. elegans was grown under induced and non-induced conditions, and during its mycoparasitic interaction with R. solani. Characterization of the full chitinolytic system of S. elegans at the gene level is important to find the enzymes effect in pre-contact gene expression and to understand the relevance of this mechanism to biocontrol (Morissette et al. 2003). Recent studies showed that elevated expression of some genes belonging to the mycoparasite and the host play an important role in this interaction. Differential gene expression of eight mycoparasitism-induced genes and eight hostresponse genes was monitored during in vivo interactions between the S. elegans and hyphae and sclerotia of the host, Rhizoctonia solani (Chamoun and Jabaji 2011). Further study of transcriptional regulation of genes on microparasitism suggests that mitogen-activated-protein kinase (MAPK) gene smkA could be implicated in the mycoparasitic process in Stachybotrys elegans. This study also tested differential expression of the gene in S. elegans in response to direct parasitism of different vegetative structures of Rhizoctonia solani (i.e., carbon-rich condition) and to nutrient starvation (i.e., carbon-poor condition) (Chamoun et al. 2013). Stachybotrys for compound discovery, especially S. chartarum Stachybotrys species are responsible for producing a number of toxic chemicals, which mainly cause pathogenesis among other animals and humans. Stachybotrys chartarum produces a variety of secondary metabolites including trichothecenes (Jarvis 1991; Abbas et al. 2012), triprenylated phenolics, and a new class of diterpenoids called atranones (Hossain et al. 2004; Kuhn and Ghannoum 2003; Nielsen 2003). Trichothecenes are a large group of sesquiterpenoid fungal toxines, which share a common core comprised of a rigid tetracyclic ring system, and variations in toxicity are observed among different species for each individual compound (Jarvis 1986, 1991; Pestka et al. 2007). Stachybotrys chartarum is able to produce some of the most potent mycotoxins, macrocyclic trichothecenes and related trichoverroids: roridin E and L-2; satratoxins F, G, and H; isosatratoxins F, G, and H; verrucarins B and J; and the trichoverroids, trichoverrols A and B and trichoverrins A and B. Stachybotrys (Memnoniella) echinata is mycotoxigenic producing trichodermol, trichodermin, dechlorogrisseofulvins, memnobotrins A and B, memenoconol, memnoconone (Jarvis 2002). Trichothecenes have widespread phytotoxicity and cytotoxicity effects throughout the cell. Moreover, these toxins can inhibit protein synthesis through an interaction with eukaryotic ribosomes and members of this class of toxins have been undergoing experimental research on novel class of eukaryote-specific antibiotics lead novel compounds (Gilly et al. 1985; Shanks et al. 2011; Abbas et al. 2012). Eleven naturally occurring atranones (A–K) have been characterized (Hinkley et al. 1999, 2000, 2003). In atranones there are special ring arrangements where as normal bicyclo ring of the dolabellanes has been elaborated by a fused enol lactone (ring D), a most unusual structural feature, especially since it lacks any further conjugative stabilization (Hinkley et al. 2000; Jarvis 2003). A study indicate that interactions between co-cultivation of S. californicus and S. chartarum stimulate the production of an unidentified cytostatic compound(s) capable of inducing mitochondria mediated apoptosis and cell cycle arrest at S-G2/M and give evidence to potential ability to immunotoxic effects similar to those by chemotherapeutic drugs (Penttinen et al. 2006). The detected immunotoxic effects caused by cytostatic compound(s) can impair the ability of macrophages to protect the host against bioaerosols including infectious microorganisms present in indoor air (Penttinen et al. 2006). Fibrinolytic therapy using tissue-type plasminogen activator (tPA) is one of the most effective treatments for acute ischemic stroke. Stachybotrys microspora triprenyl phenol (SMTP)-7 (orniplabin; CAS registry no. 273379-50-9), a novel fibrinolytic agent, is an analog of triprenyl phenol, designated staplabins which was isolated from a fungal culture of S. microspora (IFO 30018) (Hu et al. 2004). Staplabin is a low-molecular-weight compound that stimulates plasminogen-fibrin binding. Results suggest that SMTP-7 has an intrinsic neuroprotective effect on ischemia/ reperfusion injury through the suppression of oxidative stress and MMP-9 activation (Akamatsu et al. 2001). In addition to several small molecule metabolites (MW <800), S. chartarum produces proteins with emolysin (Vesper et al. 2001) and proteinase (Kordula et al. 2002) activities, both of which have been suggested as possible contributing agents in the IPH syndrome in infants. Stachybotrys sp. RF-7260 was found to produce stachyflins, novel anti-influenza virus agents, under solidstate fermentation conditions. Feeding DL-lysine to a culture of Stachybotrys sp. RF-7260 induced the formation of the novel compounds, SQ-02-S-L2 and -L1, and feeding DLvaline the formation of SQ-02-S-VI and -V2 (Minagawa et al. 2002a). In addition, acetylstachyflin, novel antiinfluenza A virus substances were also produced by Stachybotrys sp. RF-7260 (Minagawa et al. 2002b). A recent study revealed that stachyflin inhibit the growth of H1, H2, H5, and H6 influenza viruses by binding the site of the HA2 and preventing the hemagglutinin (HA) from fusion of the virus envelope with cellular membrane (Motohashi et al. 2013). Modern scientific research led to the synthesis of this novel anti-influenza viral agent with more effective manner (Sakurai et al. 2011) Fungal Diversity Novel research on isolation of bioactive molecules has led to isolation and purification of a transglucosilating betaglucosidase system composed of five beta-glucosidases from a Stachybotrys strain (Saibi et al. 2007). The purified enzyme is a monomeric protein of 78 kDa molecular weight and exhibits optimal activity at pH 6.0 and at 50 °C. Recently, the thermostable family 3 β-glucosidase from S. microspora was analysed for its molecular characterization, and mRNA expression (Abdeljalil et al. 2014). Taxonomic history of Stachybotrys and Memnoniella The asexual genus Stachybotrys Corda (1837) was proposed with one species, S. atra (= S. chartarum (Ehrenb.) S. Hughes) collected from wallpaper in a home in Prague. Its related genus, Memnoniella von Höhnel (1923) was erected nearly a century later with the type species Memnoniella aterrima Höhn. Stachybotrys has seven synonyms [Synsporium Preuss (1849), Fuckelina Sacc. (1875), Gliobotrys Höhn. (1902), Memnoniella (1923) (considered separate by Jong and Davis 1976 and Ellis 1976), Spinomyces Saito 1939, nom. inval. Art. 36, (=Memnoniella), Hyalobotrys Pidopl. (1948), Hyalostachybotrys (1958)] (Seifert et al. 2011). Since its introduction, Stachybotrys has been the subject of controversy and debate. The type species, S. atra described by Corda (1837) with 2-celled conidia, is one of these controversies. Based on his critical and extensive study of cultures and herbarium specimens Bisby (1943) indicated that Corda’s generic and specific descriptions were inaccurate. Bisby (1943) revised both generic and species descriptions from 2celled conidia to 1-celled conidia and retained the name S. atra. Bisby redescribed the genus Stachybotrys as “Hyphae, phialophore, and phialides hyaline, brightly coloured, or dark; strands or ropes of hyphae may be produced. Conidia (slime-spores) one celled, normally dark and accumulating into a cluster. The distinctive characteristic of the genus is the septate phialophore or simple conidiophore bearing a crown of phialides and generally becoming dark. A phialophore arises directly from a hypha or frequently, from another phialophore.” As the type material was not available to Bisby, he speculated that the two guttules in the conidia made them appear 2-celled, thus leading to Corda erroneously describing them as 2-celled. Our observations in the last 10 years showed that some isolates of S. chartarum (= S. atra) do develop biguttulate conidia when young (Li per comment), although biguttulation in the conidia disappears when the conidia mature. These observations support Bisby’s speculation. Up to the present time, all accepted members of Stachybotrys produce unicellular conidia without exception (Jong and Davis 1976). Subsequent examination of Corda’s type by De-Wei Li herein has shown the conidia to be 1-celled. Stachybotrys atra was predated by Stilbospora chartarum Ehrenb. (Sylv. mycol. berol.: 9, 21 (1818)), (≡ Oidium chartarum (Ehrenb.) Link, Linné Species Plantarum, Edn 4 6(1): 124 (1824) and Oospora chartarum (Ehrenb.) Wallr., Fl. crypt. Germ. 2: 184 (1833)). Stilbospora chartarum thus has priority over Stachybotrys atra, published in 1837. After examining the type materials of both Stilbospora chartarum and Stachybotrys atra, Hughes (1958) combined the two as Stachybotrys chartarum (Ehrenb.) S. Hughes. However, this new combination was not widely recognized until the 1990’s. Hughes did not make any written observation of the type material to update the previous descriptions. Bisby (1943) clarified some confusion surrounding the genus and its type species by revising the generic and specific descriptions of Stachybotrys, but his treatment of other names was questionable. Bisby (1943) reduced the number of species from over 20 to two—S. atra and S. subsimplex—based on his belief that great variability existed in S. atra. He was correct to demote S. cylindrospora to synonym status according to Li (2007), but was incorrect to demote S. dichroa Grove, and to exclude S. papyrogena (=Memnoniella echinata). Later, he corrected himself and revived S. dichroa and recognized it as a valid species (Bisby and Ellis 1949). Bisby (1943) indicated that Gliobotrys alboviridis may be a synonym of S. subsimplex. However, at present G. alboviridis is considered a synonym of S. albipes and S. albipes is a recognized valid species (Booth 1957; Jong and Davis 1976). Bisby (1943) considered “M. echinata (Rivolta) Galloway, 1933, as similar (to S. subsimplex) except in having catenulate spores.” At present these two are accepted as separate species. Stachybotrys elongata Peck published in 1890 is not a Stachybotrys species. Verona and Mazzucchetti (1968) published a monograph of Stachybotrys and Memnoniella. They accepted 16 species of Stachybotrys and three species of Memnoniella. Jong and Davis (1976) published their culture studies of 11 species of Stachybotrys and two species of Memnoniella. These publications covered <1/3 of currently accepted species due to the exclusion of a number of valid species either published previously or since then. Thirty-three new names of Stachybotrys and five new names of Memnoniella have been proposed since Jong and Davis (1976) published their paper according to Index Fungorum and MycoBank. These species were mainly collected from Asian and Pacific countries, Australia, China, Cuba, French Polynesia, India, Iraq, Japan, Mexico, New Zealand, Solomon Island as well as from Thailand and the USA (Jiang and Zhang 2009; Kong 1997; Kong et al. 2007; Li and Yang 2004a; Matsushima 1995; McKenzie 1991; Misra 1976; Misra and Srivastava 1982; Morgan-Jones and Karr 1976; Morgan-Jones and Sinclair 1980; Whitton et al. 2001). Among these new species, Matsushima (1985, 1989, 1995) published S. queenslandica Matsush., S. ruwenzoriensis Matsush., S. verrucispora Matsush., and Fungal Diversity S. zuckii K. Matsush. & Matsush. McKenzie (1991) published three species: S. breviuscula McKenzie, S. freycinetiae McKenzie and S. nephrodes McKenzie. Stachybotrys kapiti Whitton, McKenzie, and Hyde 2001, S. reniverrucosa Whitton, McKenzie, and Hyde 2001, S. suthepensis Photita, P. Lumyong, KD Hyde & McKenzie and S. waitakere Whitton, McKenzie, and Hyde 2001 were published by McKenzie and co-workers (Whitton et al. 2001; Photita et al. 2003). Pinruan et al. (2004) published two new species, Stachybotrys palmae Pinruan from Thailand and S. cordylines McKenzie from New Zealand, with a key to 55 accepted species. Several species were published by other authors. Zhang and his students described seven species: S. jiangziensis YM Wu & TY Zhang, S. nielamuensis YM Wu & TY Zhang, S. pallescens YL Jiang & TY Zhang, S. terrestris Kong, Zhang, & Zhang, S. variabilis H.F. Wang & TY Zhang and S. xigazenensis Y.M. Wu & TY Zhang and S. zhangmuensis YM Wu & TY Zhang (Jiang and Zhang 2009; Kong et al. 2007; Wang and Zhang 2009; Wu and Zhang 2009, 2011). Stachybotrys thaxteri D.W. Li and S. subreniformis Q.R. Li & Y.L. Jiang are the newest members of this genus (Li 2011; Li and Jiang 2011). In the past 30 years, no comprehensive monographic work has been published on this genus. Index Fungorum (2014) lists 104 epithets of Stachybotrys while MycoBank (2014) records 109 epithets. Seifert et al. (2011) accepted 38 species including Memnoniella based on the literature published up to 2009. Pinruan et al. (2004) accepted 55 taxa in Stachybotrys (including seven taxa formerly placed in Memnoniella). The reality is that it is uncertain how many species should be accepted without a detailed study of type materials, nomenclature, and phylogeny. Among the previously or presently accepted species, a number of species are obscure, doubtful or misplaced due to sketchy diagnoses or type materials that are unknown, lost, or not accessible. For example, Stachybotrys cannae Bat. is Periconiella portoricensis (F. Stevens & Dalbey) M.B. Ellis following the examination of type specimens of S. cannae and P. portoricensis (Li 2011) and Stachybotrys elata Sacc. is Sterigmatobotrys macrocarpa (Corda) S. Hughes (Hughes 1958). There are several more species that might be misplaced and need to be studied; Stachybotrys elongata Peck is one of them. Taxonomy of the genus Stachybotrys, especially the delineation of some species, even the type species, remains controversial due to the poor condition of the holotypes associated with the type species of the genus and with some other species (Koster et al. 2003). Some species were published invalidly. Hyalostachybotrys sacchari Sriniv. and H. bisbyi Sriniv. are nom. inval. (Art. 8), as the types of these two species. Both type cultures of S. bisbyi (SBI 696) and S. sacchari (SBI 781) are currently not maintained at Sugarcane Breeding Institute (SBI), India, according to Dr. R. Viswanathan, Principal Scientist (Plant Pathology) & Head, Division of Crop Protection, Sugarcane Breeding Institute (Pers. Comm. 2012). Since type specimen for Hyalostachybotrys bisbyi, type species of Hyalostachybotrys is a culture, this genus is also invalid (Art. 8). A problem among species of Stachybotrys described to date is that some are described from natural substrates and others from cultures. Another problematic aspect is that the type species, S. chartarum, is loosely interpreted and delineated, which allows for great morphological variation in the species. Twenty-two names proposed in the past are considered to be synonyms of S. chartarum (Bisby 1943; Hughes 1958; Jong and Davis 1976; Li 2007). Differences in the sources of specimens from either natural substrates or culture may be very significant as observed in the ornamentation of conidia of S. sphaerospora as seen from an examination of its type specimen. Stachybotrys chartarum, as described in the literature, shows a significant morphological variability in size, shape, and ornamentation of conidia (Bisby 1943; Ellis 1971; Jong and Davis 1976). Characters of conidia of S. chartarum in the description of Bisby (1943) are “conidia 8–12 (14) × 4–9 (12) μm, elliptical to oval on younger growth of the fungus, often subglobose on older growth” according to examination of three cultures, while that of Jong and Davis (1976) based on the observation of 21 isolates are “when mature, dark olive gray, more or less opaque, smooth-walled or showing banded or ridged, ellipsoidal, unicellular, 7–12×4–6 μm.” These studies really showed just how variable this species can be. The type specimen of S. atra was examined and only three 1-celled conidia were present on the holotype. The type has deteriorated into three tiny clumps, each about 1 mm in diam. This examination did not help clarify the species concept of S. chartarum, the type species of Stachybotrys. According to the type specimen of S. chartarum (≡ Stilbospora chartarum, the basionym of the type species), which was examined, the species concept described by aforementioned authors, especially the one illustrated by Ellis (1971) is too broad and overlaps with S. sphaerospora Morgan-Jones & Sinclair. Saito (1939) created a monotypic genus Spinomyces Saito (J. Ferm. Tech. 17(1): 2) typified with Spinomyces japonicas Saito without a Latin diagnosis (nom. inval., Art. 36). According to the original description (in Japanese) and illustration and an examination of the ex-type culture (CBS 344.39) and the type specimen (CBS H-7327), which Saito deposited to CBS, S. japonicus is Memnoniella echinata (Rivolta) Galloway. Thus, S. japonicus is one of synonyms of M. echinata. Spinomyces Saito (non Spinomyces Bat. & Peres) is monotypic, and thus, Spinomyces is a synonym of Stachybotrys (Memnoniella). A homonym, Spinomyces Bat. & Peres was introduced in 1961 and typified with S. genipae Bat. & Peres (Batista 1961). The current name of Spinomyces Bat. & Peres is Echinoplaca Fée, Essai Fungal Diversity Crypt. Exot. (Paris) 1: l, xciii (1824). Memnonium effusum Corda 1833 is not a Stachybotrys species as von Höhnel (1923) suggested. It is Trichosporum. However, Trichosporum Fr., Syst. orb. veg. (Lundae): 306 (1825) is nom. illegit. (Art. 53.1) and a homonym of Trichosporum D. Don 1822 (Gesneriaceae). To date ten names have been published in Memnoniella according to Index Fungorum (2014) and MycoBank (2014). The only difference between Memnoniella and Stachybotrys is the conidia disposition, in a slimy conidial mass for Stachybotrys and in dry conidial chains for Memnoniella. Five to seven species of Memnoniella were accepted among ten names published (Kirk et al. 2008; Li et al. 2003; Manoharachary et al. 2006). Memnoniella echinata is an important mycotoxigenic fungus indoors (Nielsen 2003). Zuck (1946) observed that M. echinata developed conidia in both dry chains and in slimy aggregates. Our observation of a number of isolates of M. echinata confirmed his results (Li per comment). Memnoniella longistipitata also shows both types of conidial development. Zuck (1946) considered that the isolates developing both types of conidia were intermediate between Stachybotrys and Memnoniella. More intermediate types have been found and described. Stachybotrys zuckii K. Matsush. & Matsush. was described with M. subsimplex-like conidia (Matsushima 1995). Li et al. (2003) described Memnoniella longistipitata D.W. Li et al. with Stachybotrys-type conidia according to morphological and ITS sequence data. Li and Yang (2004b) found an isolate of M. echinata developing Stachybotrys-type conidia. These intermediate isolates/ species provide good material for not only studying the phylogenetic relationships of Stachybotrys and Memnoniella, but also for studying the evolutionary direction of slimy (Stachybotrys) and dry conidia (Memnoniella). Segregating Memnoniella from Stachybotrys has been controversial over the past 50 years. Smith (1962) opinioned that conidial disposition of dry chains and slimy masses is not of sufficient significance to segregate these two genera and demoted Memnoniella to a synonym of Stachybotrys, according to the priority of the two names. Barron (1968), Carmichael et al. (1980), and Seifert et al. (2011) agreed with Smith’s treatment, but not Jong and Davis (1976) and Ellis (1971). Haugland et al. (2001) studied nine species of Stachybotrys and three species of Memnoniella and concluded that Memnoniella is paraphyletic to Stachybotrys. They proposed to demote Memnoniella to synonym status of Stachybotrys based on their phylogenetic study using ITS data. However, ITS may not be sufficient for studying phylogenetic relationships above species level of Stachybotrys and Memnoniella. The type species M. aterrima was also not in their study. Jong and Davis (1976) considered M. aterrima to be a synonym of M. echinata, but there is no indication that the type material was examined. The intermediate species/isolates: M. longistipitata, M. echinata, and S. zuckii provided additional evidence that Memnoniella should be demoted. However, the hypothesis needs to be tested with multiple genes/regions on whether Memnoniella should be segregated from Stachybotrys and whether Memnoniella evolved out of Stachybotrys or vice versa. It is necessary to include more species, especially the intermediate species/isolates and aforementioned morphological and phylogenetically allied genera, using multiple genes in future studies to determine the phylogenetic relationships of these genera and the taxonomic fate of Memnoniella. Ornatispora was created by Hyde et al. (1999) with Ornatispora palmicola K.D. Hyde et al. which possessed the morphological characters of “ascomata superficial, globose, collabent when dry, black, coriaceous, lacking or covered in numerous setae, papillate”. Up to now, seven species have been reported (Hyde et al. 1999; Dulymamode et al. 2001; Whitton et al. 2012). The family placement of Ornatispora is still ambiguous (Hyde et al. 1999). Among them, the asexual morphs of Ornatispora palmicola and O. nepalensis were observed with terminal phialides (Hyde et al. 1999; Whitton et al. 2012), which was similar to Stachybotrys morphology. Melanopsamma pomiformis (Pers.) Sacc. (1878) (≡ Sphaeria pomiformis Pers. 1801) is the type species of Melanopsamma Niessl. Melanopsamma has Stachybotrys and Custingophora-like asexual states and Kirk et al. (2008) estimated that 41 of 130 Melanopsamma epithets were acceptable species. The asexual morph of Me. pomiformis is Stachybotrys albipes (Berk. & Broome) S.C. Jong & Davis. Seifert et al. (2011) stated that the sexual state of Stachybotrys is Melanopsamma, but also listed Chaetosphaeria with the citation of Samuels and Barr (1997). However, the latter authors redescribed Me. pomiformis and listed Chaetosphaeria pomiformis as one of its synonyms. They did not treat other species of Chaetosphaeria. There are contradictions as to which family Melanopsamma belongs. Melanopsamma has been accepted in the Nectriaceae (Munk 1957) the Niessliaceae (Barr 1990), Tang et al. (2007) and Chaetosphaeriaceae (Cannon and Kirk (2007). However, when Chaetosphaeriaceae was proposed (Réblová et al. 1999), Melanopsamma was not treated, and rather Melanopsammella was included in the family. A literature search on Melanopsamma did not yield additional publications to support the opinion of Cannon and Kirk (2007). MycoBank (2014) and Index Fungorum (2014) take different positions on this issue. This is a major confusion which needs to be clarified. Castlebury et al. (2004) opinioned that Melanopsamma pomiformis and asexual Stachybotrys species belong to an undescribed family within the order Hypocreales. Tang et al. (2007) showed that both Melanopsamma pomiformis and Myrothecium inundatum are allied to the Hypocreales and Fungal Diversity accepted them in Niessliaceae and Bionectriaceae, respectively. Kirk et al. (2008) and Index Fungorum (2014) considered it as incertae sedis. Whitton et al. (2012) discovered that Ornatispora K.D. Hyde, Goh, Joanne E. Taylor, J. Fröhl. has Stachybotrys asexual states and reported asexual-sexual relationships of two species, Ornatispora nepalensis Whitton, K.D. Hyde & McKenzie with a Stachybotrys sp. asexual state and Ornatispora novae-zealandiae Whitton, K.D. Hyde & McKenzie with a Stachybotrys freycinetiae asexual state. It is rather significant to discover another sexual genus linked to Stachybotrys. The finding of Whitton et al. (2012) will assist in determining the taxonomic placement of Stachybotrys. Lechat et al. (2013) described S. oleronensis as a new species with only the sexual stage based on sequence analysis. They discovered Stachybotrys oleronensis resembled Nectriella funicola by the intertwined hyphal wall, which belonged to Bionectriaceae (Lechat et al. 2013). Since Melanopsamma and its Stachybotrys and Custingophora asexual morphs have been placed into different families, Chaetosphaeriaceae, Niessliaceae or an undescribed family in Hypocreales, it is necessary to conduct a phylogenetic and morphological study of Stachybotrys, its sexual Melanopsamma, and morphologically and phylogenetically allied genera using multiple genes/regions to clarify the contradictory treatments and to determine their correct higher level placement. History of molecular phylogenetic development of Stachybotrys and allied genera Morphological characters have been used for over 300 years to identify and classify fungi. Although relatively easy to observe and record, and to assist in the differentiation process (Talbot 1971), morphological characters may not reflect phylogenetic relationships as many characters are subject to plasticity, parallelism, and reversal (homoplasy) (Judd et al. 2002). Morphological species concepts, in many instances, have either over-estimated or under-estimated speciation events (Jeewon et al. 2004). Mayden (1997) characterized fungal species concepts as either theoretical or operational. He reviewed the only primarily theoretical species concept, the Evolutionary Species Concept (ESC). However, when it came to identifying fungal species, the ESC was not helpful because it has no recognition criteria (Taylor et al. 2000). Taylor’s research group further pointed out that three operational species concepts (Morphological Species Concept, Phylogenetic Species Concept and Biological Species Concept) did specify criteria for recognizing species. Among them, “Phylogenetic Species Concept (PSC) performed best because, once progeny evolutionary species have formed from an ancestor, change in gene sequences occurs and can be recognized before changes have occurred in mating behavior or morphology” (Taylor et al. 2000). For Stachybotrys, the main method to delimite species was still on the basis of Morphological Species Concept. Gene sequence data are resolving taxonomic problems at the species or higher taxonomic levels (Shenoy et al. 2007, 2010; Hyde et al. 2013; Wu et al. 2014). Haugland and Heckman (1998) first identified nuclear ribosomal DNA sequences as the specific PCR primers for detection of the toxigenic fungal species S. chartarum. Then 26 internal transcribed spacers (ITS) sequences representing 11 Stachybotrys/ Memnoniella species and one unidentified Stachybotrys strain (it was later described as M. longistipitata) were analyzed to evaluate the morphological features for Memnoniella species identification (Haugland et al. 2001). The results indicated that Memnoniella is synonymous with Stachybotrys, and Memnoniella echinata and M. subsimplex should be recognized as species of Stachybotrys, bringing to 13 the total number of morphologically recognized Stachybotrys species. Taylor et al. (2000) stated that one gene genealogy could not be sufficient to recognize phylogenetic species. Thus, using three polymorphic protein coding loci [trichodiene synthase 5 fragment (tri5), beta-tubulin 1 fragment (tub1) and chitin synthase 1 fragment (chs1)], Cruse et al. (2002) inferred that two distinct phylogenetic species exist within the single described morphological species of S. chartarum. Andersen et al. (2003) described this new species as S. chlorohalonata and recognized two chemotypes of S. chartarum on the basis of the three above fragments and phenotypic analyses. Koster et al. (2003) used molecular markers based on β-tubulin, calmodulin, elongation factor-1 alpha (EF1-α), and tri5 genes, as well as ITS rDNA to investigate genetic variation among 52 morphologically and geographically diverse, indoor and outdoor isolates of S. chartarum sensu lato. This proved concordant in dividing all isolates into two strongly supported clades. However, they were not able to provide names for the two species represented by the two clades due to the absence of ex-type cultures and unavailability of the type specimens for examination. Clearly, Stachybotrys chartarum is a species complex. A quantitative polymerase chain reaction (QPCR) test was conducted to re-examine the holotype of S. cylindrospora and a new taxon, S. eucylindrospora was described (Li 2007). Multi-gene phylogeny for Stachybotrys implied tris5 was highly conserved at the amino acid level suggesting that identity at variable sites, among otherwise divergent taxa might be the results of chance events and tri5 was a poor marker choice for phylogenetic reconstruction at the genus level (Koster et al. 2009). Jie et al. (2013) also described a new Stachybotrys species from soil by morphology and analyses of three gene regions (ITS, EF1-α, RPB2). The results also indicated that single ITS sequences were obviously insufficient in species identification. Fungal Diversity Despite the economic importance of Stachybotrys, higher level phylogenetic relationships of the genus have seldom been investigated (Castlebury et al. 2004). Castlebury et al. (2004) reported that the hypocrealean dataset, Stachybotrys/ Melanopsamma, Myrothecium, and Peethambara/ Didymostilbe formed a strongly supported, previously undiscovered sister lineage to all other families currently accepted in the Hypocreales—Bionectriaceae, Clavicipitaceae, Hypocreaceae, Nectriaceae, Niessliaceae—based on six gene analyses [small subunit ribosomal DNA (nrSSU) and the nuclear encoded large subunit ribosomal DNA (nrLSU), RNA polymerase II largest subunit (RPB2), EF1-α and mitochondrial ATP synthase 6 (ATP6)]. Stachybotrys and Myrothecium were well-supported as monophyletic asexual genera and the relationship of Melanopsamma pomiformis with Stachybotrys was confirmed. Importantly, this relationship provides the clue for resolving the taxonomic position of Melanopsamma (and hence Stachybotrys) in the Hypocreales. However, only three Stachybotrys species (S. chartarum, S. echinata and S. subsimplex) were included, and S. echinata (UAMH 6594) is designated as an epitype isolate (Haugland et al. 2001), but we believe their decision is problematic. Schroers et al. (2005) used LSU rDNA, beta-tubulin, and ITS rDNA sequences to analyze hypocrealean taxa, which revealed that phylogenetic relationships of Myrothecium inundatum, Peethambara sundara, S. chartarum, S. echinata and S. bisbyi was still problematic. The above research suggests that two key questions are not resolved, viz. “Is Stachybotrys monophyletic or polyphyletic, and how does this genus affront One fungus = One Name?” Phylogenetic studies on Stachybotrys Stachybotrys species at the species level The ITS region is sequenced DNA region and is considered to be a universal DNA barcode marker for Fungi (Nilsson et al. 2014; Peay et al. 2008; Schoch et al. 2012). It has typically been most useful for molecular systematics at the species level, and even within species (e.g., to identify geographic races). For the genus Stachybotrys, 270 ITS sequences were found in GenBank covering 17 Stachybotrys species: S. chartarum, S. chlorohalonata, S. dichroa, S. echinata, S. elegans, S. eucylindrospora, S. kampalensis, S. longispora, S. microspora, S. nephrospora, S. oenanthes, S. parvispora, S. subcylindrospora, S. subreniformis, S. subsimplex, S. theobromae and S. zeae. In addition, sequences were available for Melanopsamma pomiformis, four uncultured Stachybotrys isolates and 17 isolates named as Stachybotrys sp. from NCBI website. Additionally, we have obtained 122 Stachybotrys strains with the help of Prof. TianYu Zhang’s research group, but unfortunately few are ex-type or epitype strains. We have built up the parsimonious tree (Fig. 1) based on GenBank sequences plus our 22 ITS sequences with Cordyceps heteropoda as outgroup. The Stachybotrys/Melanopsamma-Memnoniella isolates clustered together with a high bootstrap value (100 %), meaning that we have some proof of a monophyletic status for Stachybotrys and Memnoniella. Stachybotrys eucylindrospora (ATCC 18851, ex-type) displayed a distant phylogenetic relationship with other Stachybotrys spp. The large clade obtained includes at least seven species-groups (chartarum, dichroa, elegans, echinata, microspora, nephrospora, parvispora and subsimplex). At least one undescribed Stachybotrys isolate (LA227) and S. longistipitata (ATCC 22699) had distinct relationships with other Stachybotrys-Memnoniella spp. Chartarum species-group This grouping obtained a 99 % bootstrap value with S. chartarum as representative. Many strains belonged to S. chartarum, as a cellulolytic saprobe with a worldwide distribution (Li and Yang 2005). However, according to previous experience (Andersen et al. 2003; Cruse et al. 2002; Koster et al. 2003, 2009), S. chartarum isolates (Fig. 1) certainly include some divergent lineages based on multi-gene analysis, and some taxa might be synonyms of S. chartarum, just like S. atra. Thus, for S. chartarum it is vital to nominate an epitype. S. chartarum showed a close relationship with the epitype of S. microspora, but we believe the latter is a problematic species, because the definition of Jong and Davis (1976) deviated significantly from the original description of B.L. Mathur & Sankhla, and eight adjoined Stachybotrys isolates. We discovered that some of our strains were identified as S. breviuscula, S. eucylindrospora, S. subreniformis and S. yunnanensis, but did not obtain sequence support. Four strains (HGUP 0143, HGUP0155, HGUP0120 and HGUP0201) clustered together with 97 % bootstrap support as a sister taxon to S. chartarum. Microspora species-group Two isolates of S. microspora (ATCC 18852 and HGUP0234) were placed together with 100 % bootstrap support, and formed an independent branch. According to Jong and Davis (1976), ATCC 18852 in full agreement with the type material is nominated as the epitype of S. microspora. Dichroa species-group This clade had high bootstrap support (96 %). ATCC 22844 is the type culture of S. oenanthes and we propose ATCC 18913 as epitype for S. dichroa, whose morphology is consistent with (See comment of Jong and Davis 1976). Fungal Diversity Fig. 1 The ITS phylogenetic tree of Stachybotrys spp. based on MP method with Cordyceps heteropoda (IFO 33060) as outgroup. Bootstrap values ≥ 50% are indicated above the branches. Ex-type strains were labeled with “T” after strain number. Epi-type strains were labeled with “epi” after strain number (Table S1) Echinata species-group S. subsimplex, in future which might form a separate species-group. ATCC 18839 is the type of S. nephrospora (Jong and Davis 1976). Stachybotrys reniformis is probably a synonym of S. nephrospora but, unfortunately, no culture is available for S. reniformis (Jong and Davis 1976). However, we cannot determine the taxonomic position of SAP 155 as only an ITS sequence is available and no other valuable information could be used. This clade had credible bootstrap support (78 %), and included Memnoniella echinata (DAOM 173162), S. kampalensis (UAMH 7746) and one undescribed strain (LA227). They displayed an obvious phylogenetic distinction. Subsimplex species-group The subsimplex species-group of four S. subsimplex strains, S. nephrospora (ATCC 18839) and SAP155 was obtained with a moderate bootstrap value (57 %). Jong and Davis (1976) found ATCC 18838 and ATCC 22700 to fit with the description of Memnoniella subsimplex (Deighton 1960), but they did not see a Stachybotrys-like phase in these cultures, suggesting that clarifying the relationship between Memnoniella subsimplex and Stachybotrys was more perplexing. However, this clade is obviously a Stachybotrys. Among them, ATCC 18839 and SAP 155 formed a branch with 86 % bootstrap support as the sister branch of Parvispora species-group This group comprised six isolates (ATCC 18905, ATCC 18873, ATCC 32451, ICMP 15920, NRRL 54531, ATCC 18877). Obviously, ICMP 15920 is not S. chartarum and may need to be described as a new species. We propose ATCC 18905 as the epitype of S. theobromae. Jong and Davis (1976) pointed out that both ATCC 18877 fit Hughes’ original description of S. parvispora, thus ATCC 18877 is designated as epitype. The phylogenetic positions of Fungal Diversity S. longispora and Me. pomiformis were resolved in this parsimonious tree. ITS (data not shown), which has been supported by previous studies (Liu and Hall 2004; Reeb et al. 2004; Geiser et al. 2004). Elegans species-group Stachybotrys at the generic level The bootstrap support value for the elegans species-group is 99 %. Three S. elegans strains (ATCC 18825, ATCC 66760 and DAOM 225565) displayed phylogenetic difference, which may indicate S. elegans is a species complex. Stachybotrys bisbyi, synonym of S. elegans in this paper, was also accommodated in this clade, but did not show a close relationship to the others. KUC 5202 should be a new taxon after complementing morphological observation. In order to compare the applied foreground of different DNA markers, we also built the parsimonious trees of RPB2 and EF1-α using the Stachybotrys sequences in NCBI (Figs. 2 and 3). RPB2 and EF1-α genes hold more parsimonious information than Fig. 2 One of the four equally most parsimonious trees of the analyzed EF1-α region. Bootstrap support values <50 % are not shown. The tree is rooted with Myrothecium verrucaria. Ex-type strains were labeled with “T” after strain number The sexual state for most Stachybotrys spp. has never been reported. As asexually reproducing fungi they are, at least theoretically, a clonal organism. An increasing body of evidence suggests that despite the lack of a sexual state, several fungi known to reproduce only asexually appear to be undergoing the equivalent of sexual reproduction resulting in genetically diverse populations (Taylor et al. 2000, 1999; Bidochka and De Koning 2001). In addition, many asexually reproducing fungi are derived from within groups that include sexually reproducing species (O’Donnell et al. 1998; Chaverri et al. 2003). Only Fungal Diversity Fig. 3 One of the three equally most parsimonious trees of the analyzed RPB2 region. Bootstrap support values <50 % are not shown. The tree is rooted with Myrothecium verrucaria (ATCC 9095). Ex-type strains were labeled with “T” after strain number four species of Stachybotrys have been linked to a sexual state. Melanopsamma pomiformis was reported as the sexual state of S. albipes (Booth 1957). Whitton et al. (2012) described two new species, Ornatispora nepalensis Fungal Diversity and O. novae-zealandiae. The former was said to have an asexual state belonging to Stachybotrys, and the latter was stated to be the sexual state of S. freycinetiae. Ornatispora is thought to be a member of Niessliaceae, but this has not been confirmed by molecular data (Hyde et al. 1999). Some fungi with pleomorphic life-cycles still have two names despite more than 20 years of molecular phylogenetics that have shown how to merge the two systems of classification, the asexual “Deuteromycota”, and the sexual “Eumycota” (Taylor 2011). To evaluate the genetic position of Stachybotrys in Hypocreales, we supplemented LSU rDNA sequences of 14 isolates distributed in four species-groups and downloaded 11 Stachybotrys sequences from NCBI, although some of the latter have only short fragments (approximately 200–300 bp). Twenty-five Stachybotrys sequences were placed in a backbone parsimonious tree with Verticillium dahliae (ATCC 16535) as the outgroup taxon. Stachybotrys grouped with Myrothecium leucotrichum, Didymostilbe echinofibrosa and Peethambara spirostriata Fig. 4 One of the nine equally most parsimonious trees of the analyzed LSU region (140 of the 694 characters were parsimony informative). Bootstrap support values <50 % are not shown. The tree is rooted with Verticillium dahliae. Ex-type strains were labeled with “T” after strain number Fungal Diversity Fig. 5 Stachybotrys albipes (Jong and Davis 1976) with 55 % bootstrap support (Fig. 4), which was consistent with the studies of Castlebury et al. (2004) and Schroers et al. (2005). If ATCC 32888 is removed from Stachybotrys group, the genus can be considered as monophyly (Fig. 4), this however, goes against the ITS result (Fig. 1). In Fig. 4, the phylogenetic clade of Stachybotrys group showed a close relationship with Didymostilbe echinofibrosa, Myrothecium inundatun and Peethambara sundara, but with a moderate bootstrap support (52 %), all of which belong to Bionectriaceae (Schroers et al. 2005). Lechat et al. (2013) also place their new species, Stachybotrys oleronensis (sexual state only) close to Myrothecium and say it is very similar to Nectriella (Bionectriaceae), although no species of Nectriella have been sequenced (Lechat et al. 2013). According to Rossman et al. (2001), Myrothecium inundatum, the type species of Myrothecium, Peethambara sundara (asexual Didymostilbe sundara), Albosynnema elegans, and Didymostilbe echinofibrosa are all linked to the Bionectriaceae according to morphological evidence, and formed a paraphyletic assemblage at the base of a clade comprising taxa of the Hypocreaceae, Clavicipitaceae, and Bionectriaceae based on LSU rDNA sequence data. This might provide proof for the taxonomic position of Stachybotrys group in Bionectriaceae. Interestingly, Schroers et al. (2005) included S. bisbyi, S. chartarum, S. echinata, Myrothecium inundatun, and Peethambara sundara in an independent clade supported by 60 % statistic bootstrap in the root of hypocrealean taxa tree based on LSU rDNA gene region. However, the phylogenetic tree of combined nrSSU and nrLSU gene regions (Castlebury et al. 2004) was basically consistent with our study. Crous et al. (2014) further proposed the new family Stachybotriaceae to accommodate Myrothecium, Peethambara and Stachybotrys since these genera formed an independent lineage distinct from other families in the Hypocreales (Castlebury et al. 2004; Crous et al. 2014; Summerbell et al. 2011). We believed the new family was a better choice for us to explain the taxonomic confusion of Stachybotrys and related genera. Fungal Diversity Fig. 6 Stachybotrys bambusicola (Rifai 1964) Taxonomic confusion and search for suitable genes for barcoding The morphological descriptions, together with further observations on the described species, represent a valuable and comprehensive source of information, which is still extensively used today. Nevertheless, relying solely on morphological characters in the identification process can be controversial. This is true because of the scarcity and plasticity of discriminatory, yet easily accessible, morphological characters (Begerow et al. 2010). Therefore, molecular tools were readily embraced by the mycological community when they became available. We are interested in how to rapidly identify Stachybotrys strains. DNA barcoding uses a short genetic marker in an organism’s DNA to assign it to a particular species (Hebert et al. 2003). It is different from molecular phylogeny in that the main goal is to identify an unknown sample in terms of a known classification, rather than to determine its phylogeny. Another function of DNA barcoding for fungi is to link the sample or isolate to its voucher information so that the isolate and the name it bears can be re-examined, if necessary. It will provide a mechanism with which erroneous information deposited to a database (eg, GenBank 2014) can be corrected and/or annotated in the future. For fungi, including Stachybotrys, ITS is used almost universally and was selected as the standard DNA barcode region, over which we have nearly reached a consensus (Schoch et al. 2012; Nilsson et al. 2014). However, it is not a good region for phylogenetics as ITS variation does not necessarily match phylogenetic species (Rossman 2007). Rossman (2007) suggests that ITS can be used as the first step or first key followed by sequencing of a second gene for a precise identification. Such an approach has been used successfully in the TrichoKey DNA barcoding system. Fungal Diversity Fig. 7 Stachybotrys breviuscula (McKenzie 1991) Koster et al. (2009) used trichodiene synthase (tri5) (trichothecene biosynthetic gene), rpb2H, tef1H and SSU genes to study phylogeny of Stachybotrys. Results showed that tri5 gene lacked resolution in their tree and concluded that it is a poor marker choice for phylogenetic reconstruction at the genus level. Nielsen et al. (2002) found that some strains of Stachybotrys chartarum do not produce macrocyclic trichothecene, and it is, therefore, unlikely that tri5 can serve as a good marker for studying phylogenetic relationship of Stachybotrys including Memnoniella at species level. Unfortunately, this means that the tri5 gene cannot by selected as the secondary DNA maker for barcoding analyses. The fact that few sequences are available, especially those originating from ex-type or epitype strains is a great limitation in evaluating the utility of other gene markers for precise identification of Stachybotrys species. We also sequenced the tef1 of all our Stachybotrys isolates (data not shown), and found that tef1 marker has more informative characters than the ITS region. This gene has the potential of being the second barcoding marker for this genus. However, its use must be determined only after an evaluation of the backbone tree of Stachybotrys by multi-gene analyses. Taxonomy of Stachybotrys Stachybotrys Corda, Icon. fung. (Prague) 1: 21 (1837) Synonyms: Melanopsamma Niessl, Verh. nat. Ver. Brünn 14: 200 (1876) Memnoniella Höhn., Zentbl. Bakt. Parasit Kde, Abt. II 60: 16 (1923) [1924] Ornatispora K.D. Hyde, Goh, Joanne E. Taylor & J. Fröhl., Mycol. Res. 103(11): 1432 (1999) Sexual morph: Ascomata superficial, globose, collabent when dry, black, coriaceous, lacking or covered in numerous setae, papillate. Papilla short, beaklike, black, shiny, periphysate. Sterile tissue filiform, aseptate, flexuose, deliquescing in dried material. Asci 8-spored, clavate, pedicellate, thin-walled, unitunicate, lacking an apical apparatus, deliquescent at maturity. Ascospores 2-3-seriate, ellipsoidal, 1-septate, hyaline, verrucose and surrounded by a mucilaginous sheath (Hyde et al. 1999). Asexual morph: Vegetative hyphae usually hyaline. Conidiophores macronematous, simple or cymosely branched, with apical clusters of several Fungal Diversity Fig. 8 Stachybotrys chartarum (Jong and Davis 1976) ellipsoidal or subclavate phialides formed in succession. Conidia and phialides hyaline or pigmented. Conidia released in basipetal succession through an opening in the rounded phialide apex which has hardly prominent collarettes, held together in slimy drops or in chains, one-celled, ellipsoidal, cylindrical, reniform or fusiform, mostly ornamented, in some species smooth-walled, pigmented or hyaline. New conidia arise after the previous ones are mature and have been released from the phialide neck. Notes: Stachybotrys is characterized by macronematous, mononematous, single or branched conidiophores, with discrete phialidic conidiogenous cells, and 0-septate conidia, produced in a slimy mass, usually being dark in color (Jong and Davis 1976; Mercado-Sierra et al. 1997). We accept a wider generic concept to accommodate some ambiguous genera, viz. conidia in a slimy mass and chains both belong to Stachybotrys. Therefore, we agree with Smith (1962) and Carmichael et al. (1980) to combine Stachybotrys and Memnoniella under the older name of Stachybotrys. The phylogenetic analysis based on ITS gene region (Fig. 2) also supported the opinion that Melanopsamma is the sexual state of Stachybotrys. Species of the sexual genera Melanopsamma and Ornatispora have been linked to Stachybotrys while Stachybotrys oleronensis is only known in it sexual state (Lechat et al. 2013). Melanopsamma is a unitunicate ascomycete genus (Wang 2011) and Kirk et al. (2008) estimated there Fungal Diversity Fig. 9 Stachybotrys chlorohalonata (Andersen and Thrane 2003) are 41 species in the genus. The asexual state of Melanopsamma pomiformis is S. albipes, which is the type species of Melanopsamma. Thus, Melanopsamma is a later synonym of Stachybotrys. The link between S. albipes and Melanopsamma pomiformis has been discussed by Castlebury et al. (2004) and is confirmed in this paper by ITS sequence analysis (Fig. 2). Lechat et al. (2013) observed only a sexual morph on leaf of Iris pseudacorus, but ITS sequence analysis indicated it belonged to Stachybotrys and the species was introduced as S. oleronensis. Ornatispora was introduced by Hyde et al. (1999) with O. palmicola K.D. Hyde, Goh, Joanne E. Taylor & J. Fröhl. as the type species. Unfortunately there is no molecular data available for the genus, but O. taiwanensis which was also introduced in the same paper is illustrated with a Stachybotrys asexual state. The characters of Ornatispora are similar to those of Melanopsamma and are therefore also synonymous with Stachybotrys. Ornatispora gamsii was reported to have a Didymostilbe aurantiospora asexual state by Hyde et al. (1999), but this is not illustrated and was based on its close association on the host and needs confirmation. Thus, we accepted Ornatispora as the synonym of Stachybotrys, and all seven Ornatispora species are transferred to Stachybotrys. Why choose Stachybotrys over other names? The proposal to build a system of adopting one name for each fungal species and end the system of dual nomenclature was adopted at the Eighteenth International Botanical Congress Melbourne, Australia, July 2011 and by the International Code of Nomenclature for algae, fungi, and plants (Melbourne Code) (Hibbett and Taylor 2013). Thus, we must choose between Stachybotrys, Memnoniella, Melanopsamma and Ornatispora to name the genus. When selecting the generic name we must adopt the principle of priority of publication (Hawksworth et al. 2011), unless we can make a case for using the younger name. Stachybotrys is both the oldest better known and commonly used name and, therefore, should be used for the genus. Thus, Memnoniella, Melanopsamma and Ornatispora are designated as synonyms. Type species: Stachybotrys chartarum (Ehrenb.) S. Hughes, Can. J. Bot. 36: 812, 1958 Fungal Diversity Fig. 10 Stachybotrys cordylines (Pinruan et al. 2004) Key to the genus Stachybotrys Note: Conidia have not been recorded for Stachybotrys frondicola, S. oleronensis, S. punctata, and S. taiwanensis and hence these species are not included in the key. The description of the asexual stage of S. palmicola is rudimentary (conidiophores ca. 200×5–6 μm; conidia 10×4 μm) and it is also not included in the key. Stachybotrys gamsii, reported to have a Didymostilbe aurantiospora asexual state, is also not included. 1. Conidiophores synnematous..........................................2 1. Conidiophores mononematous......................................3 2. Conidia subglobose, 7–12 μm diam., distal end black and verrucose, basal end pale and smooth......S. leprosa 2. Conidia globose to slightly angular, 3–5 μm diam.......................................................S. stilboidea 3. Conidia catenate, produced in dry chains, sometimes also non-catenate in slimy heads...................................4 3. Conidia non-catenate, aggregated in slimy heads.........11 4. Conidia smooth..............................................................5 4. Conidia roughened.........................................................6 5. Conidia catenate, globose or sometimes hemiglobose, 4– 6 μm diam.....................................................S. levispora 5. Conidia catenate, ellipsoid, 5.5–6.5 × 2.5–3.8 μm diam........................................................S. mohanramii 6. Conidia bimorphic (catenate and non-catenate).............7 6. Conidia monomorphic (catenate), globose to subglobose......................................................................8 7. Catenate conidia globose, 5–5.5 μm diam.; non-catenate conidia oblong, 7–10×3.5–5 μm.......................S. zuckii 7. Catenate conidia globose, 5.8–8.5×6.3–8.3 μm; noncatenate conidia oblong or ovoid, 10.6–11.9×4.8– 5.7 μm...................................................S. longistipitata 8. Conidia thick-walled, 5–6 μm diam.; conidiophores regularly branched, covered in part by granules.....................................................S. indicoides 8. Conidia thin-walled, conidiophores not regularly branched.......................................................................9 9. Conidiophores 38–50 μm long; phialides 6.3–14.7× 4.2–5.2 μm; conidia verrucose, 4.4–6.8 μm diam...........................................................S. zingiberis 9. Conidiophores more than 50 μm long.......................10 Fungal Diversity Fig. 11 Stachybotrys dichroa (Jong and Davis 1976) 10. Conidia verrucose, 4–6 μm diam.; conidiophores 50– 100 μm long, covered in part by granules; phialides 7– 9×3–5 μm....................................................S. echinata 10. Conidia verrucose 6–9 μm diam.; conidiophores commonly 100–140 μm long, smooth or minutely verruculose..............................................S. subsimplex 11. Conidia reniform (sometimes very slightly and appearing other shapes in face view) or curved........12 11. Conidia not reniform or curved..................................21 12. Conidia tightly curled or tightly reniform (almost globose in face view), with dark outer zone, (11–)13– 15(−15.5)×(10.5–)12–14 μm...................S. nephrodes 12. Conidia not tightly curled...........................................13 13. Conidiophores regularly proliferating 2–4 times through apex, conidia 5.5–7 × 4–5 μm, smooth......................................................S. proliferata 13. Conidiophores not proliferating.................................14 14. Conidiophores sinuous or even circinate, repeatedly branched; conidia 8–12 × 6–7 μm, verrucose..............................................S. sinuatophora 14. Conidiophores not sinuous, unbranched or branched...............................................................15 15. Conidia smooth, mostly less than 5 μm wide and less than 7 μm long...........................................................16 15. Conidia smooth or roughened, mostly more than 5 μm wide and more than 7 μm long..................................17 Fungal Diversity Fig. 12 Stachybotrys echinata (Jong and Davis 1976) 1 5 .C o n i d i a 7 – 9 × 3 – 5 μ m , c o n t a i n i n g 1 – 2 o il droplets.......................................................S. aloeticola 16. Conidia 4.5–7×3–4.5 μm; phialides smooth, 7–11× 4.5–6 μm; conidiophores smooth, unbranched......................................“S. renisporoides” 16. Conidia 5.2–7×3.5–5.2 μm; phialides minutely verrucose, 7.5–9.3×3–4.5 μm; conidiophores minutely verrucose, branched.........................................S. renispora 17. Conidia often ellipsoid, globose or only slightly reniform.....................................................................18 17. Conidia more obviously reniform..............................19 18. Conidia often ellipsoid, smooth or verrucose, 9–12× 4.5–8 μm; phialides smooth or verruculose, 12–21×4– 7 μm; conidiophores smooth, 120–180 × 6– 9 μm..........................................................S. oenanthes 18. Conidia often globose, verruculose, 6–9.5 × 4.5– 7.5 μm; phialides smooth, brown, 8–11.5 × 4.5– 19. 19. 20. 20. 21. 6 μm; conidiophores smooth, hyaline to dark brown, 48–98×5–7.5 μm..............................S. subreniformis Conidia smooth, 8–11×4.5–6 μm; phialides smooth, 10–12×5–6 μm; conidiophores smooth or roughened (especially near apex), branched..........S. nephrospora Conidia roughened....................................................20 Conidia tuberculate, often globose, (9–)10–12×5– 6(−7) μm, usually biguttulate; phialides smooth, hyaline, 10–13×4–6 μm; conidiophores roughened towards apex, hyaline, (70–)100–130(−200) × 3– 4.5 μm......................................................S. reniformis Conidia coarsely ornamented, 10–13.5×6–9.5 μm; phialides smooth, pale grey or pale brown, 10–14× 5–6 μm; conidiophores smooth, subhyaline near base becoming slightly darker towards apex, 86–137 μm long.....................................................S. reniverrucosa Conidiophores regularly sympodially branched.......22 Fungal Diversity Fig. 13 Stachybotrys eucylindrospora (Jong and Davis 1976) 21. Conidiophores not regularly sympodially branched (but may be abundantly branched)...........................25 2 2 . C o n i di a s m o o t h , s ub g l o b o s e or e l l i p s o i d , 7–9 × 6–7 μm; conidiophore branches 6–8 μm long...............................................................S. ramosa 22. Conidia roughened..................................................23 23. Conidia ovoid, ellipsoid or subglobose, tuberculate, 9– 12.5×7.5–10 μm; conidiophore branches 60–100 μm long.......................................................S. xigazenensis 23. Conidia globose to subglobose.................................24 24. Conidia globose, 4.5–8 μm diam., roughened; phialides in groups of 2–5 at apex of each branch; conidiophore branches 23–46 μm long, 2.5–3.5 μm wide near base tapering to 1–1.5 μm at apex..............................................................S. globosa 24. Conidia globose to subglobose, 6–9 μm diam., tuberculate; phialides in groups of 6–8 at apex of each branch; conidiophore branches 60–80 μm long, 4– 5 μm wide near base.............................S. jiangziensis 25. Conidia when mature globose or subglobose (other shapes may be present).............................................26 25. Conidia of other shapes.............................................32 26. Conidia hyaline, globose to subglobose and 9–12.5 μm diam., or broadly ellipsoid to limoniform and 10.5– 16×8–12 μm; conidiophores up to 120 μm long, (3.5–)5–7 μm wide; phialides hyaline, 9.5–16×3– 5 μm.........................................................S. mexicanus 26. Conidia pigmented....................................................27 27. Conidia more than 10 μm diam................................28 27. Conidia mostly less than 10 μm diam......................30 28. Conidia 11–12 μm diam., ornamented with irregular, prominent ridges; conidiophores abundantly branched.............................................S. sphaerospora 28. Conidia more than 15 μm diam.................................29 29. Conidia 16–18 μm diam., verruculose............S. crassa 29. Conidia (15.4–)21–25.2(−28) μm diam., tuberculate.................................................S. nilagirica 30. Conidia 5–6 μm diam. (young conidia ellipsoid or pyriform), coarsely roughened, base not papillate; conidiophores sometimes branched; phialides 8–9×4– 5 μm........................................................S. microspora 30. Conidia papillate at base............................................31 31. Conidia 6–8 μm diam., inconspicuously verruculose; conidiophores unbranched; phialides 10–15 × 4.6– 6 μm..................................................S. ruwenzoriensis 31. Conidia 7.5–10.5 μm diam., verrucose; conidiophores sometimes branched; phialides 10–15 × 4– 6 μm.................................................................S. kapiti 32. Conidia hyaline, or pink in mass...............................33 32. Conidia pigmented.....................................................36 33. Conidia verrucose, ellipsoid, truncate at each end, hyaline, 10–15×5–7.5 μm; phialides hyaline, 11– 12.5×6–7.5 μm; conidiophores brown, apical cell hyaline, (80–)110–230×6.3–10 μm.................S. palmae 33. Conidia smooth..........................................................34 34. Conidiophores not proliferating, hyaline, up to 200 μm long, 3–4 μm wide; conidia limoniform to fusiform, hyaline, 1–3 guttulate, 8–14 × 6– Fungal Diversity Fig. 14 Stachybotrys freycinetiae (McKenzie 1991) 9 μm; phialides hyaline, 10–17 × 4– 6 μm......................................................S. elegans 34. Conidiophores may proliferate several times through apical cell, dark reddish brown, paler towards apex, apical cell hyaline.......................................................35 35. Conidia ellipsoid to subfusiform, 13.8–18.4 × 4– 5.8 μm, containing 2 large oil globules; phialides hyaline, 9–12.5×4.5–5.5 μm; conidiophores up to 275 μm high............................................S. palmijunci 35. Conidia obovoid, rarely ellipsoid, slightly flattened base, 10–15.5 × 6.5–8 μm; phialides usually hyaline, occasionally pale brown, 17– 23 × 4.5–7 μm; conidiophores 125–396 μm high................................................S. bambusicola 36. Mature conidia ornamented with striations...............37 36. Mature conidia without obvious striations, variously roughened or smooth.................................................39 37. Conidia with diagonal striations but becoming irregular with age, cylindrical or ellipsoid, sometimes slightly constricted at centre, (12.5–)15– 17.5(−19)×(6–)6.5–7.5(−8) μm............S. thaxteri 37. Conidia with longitudinal striations..........................38 38. Conidia cylindrical, (10.3–)12.8–16(−18.5)×(2.5–)3.4– 5.5(−6.6) μm.....................................S. eucylindrospora 38. Conidia ellipsoid to fusiform, 12–13 × 5– 5.5 μm............................................................S. virgata 39. Conidia 20–28×14–17 μm, ellipsoid, papillate at base, smooth, black; conidiophores hyaline, smooth, Fungal Diversity Fig. 15 Stachybotrys globosa (Misra and Srivastava 1982) branched, up to 200 μm long, 4–6 μm wide; phialides 20–27×6–8 μm.....................................S. theobromae 39. Conidia less than 14 μm wide....................................40 40. Conidia variable in shape and size, ovoid, ellipsoid, oblong, globose or subglobose, 4–20 × 3–13 μm, smooth or coarsely roughened....................S. variabilis 40. Conidia more regular in shape and size......................41 41. Conidia mostly more than 8 μm wide.........................42 41. Conidia mostly less than 8 μm wide..........................43 42. Conidia 10–15×9.5–11(−12.5) μm, ellipsoid to obovoid, tuberculate; conidiophores hyaline, pale brown towards apex, unbranched, 80–235 × 7– 1 2 . 5 μ m ; p h i a l i d e s 11 – 1 6 × 6 – 8 μ m , p a l e brown....................................................S. verrucispora 42. Conidia 14.5–19×8–11.2 μm, ellipsoid to broadly ellipsoid, base rounded or with a broad truncate papilla, verrucose; conidiophores hyaline, pale grey towards apex, unbranched, 54–75×4.5–7 μm; phialides 10–15 × 4.5–6 μm, pale grey or darker at apex.............................................................S. waitakere 43. Mature conidia smooth...............................................44 43. Mature conidia roughened..........................................54 44. Conidia mainly cylindrical or subcylindrical.............45 44. Conidia of other shapes..............................................47 Fungal Diversity Fig. 16 Stachybotrys guttulispora (Muhsin and Al-Helfi 1981) 45. Conidia cylindrical, 8.8–12×2–2.4 μm, olivaceous; conidiophores hyaline, usually unbranched, 56–76× 4–5 μm; phialides 8–10 × 3.2–4.6 μm, subhyaline.................................................S. longispora 45. Conidia at least 2.5 μm wide.....................................46 46. Conidia ellipsoid, 6–11×2.5–4 μm, smooth or verrucose; conidiophores hyaline to pale olivaceous grey, sometimes branched, 41–69×3–4 μm; phialides 7.5– 12×3–4.5 μm............................................S. pallescens 46. Conidia ellipsoid, 7–11 × (2.5–)3.5–5 μm, usually smooth, sometimes rough; conidiophores greenish becoming grey to black, unbranched, 70–120 × 2.2– 4.5 μm; phialides 8–13×3.2–5 μm.......S. yunnanensis 47. Conidia oval, 3.5–6×3–3.5 μm, dark brown; conidiophores hyaline, up to 160 μm long, 4–7 μm wide; phialides 7–10×3–4 μm...........................S. parvispora 47. Conidia mostly more than 6 μm long........................48 48. Conidia often fusiform, with truncate base................49 48. Conidia not fusiform...................................................51 49. Conidiophores branched, subhyaline to pale brown, up to 68 μm long, 3.5 μm; conidia 8.5–15.5×3.5–5 μm, dark brown; phialides 8.5–12 × 5–7 μm, pale brown................................................S. thermotolerans 49. Conidiophores unbranched........................................50 50. Conidia 6–9×3–4 μm; conidiophores subhyaline to pale brown, up to 60 μm long, 2–3 μm wide; phialides 8–13×3–4 μm, pale brown...................S. sansevieriae 50. Conidia 7–10.2×2.5–3.5 μm, grey or dark grey with truncate base; conidiophores hyaline, up to 50 μm long, 2.5–3.2 μm wide; phialides 7–8 × 3.5–4 μm, hyaline.....................................................S. havanensis 51. Conidia containing 2 obvious guttules.......................52 51. Conidia without obvious guttules..............................53 52. Conidia 9–12 ×3.5–5 μm, olivaceous or greenish brown; conidiophores pale olivaceous, unbranched, rough over whole surface, 60–90×3–5 μm; phialides 10–15×3–4.5 μm, subhyaline..............S. guttulispora 52. Conidia 7–10×3–5 μm, pale grey; conidiophores subhyaline becoming greyish brown towards apex, unbranched, rough over whole surface, 23–75×2– 3 μm; phialides 8.5–14.5×3–4.5 μm, pale olivegrey.............................................................S. terrestris 53. Conidia 6–11×4.5–7 μm, greenish or greyish brown; conidiophores hyaline, mostly unbranched, up to 250 μm long, 8–11 μm wide; phialides 10–16×3.5– 5 μm, hyaline.................................................S. albipes Fungal Diversity Fig. 17 Stachybotrys indicoides (Keshava Prasad et al. 2003) 53. Conidia 8–10.5×4–5.5 μm, blackish green; conidiophores hyaline but dark towards apex, branched, up to 100 μm long, 4–9 μm wide; phialides 8–11×4–6 μm, hyaline at base, dark towards apex...S. chlorohalonata 54. At least some conidia ovoid......................................55 54. Conidia not ovoid......................................................57 55. Conidia ovoid, ellipsoid or oblong, 5–8×2.5–3.5 μm, verrucose, dark grey; conidiophores subhyaline, greyish brown towards apex, rarely branched, 63– 95×4.5–7 μm; phialides 5.5–11.5×2.5–3 μm, pale olive grey................................................S. mangiferae 55. Conidia at least 8 μm long........................................56 56. Conidia ovoid or ellipsoid, 8–9.5×6–7 μm, tuberculate, greyish brown; conidiophores subhyaline, greyish brown towards apex, usually unbranched, 100–120×6–9 μm; phialides 8–12×5–6 μm, verrucose, pale olive-grey..........................S. zhangmuensis 56. Conidia ovoid, sometimes slightly curved into bean shape, 8.5–11.5×4.5–6 μm, coarsely roughened, olivaceous; conidiophores hyaline, unbranched, up to 210 μm long, 4–9 μm wide; phialides 8–10×5– 6 μm, hyaline...............................................S. dichroa 57. Conidia with a ridged, banded or verrucose surface, obovoid or ellipsoid, 7–12×4–6 μm; conidiophores hyaline, dark olivaceous towards apex, sometimes branched, up to 1000 μm long, 3–6 μm wide; phialides 9–14 × 4–6 μm, dark olivaceous................................................S. chartarum 57. Conidia smooth or variously roughened (verrucose, verruculose, tuberculate, rugulose)...........................58 Fungal Diversity Fig. 18 Stachybotrys jiangziensis (Wu and Zhang 2011) 58. Conidia biguttulate, (5.5–)6–8×3–4 μm, verrucose; conidiophores hyaline, stout, unbranched, 40–70×5– 6 μm wide; phialides 7–10 × 3.5–5.5 μm, subhyaline..........................................S. queenslandica 58. Conidia without guttules...........................................59 59. Conidia usually less than 10 μm long or in diam.....60 59. Conidia usually more than 10 μm long.....................67 60. Conidia 4–5.2×2.2–3.5 μm, broadly ellipsoid, smooth or verruculose, brown to dark grey-brown; conidiophores hyaline or pale pigmentation, finely verrucose or smooth, sometimes branched, 6 8 – 1 4 6 × 4 . 2 – 5 μ m ; p h i a l i d e s 8 . 5 – 11 × 3 – 3.6 μm......................................................S. nepalensis 60. Conidia more than 6 μm long or in diam...................61 61. Conidia globose when mature, ellipsoid or pyriform when young, 6–8×4–5 μm, coarsely roughened; conidiophores hyaline to pale brown, up to 55 μm long, 2–4 μm wide, phialides 6–8×4–5 μm...S. microspora 61. Conidia cylindrical or ellipsoid when mature............62 62. Phialides 11–14(−16)×3.8–5.4 μm, hyaline; conidia ellipsoid to obovoid, 7–8.3×3.2–5.1 μm, rugulose; conidiophores hyaline, unbranched, 95–160 × 5.8– 8 μm..........................................................S. cordylines 62. Phialides less than 12 μm long..................................63 63. Mature conidia ellipsoid.............................................64 63. Mature conidia cylindrical to ellipsoid......................66 64. Conidia 6–11×2.5–4 μm, smooth or verrucose, pale olivaceous-brown; conidiophores hyaline to pale olivaceous-brown, smooth, sometimes branched, 41–69 × 3–4 μm; phialides 7.5–12 × 3– 4.5 μm.......................................................S. pallescens 64. Conidia darker (brown to dark brown).......................65 65. Conidia 7–9×3–5 μm, roughened, dark with age; conidiophores hyaline, smooth, mostly unbranched, 72– 143×3–5 μm; phialides 8–11×4–6 μm.....S. aloeticola 65. Conidia 7–9×3.5–4.5 μm, verruculose, brown to dark brown; conidiophores hyaline, smooth, unbranched, 48–85×3–5 μm; phialides 9–11×4–5 μm.........S. zeae Fungal Diversity Fig. 19 Stachybotrys kampalensis (Hansford 1943) 66. Conidia 7–9×3–6 μm, verruculose, olivaceous-brown to brown; conidiophores pale brown, darker towards apex, sometimes branched, 48–94× 4–5 μm; phialides 7.5–11×3 μm, pale brown.....................S. suthepensis 66. Conidia (6.5–)7–9(−9.5) × (2–)2.25–3(−3.5) μm, smooth or verrucose, dark olivaceous-grey; conidiophores hyaline, branched, up to 130 μm long×3.25– 8.5 μm wide; phialides 8.5–12×3.5–5 μm, hyaline or pale straw coloured................................S. breviuscula 67. Conidia less than 5 μm wide......................................68 67. Conidia more than 5 μm wide....................................69 68. Conidia cylindrical, (10–)11–13(−15) × (3.5–)4– 4.5(−5.25) μm, coarsely verrucose; conidiophores hyaline or pale straw coloured, sometimes branched, up to 320 μm long×5.5–8.5 μm wide; phialides 10–14× 4.5–5.25 μm, hyaline to pale brown......S. freycinetiae 68. Conidia cylindrical to cylindrical-ellipsoid, (9.7–)11.6– 13.8(−14.7)×(2.9–)3.4–4.4(−4.6) μm, coarsely roughened, dark olive-grey; conidiophores hyaline, smooth, sometimes branched, 52–88×2.4–4.3 μm; phialides (8.4–)9.6–12.6(−14.3) × (4–)4.3–5.5(−6.1) μm..................................................S. subcylindrospora 69. Conidia ellipsoid to cylindrical, 10–19×5–6.5 μm, verrucose; conidiophores olivaceous, branched, 120– 260×3–3.5 μm; phialides 7.5–11×4.5–5.5 μm, pale brown.................................................S. xanthosomatis 69. Conidia at least 6 μm wide........................................70 70. Conidia oblong, 10–14 × 6–8 μm, verrucose, black; conidiophores hyaline to olivaceous, 150– 250 × 8–9 μm; phialides 10–13 × 5–6 μm, subhyaline............................................S. kampalensis 70. Conidia clavate or oblong, 11–16×6–9 μm, tuberculate, greyish brown; conidiophores hyaline, smooth, rarely branched, 200–250×4.5–7 μm; phialides 13– 15×6–8 μm, pale olive-grey................S. nielamuensis Species accepted in Stachybotrys Seventy-four species are accepted by the authors. 1. Stachybotrys albipes (Berk. & Broome) S.C. Jong & Davis, Mycotaxon 3: 425, 1976 (Fig. 5) ≡ Sporocybe albipes Berk. & Broome, Ann. Mag. nat. Hist., Ser. 2, 8: 19, 1851 Fungal Diversity Fig. 20 Stachybotrys kapiti (Whitton et al. 2001) ≡ Fuckelina albipes (Berk. & Broome) Höhn., Zentbl. Bakt. ParasitKde, Abt. II 60: 14,1923 [1924] = Melanopsamma pomiformis (Pers.) Sacc., Michelia 1: 347, 1878 [sexual morph] ≡ Sphaeria pomiformis Pers., Syn. meth. fung. 1: 65, 1801 = Gliobotrys alboviridis Höhn. [as albo-viridis], Sber. Akad. Wiss. Wien, Math.-naturw. Kl., Abt. 1 111: 1048 [62 of repr.], 1902 = Stachybotrys socia (Sacc.) Sacc. in Ferraris, Annls mycol. 7(3): 283, 1909 Note. This species has been linked to its sexual state, Melanopsamma. Its conidia are olivaceous, smooth, ovoid, mostly 7–9×5–6 μm. No ex-type sequence data in GenBank (2014), but AF081478 (ITS) might be designated as a candidate after detailed morphological description. 2. Stachybotrys aloeticola L. Lombard & Crous, Persoonia 32: 123, 2014 Note. Conidia of S. aloeticola aggregate in slimy masses, allantoid to fusiform, 7–9 × 3–5 μm and contain 1–2 oil droplets. This species was first isolated from Aloe spp. in South Africa (Crous et al. 2014). In morphology, S. aloeticola differs from S. nephrospora by producing longer conidiophores and smaller conidia (Hansford 1943). 3. Stachybotrys bambusicola Rifai, Trans. Br. mycol. Soc. 47: 270, 1964 (Fig. 6) gNote. Its conidia are smooth, colourless or pale pink, obovoid or rarely ellipsoidal with a slightly flattened base, 10–15.5×6·5–8 μm, in pink conidial masses and conidiophores elongate by proliferation (Rifai 1964). Compare it with S. elegans for similarities and differences. D.W. Li’s attempts to borrow the type material (M.A. Rifai 297 (BO) type) have failed. No ex-type sequence data in GenBank (2014). Fungal Diversity Fig. 21 Stachybotrys longispora (Matsushima 1975) 4. Stachybotrys breviuscula [as breviusculus] McKenzie, Mycotaxon 41: 180, 1991 (Fig. 7) Note. Conidia of S. breviuscula are 6.5–9.5×2– 3.5 μm (McKenzie 1991). Compare it with S. parvispora (3–5.5×2.5–3.5 μm), S. freycinetiae (10– 15×3.5–5.5 μm) and S. zeae (7–9×3.5– 4.5 μm) (Ellis 1971a; Jong and Davis 1976; Morgan-Jones and Karr 1976; McKenzie 1991). No ex-type sequence data in GenBank (2014). 5. Stachybotrys chartarum (Ehrenb.) S. Hughes, Can. J. Bot. 36: 812, 1958 (Fig. 8) ≡ Stilbospora chartarum Ehrenb., Sylv. Mycol. Berol.: 9, 21, 1818 ≡ Oospora chartarum (Ehrenb.) Wallr., Fl. Crypt. Germ. 2: 184, 1833 = Stachybotrys atra Corda, Icon. Fung. 1: 21, 1837 = Stachybotrys atra f. atra Corda, Icon. Fung. 1: 21, 1837 Fungal Diversity Fig. 22 Stachybotrys longistipitata (Li et al. 2003) = Stachybotrys lobulata var. angustispora Moreau & V. Moreau, Revue de Mycologie 6(3–4): 83, 1941 = Stachybotrys atra f. genuina Verona = Stachybotrys atra f. lobulata Verona, Cellulosa Carta 2: 94, 139, 1939 = Stachybotrys atra var. brevicaulis (as brevicaule) Verona, Studio sulle cause microbiche che danneggiano la carta ed i libri [Study of the microbiological causes of damage to paper and books]: 41, 1939 = Synsporium biguttatum Preuss, Klotzschii Herb. Viv. Mycol.: no. 1285, 1849 = Stachybotrys lobulata (Berk.) Berk., Outl. Brit. Fung.: 343, 1860 ≡ Sporocybe lobulata Berk., Ann. Mag. Nat. Hist., Ser. 1, 6: 434, 1841 = Stachybotrys lobulata var. angustispora Moreau & V. Moreau, Revue Mycol. 6: 83, 1941 = Stachybotrys lobulata var. macra Pidopl., [Fungus Flora of Coarse Fodders]: 259, 1953 = Stachybotrys alternans Bonord., Handb. Allgem. Mykol.: 117, 1851 ≡ Stachybotrys alternans var. alternans Bonord., Handb. Allgem. mykol.: 117,1851 = Stachybotrys alternans var. atoxica Pidopl., 1946, nom. inval., Art. 36.1[Also invalidly published in 1953] = Memnonium sphaerospermum Fuckel, Jb. nassau. Ver. Naturk. 23–24: 358, 1870 [1869–70] fide Hughes (Hughes 1958) = Stachybotrys scabra Cooke & Harkn., Grevillea 12: 96, 1884 = Stachybotrys atrogrisea Ellis & Everh., J. Mycol. 4: 106, 1888 = Stachybotrys verrucosa Cooke & Massee, in Cooke, Grevillea 16: 102, 1888 = Stachybotrys asperula Massee, Grevillea 16: 26, 1893 ?=Stachybotrys gracilis É.J. Marchal, Bull. Soc. Belg. Micr. 20: 265, 1894 = Stachybotrys pulchra [as pulcra] Speg., Revta Fac. Agron. Vet. Univ. nac. La Plata 2: 248, 1896 = Stachybotrys elasticae Koord., Verh. K. ned. Akad. Wet., Afd. Natuurkunde, Tweede Reeks 13(4): 227, Fungal Diversity Fig. 23 Stachybotrys mangiferae (Misra and Srivastava 1982) 1907 = Stachybotrys cylindrospora C.N. Jensen, Bull. Cornell Univ. Agric. Exp. Stn 315: 496, 1912 = Stachybotrys dakotensis Sacc., Atti Mem. R. Accad. Sci., Lett., Arti, Padova 33:174, 1917 = Stachybotrys voglinii Cif., Annls mycol. 20: 48, 1922 [fide Bisby (1943), Jong and Davis (1976)] = Stachybotrys klebahnii [as klebahni] G. Burchard, Phytopath. Z. 1: 314, 1930 = Synsporium furcatum Losa, Collectanea Botanica 4: 139, 1954 Notes. This is the type species of Stachybotrys and the most common species of Stachybotrys, especially in indoor environments. Despite the segregation of S. chlorohalonata, it remains a species complex. Examination of type materials of Stachybotrys atra Corda and Stilbospora chartarum Ehrenb. by D.W. Li failed to lead to the conclusion that the two epithets are conspecific due to poor condition of Corda’s type. This species develops conidia that are olivaceous to black, mostly ellipsoidal, banded, ridged or verrucose, 7–12× 4–6 μm (Jong and Davis 1976). No ex-type sequence data in GenBank (2014). 6. Stachybotrys chlorohalonata B. Andersen & Thrane, in Andersen, Nielsen, Thrane, Szaro, Taylor & Jarvis, Mycologia 95: 1228, 2003 (Fig. 9) Note. This species was segregated from S. chartarum as a cryptic species according to sequence data. It is differentiated from S. chartarum by its smooth conidia and releasing pale greenish pigment into culture media Fungal Diversity Fig. 24 Stachybotrys microspora (Jong and Davis 1976) (Andersen et al. 2003). It is rather a challenge to differentiate it from S. chartarum based only on morphological characters. Ex-type sequence data = AY180261 (ITS) 7. Stachybotrys cordylines McKenzie, in Pinruan, McKenzie, Jones & Hyde, Fung. Divers. 17: 146, 2004 (Fig. 10) Note. S. cordylines is similar to S. albipes in conidial size and shape. However, the latter species develops smooth conidia, whereas those of S. cordylines are rugulose (Pinruan et al. 2004). No ex-type sequence data in GenBank (2014). 8. Stachybotrys crassa Marchal, Bull. Acad. R. Sci. Belg., Cl. Sci., sér. 5, 34: 140, 1895 Note. It is a distinct species. S. crassa conidia are globose, verruculose, black, 16–18 μm diam., phialides ovoid, colourless, 17–21×10–12.5 μm (Marchal 1895). It is similar to S. kapiti, S. microspora, S. nilagirica, S. ruwenzoriensis and S. sphaerospora. It has much smaller conidia than those of S. nilagirica (15.4– 28 μm diam.), but they are much bigger than those of S. kapiti (7.5–10.5×7–10.5 μm), S. microspora (5–6 μm diam.), S. ruwenzoriensis (6–8 μm diam.) and S. sphaerospora (11–12 μm diam.) (Jong and Davis Fungal Diversity Fig. 25 Stachybotrys mohanramii (Manoharachary et al. 2006) 1976; Morgan-Jones and Sinclair 1980; Subramanian 1954; Whitton et al. 2001). No ex-type sequence data in GenBank (2014). 9. Stachybotrys dichroa Grove, J. Bot., Lond. 24: 201, 1886 (Fig. 11) Note. It is distinct taxon. However, cultures and specimens under this name in herbaria and culture collections around the world are not the one taxon. Molecular data reported in several papers showed that some cultures are misidentified. Its conidia were originally reported as oblong, 1-septate, 10 × 5 μm (Grove 1886). This is the second species mistakenly described as possessing ‘1-septate’ conidia. Examination of type material (IMI18006) showed that ‘1-septate’ is erroneous, possibly, due to guttules in the conidia. The key morphological characters of this species are that its conidiophores are thickwalled, smooth, colourless, and conidia are olivaceous, coarsely roughed, thick-walled, ovate, some are slightly curved into bean shape, single-celled, 8.5–11.5×4.5–6 μm based on the measurement of type material. No ex-type sequence data in GenBank (2014), but in this paper we designate ATCC 18913 as the epitype culture, thus the epitype sequence is AF081472 (ITS). 10. Stachybotrys echinata (Rivolta) G. Sm., Trans. Br. mycol. Soc. 45: 392, 1962 (Fig. 12) 11. Stachybotrys elegans (Pidopl.) W. Gams, Compendium of Soil Fungi: 746, 1980 ≡ Hyalobotrys elegans Pidopl., Gribnaja Flora Grubych Kormov [Fungus flora of coarse fodder]: 186, 1948 = Stachybotrys pallida Orpurt, Studies on the soil microfungi of Wisconsin prairies, Diss. Univ. Wisconsin: 95, 1954, [nom. inval., Art. 30.5, 36.1] = Hyalostachybotrys bisbyi Sriniv., J. Indian bot. Soc. 37: 340, 1958 [nom. inval., Art. 8] ≡ Stachybotrys bisbyi (Sriniv.) G.L. Barron, Mycologia 56: 315 1964 [nom. inval., Art. 8] = Hyalostachybotrys sacchari Sriniv., J. Indian bot. Soc. 37: 341, 1958 [nom. inval., Art. 8] ≡ Stachybotrys sacchari (Sriniv.) G.L. Barron, Mycologia 56: 315, 1964 = Stachybotrys aurantia G.L. Barron, Can. J. Bot. 40: 258, 1962 Note: It develops colourless, smooth, limoniform or fusiform, guttulate conidia, 8–14×6–9 μm and salmon coloured colonies. Both molecular data and morphological characters showed that it is a species complex (Pers Observ., Gams pers. comm. 2012). Further studies are necessary to clarify the confusion of this species. Both type cultures of Hyalostachybotrys bisbyi (SBI 696) and H. sacchari (SBI 781) are currently not maintained at Sugarcane Breeding Institute (SBI), India according to Fungal Diversity Fig. 26 Stachybotrys nephrodes (McKenzie 1991) Dr. R. Viswanathan, Principal Scientist (Plant Pathology) & Head, Division of Crop Protection, Sugarcane Breeding Institute (pers. comm. 2012). Since type specimen for H. bisbyi, type species of Hyalostachybotrys is a culture, this genus is invalid (Art. 8, St. Louis Code). This species is occasionally isolated from indoor environments. No ex-type sequence data in GenBank (2014). 12. Stachybotrys eucylindrospora D.W. Li, Mycologia 99: 333, 2007 (Fig. 13) = Stachybotrys striatispora Orpurt, Studies on the soil microfungi of Wisconsin prairies, Diss. Univ. Wisconsin: 93, 1954 [Invalid, Art. 30.5, 36.1] Note: The conidia of this species are olivaceous, cylindrical, with longitudinal striations, and measure 12.8–16×3.4–5.5 μm (Li 2007). S. cylindrospora had been misapplied to this species for many years and many herbarium specimens of this species are still under the name of S. cylindrospora. There is no doubt that Orpurt (1954) found this species but, unfortunately, it was not validly published in his Ph.D. dissertation to meet requirements of International Code of Botanical Nomenclature (Stockholm Code 1952). Ex-type sequence data = AF081474 (ITS). 13. Stachybotrys freycinetiae McKenzie, Mycotaxon 41: 183, 1991 (Fig. 14) Fungal Diversity Fig. 27 Stachybotrys nephrospora (Jong and Davis 1976) = Ornatispora novae-zealandiae Whitton, McKenzie & K.D. Hyde, Fungi Associated with Pandanaceae. Fungal Divers. Res. Ser. 21: 88, 2012 [sexual state] Note. This taxon is characterised by black, 10–15× 3.5–6 μm, coarsely verrucose, somewhat cylindrical conidia, rounded at the apex and rounded or truncate at the base (McKenzie 1991). Whitton et al. (2012) connected S. freycinetiae to the Ornatispora sexual state because of the hyphal subiculum containing fertile conidiophores that surrounds the ascomata. No ex-type sequence data in GenBank (2014). 14. Stachybotrys frondicola (K.D. Hyde, Goh, Joanne E. Taylor & J. Fröhl.) Yong Wang bis, K.D. Hyde, McKenzie, Y.L. Jiang & D.W. Li, comb. nov. = Ornatispora frondicola K.D. Hyde, Goh, Joanne E. Taylor & J. Fröhl., Mycol. Res. 103(11): 1438 (1999) MycoBank MB 809095 Note. No ex-type sequence data in GenBank (2014). 15. Stachybotrys gamsii (K.D. Hyde, Goh, Joanne E. Taylor & J. Fröhl.) Yong Wang bis, K.D. Hyde, McKenzie, Y.L. Jiang & D.W. Li, comb. nov. = Ornatispora gamsii K.D. Hyde, Goh, Joanne E. Taylor & J. Fröhl., Mycol. Res. 103(11): 1432 (1999) = ?Didymostilbe aurantiospora Seifert & G. Okada, Stud. Mycol. 27: 133 (1985) (asexual state) MycoBank MB 809096 Note. The connection of sexual and asexual states of this species needs to be studied to determine which epithet should be used, since ‘aurantiospora’ has priority over ‘gamisii’. No ex-type sequence data in GenBank (2014). 16. Stachybotrys globosa P.C. Misra & S.K. Srivast., Trans. Br. Mycol. Soc. 78: 556, 1982 (Fig. 15) Note. Conidia of this species are globose or subglobose, dark olivaceous, rough, 5.7–7.1 × 5.6– 7.4 μm (Misra and Srivastava 1982). It is very similar to S. microspora in conidial shape and size and these two species should be further studied to determine if they are Fungal Diversity Fig. 28 Stachybotrys nielamuensis (Wu and Zhang 2009) Fig. 29 Stachybotrys nilagirica (Subramanian 1957) Fungal Diversity Fig. 30 Stachybotrys oenanthes (Ellis 1971) conspecific. Matsushima (1985) indicated that S. ruwenzoriensis is similar to S. globosa in conidial size and shape, but in the latter species the conidiophores branch sympodially. Conidia of S. ruwenzoriensis are papillate at the base. No ex-type sequence data in GenBank (2014). 17. Stachybotrys guttulispora Muhsin & Al-Helfi, Sydowia 34: 133, 1981 (Fig. 16) Note. This is a distinct species. The conidia of this taxon are ellipsoid, olivaceous or greenish brown, smooth, biguttulate, 9–12 × 3.5–5 μm (Muhsin and Al-Helfi 1981). Its conidia are bigger than those of S. albipes (7–9×5–6 μm) (Jong and Davis 1976). No ex-type sequence data in GenBank (2014). 18. Stachybotrys havanensis Mercado & J. Mena, Acta Bot. Cubana 55: 2, 1988 Note. Conidia are cylindrical, fusiform, or ellipsoid, colourless, smooth, grey or dark grey, 7–10.2×2.5– 3.5 μm, with a truncate base (Mercado-Sierra and Mena-Portales 1988). No ex-type sequence data in GenBank (2014). 19. Stachybotrys indicoides Yong Wang bis, K.D. Hyde, McKenzie, Y.L. Jiang & D.W. Li, nom. nov. (Fig. 17) ≡ Memnoniella indica Kesh. Prasad, Asha & Bhat, Mycotaxon 85: 341, 2003 Fungal Diversity Fig. 31 Stachybotrys pallescens (Jiang and Zhang 2009) Fig. 32 Stachybotrys palmae (Pinruan et al. 2004) Fungal Diversity Fig. 33 Stachybotrys parvispora (Hughes 1952) Non Stachybotrys indica P.C. Misra, Mycotaxon 2(1): 107, 1975 MycoBank MB 809098 Etymology: Latin, means indica-like. Note. To avoid its homonym status with Stachybotrys indica P.C. Misra, a new name is proposed. Its conidia are similar to S. echinata in roughness, size and shape. However, its conidia are 5– 6 μm diam., thick-walled and with a small papilla at the base (Keshava Prasad et al. 2003). No ex-type sequence data in GenBank (2014). 19. Stachybotrys jiangziensis Y.M. Wu & T.Y. Zhang, Mycotaxon 114: 459, 2011 (Fig. 18) Note. Conidiophores erect, branched, 2–4-septate, subhyaline near the base, greyish brown above, smooth, 60–80 μm long, 4–5 μm wide near the base; phialides borne in groups of 6–8 at the apices of conidiophores, pale brown, smooth, 8–10×5–7 μm; conidia globose to subglobose, tuberculate, brown to dark brown, 6–9 μm in diam. (Wu and Zhang 2011). No ex-type sequence data in GenBank (2014). 20. Stachybotrys kampalensis Hansf., Proc. Linn. Soc. London 155: 45, 1943 [1942–43] (Fig. 19) Note. Conidiophores erect, 150–250 × 8–9 μm; phialides 4–9, terminal, subhyaline, cylindrical or ellipsoid, 10–13×5–6 μm; conidia black, oblong, granulate, 10–14×6–8 μm (Hansford 1943). No ex-type sequence data in GenBank (2014). 21. Stachybotrys kapiti Whitton, McKenzie & K.D. Hyde, N.Z. Jl Bot. 39: 493, 2001 (Fig. 20) Note. Conidia 7.5–10.5×7–10.5 μm, broadly ellipsoidal or broadly clavate to nearly spherical at maturity, apical end broadly rounded, basal end typically with a single papilla, black, verrucose. The species that produce more or less spherical conidia are S. crassa (16–18 μm diam.), S. nilagirica (15– 28 μm diam.), and S. sphaerospora (11–12 μm diam.), S. microspora (5–6 μm diam.) and Fungal Diversity Fig. 34 Stachybotrys queenslandica (Matsushima 1989) S. ruwenzoriensis (6–8 μm diam.). The conidia of these species are either larger or smaller than S. kapiti. No extype sequence data in GenBank (2014). 22. Stachybotrys leprosa (R.F. Castañeda) R.F. Castañeda comb. nov. ≡ Memnoniella leprosa R.F. Castañeda; Fungi Cubense (La Habana): 10, 1986 MycoBank MB 809099 Note. This is a synnematous species. Its conidia are rather unique, verrucose and black at the apical side, but smooth and subhyaline at the basal side, 7–12 μm diam., which are much bigger than those (3–5 μm diam.) of S. stilboidea (Castañeda 1986). No ex-type sequence data in GenBank (2014). 23. Stachybotrys levispora (Subram.) Yong Wang bis, K.D. Hyde, McKenzie, Y.L. Jiang & D.W. Li, comb. nov. ≡ Memnoniella levispora Subram., J. Indian Bot. Soc. 33: 40, 1954 MycoBank MB 809100 Note. Conidia of this species are smooth (Subramanian 1954). Its size is similar to S. echinata. No ex-type sequence data in GenBank (2014). 24. Stachybotrys longispora Matsush., Icon. microfung. Matsush. Lect.: 145, 1975 (Fig. 21) Note: It is a rather characteristic species with conidia narrowly cylindrical, smooth, rounded at both ends, pale olivaceous, 8.8–12×2–2.4 μm (Matsushima 1975). Ex-type sequence data = AF081482.1 (ITS) in GenBank (2014). 25. Stachybotrys longistipitata (D.W. Li, Chin S. Yang, Vesper & Haugland) D.W. Li, Chin S. Yang, Vesper & Haugland comb. nov. (Fig. 22) ≡ Memnoniella longistipitata D.W. Li, Chin S. Yang, Vesper & Haugland, Mycotaxon 85: 254, 2003 MycoBank MB 809101 Note. Holotype MFC-2994, ex-holotype ATCC 22699, (BPI 843690 isotype). This species was deposited as M. subsimplex. However, analysis of ITS sequence data (AF081471) showed that it is a distinct species from S. subsimplex (Haugland et al. 2001). The conidia of M. longistipitata are similar to S. subsimplex in shape and size. However, very long conidiophores and bimorphic conidia of M. longistipitata differentiate it from S. subsimplex (Li et al. 2003). Ex-type sequence data = AF081471.2 (ITS) in GenBank (2014). 26. Stachybotrys mangiferae P.C. Misra & S.K. Srivast., Trans. Br. mycol. Soc. 78: 556, 1982 (Fig. 23) Note: Conidia are ovoid, ellipsoid or oblong, verrucose, dark grey, 5–8×2.5–3.5 μm (Misra and Srivastava 1982). No ex-type sequence data in GenBank (2014). 27. Stachybotrys mexicanus J. Mena & Heredia [as ‘mexicana’], Boln Soc. Micol. Madrid 33: 12, 2009 Note: Conidiophores solitary, erect, straight of Fungal Diversity Fig. 35 Stachybotrys ramosa (Dorai and Vittal 1987) flexuose, unbranched, colourless, smooth, 120 μm long, (3.5–)5–7 μm wide at the base, 2–3.5 μm wide at the apex; phialides obovoid or ellipsoid, smooth, colourless, 9.5–16×3–5 μm; conidia broadly ellipsoid, limoniform, subsphaerical or spherical, slightly protrudent at the base, colourless, smooth; broadly ellipsoid or limoniform conidia, 10.5–16×8–12 μm, subsphaerical or spherical conidia, 9–12.5 μm diam. (Mena-Portales et al. 2009). No ex-type sequence data in GenBank (2014). 28. Stachybotrys microspora (B.L. Mathur & Sankhla) S.C. Jong & E.E. Davis, Mycotaxon 3: 448, 1976 (Fig. 24) ≡ Stachybotrys atra var. microspora B.L. Mathur & Sankhla, Sci. Cult. 32: 93, 1966 Note: Jong and Davis (1976) redescribed this taxon based on their study of type materials. In their description the conidia of this species are ellipsoid, pyriform, 6– 8×4–5 μm when young, becoming spherical, 5–6 μm in diam. and coarsely roughened when mature. It is similar in conidial shape and size to S. globosa and these two species should be further studied to determine if they are conspecific. No ex-type sequence data in GenBank (2014), but ATCC 18852 was designated as epitype, so AF081475 (ITS) is the epitype sequence. 29. Stachybotrys mohanramii (Manohar., D.K. Agarwal, Kunwar, Sureshk. & Sharath,) Yong Wang bis, K.D. Hyde, McKenzie, Y.L. Jiang & D.W. Li, comb. nov. Fungal Diversity Fig. 36 Stachybotrys reniformis (Tubaki 1963) (Fig. 25) ≡ Memnoniella mohanramii Manohar., D.K. Agarwal, Kunwar, Sureshk. & Sharath, Indian Phytopath. 59(4): 489, 2006 MycoBank MB 809102 Note. This is the only species that develops ellipsoidal, smooth conidia in chains, 5.5–6.5×2.5–3.8 μm (Manoharachary et al. 2006). No ex-type sequence data in GenBank (2014). 30. Stachybotrys nepalensis (Whitton, McKenzie & K.D. Hyde) Whitton, McKenzie & K.D. Hyde comb. nov. = Ornatispora nepalensis Whitton, McKenzie & K.D. Hyde, Fungi associated with Pandanaceae. Fungal Divers. Res. Ser. 21: 86, 2012 [sexual stage] MycoBank MB 809104 Note. The new combination is proposed according to the Melbourne Code. Conidiophores 68–146 μm long, 4.2–5 μm wide at base, 2–3 μm wide at apex, hyaline or pale in pigmentation, smooth or finely verrucose; conidiogenous cells ellipsoid, 8.5–11×3–3.6 μm; conidia broadly ellipsoid, rounded at both ends, brown to dark grey-brown, typically smooth, sometimes verrucose and darker in colour, 4–5.2×2.2–3.5 μm. Whitton et al. (2012) connected Stachybotrys sp. to the Ornatispora sexual state because the setae, which arise from the ascomata, sometimes elongate and develop into a fertile conidiophore. No ex-type sequence data in GenBank (2014). 31. Stachybotrys nephrodes McKenzie, Mycotaxon 41: 185, 1991 (Fig. 26) Note. It is a very characteristic species. Conidia tightly curled or reniform, pale olivaceous with a 1.75–4.5 m wide black or dark brown outer zone, verrucose, 13– 15×12–14 μm. No ex-type sequence data in GenBank (2014). Fungal Diversity Fig. 37 Stachybotrys renispora (Misra 1976) 32. Stachybotrys nephrospora Hansf., Proc. Linn. Soc. London 155: 45, 1943 [1942–43] (Fig. 27) Note: Its conidia are allantoid (reniform), black, smooth, 8–11×4.5–6 μm (Hansford 1943). Jong and Davis (1976) re-described its conidia as coarsely roughened, 10–12 × 4–5 μm using ex-type culture (ATCC18839 = IFO 7076) of S. reniformis. Our examination of holotype (K(M) 16581) of S. nephrospora showed that its conidia are truly smooth, not coarsely roughened. The inconsistency between the type materials raises the questions about current treatment of these two taxa. Are S. nephrospora and S. reniformis conspecific as treated by Jong and Davis (1976)? Should the characters of conidial surface be used to separate the two species? An isolate from indoor environment developed smooth conidia and this character does not change (pers. obser. of D.W. Li). In our opinion these two are distinct species. No ex-type sequence data in GenBank (2014). 33. Stachybotrys nielamuensis Y.M. Wu & T.Y. Zhang, Mycotaxon 109: 461, 2009 (Fig. 28) Note. Conidia clavate or oblong, obviously tuberculate, greyish brown, 11–16×6–9 μm (Wu and Zhang 2011). In conidial morphology S. nielamuensis is similar to S. k amp ale nsis Hans f. (Hans ford 19 43 ), S. freycinetiae McKenzie (1991), S. verrucispora Matsush. (Matsushima 1985), S. xigazenensis Y.M. Wu & T.Y. Zhang, and S. zhangmuensis Y.M. Wu & T.Y. Zhang. However S. kampalensis, S. freycinetiae and S. xigazenensis have smaller conidia (9–13×6–7 μm, 11–13 × 4–4.5 μm, 8–9.5 × 6–7 μm, respectively). S. verrucispora develops pale brown conidiophores and phialides, and shorter but wider conidia, 10–15×9.5– 11 μm. No ex-type sequence data in GenBank (2014). 34. Stachybotrys nilagirica Subram., Proc. Indian Acad. Sci., Pl. Sci. 46: 331, 1957 (Fig. 29) Note: This is a well characterized species. Its large Fungal Diversity Fig. 38 Stachybotrys reniverrucosa (Whitton et al. 2001) globose conidia, 15.4–28 μm diam. separate it from similar species. See S. crassa for additional discussion. No ex-type sequence data in GenBank (2014). 35. Stachybotrys oenanthes M.B. Ellis, Mycol. Pap. 125: 29, 1971 (Fig. 30) Note: Conidiophores erect, simple, straight or flexuous, 1–2 septate, solitary, smooth, smoke grey to black, 120–180 μm long, 6–9 μm wide; phialides cylindrical or obovoid, colourless at first, later smoke grey to black, smooth or verrucose, 12–21×4–7 μm; conidia are reniform or ellipsoidal, sometimes attenuated at the base, 9–12×4.5–8 μm, smooth or verrucose (Ellis 1971; Ellis 1976). See S. reniverrucosa for additional discussion. Ex-type sequence data = AF081473.1 (ITS) in GenBank (2014). 36. Stachybotrys oleronensis Lechat, Hairaud & LesageMeessen, Persoonia 31: 283, 2013 Note: This species is not known to produce conidia, but it does produce Nectriella-like ascocarps and ascospores. Molecular data (ITS) showed that the species belongs in Stachybotrys (Lechat et al. 2013). Ex-type sequence data = KF777192. 37. Stachybotrys pallescens Y.L. Jiang & T.Y. Zhang, Mycosystema 28: 646, 2009 (Fig. 31) Note: Conidiophores colourless to pale olivaceous- brown, smooth, unbranched or branched, straight or flexuous, 41–69×3–4 μm. Phialides pale olivaceousbrown, smooth, 7.5–12×3–4.5 μm. Conidia cylindrical or subcylindrical, rounded at the apex, truncate at the base, smooth or verrucose, slightly olivaceousbrown when mature, 6–11× 2.5–4 μm (Jiang and Zhang 2009). It is similar to S. sansevieriae in conidial morphology. Conidia of S. sansevieriae are ellipsoid or boat-shaped, straight, truncate at the base, dark brown, smooth, 6–9 × 3–4 μm (Ellis 1976). The two taxa need to be further studied to determine if they are conspecific. Ex-type sequence data = KC305345 (LSU). 38. Stachybotrys palmae Pinruan, in Pinruan, McKenzie, Jones & Hyde, Fungal Diversity 17: 146, 2004 (Fig. 32) Note: S. palmae is one of only five species that produce colourless conidia. Among these taxa, S. palmae is the only species with rough-walled conidia. Those of the other four species, S. bambusicola, S. elegans, S. guttulispora, and S. palmijunci are smooth-walled. No ex-type sequence data in GenBank (2014). 39. Stachybotrys palmicola (K.D. Hyde, Goh, Joanne E. Taylor & J. Fröhl.) Yong Wang bis, K.D. Hyde, McKenzie, Y.L. Jiang & D.W. Li, comb. nov. ≡ Ornatispora palmicola K.D. Hyde, Goh, Joanne E. Fungal Diversity Fig. 39 Stachybotrys ruwenzoriensis (Matsushima 1985) Taylor & J. Fröhl., Mycol. Res. 103(11): 1438 (1999) MycoBank MB 809106 Note. No ex-type sequence data in GenBank (2014). 40. Stachybotrys palmijunci Rifai, Reinwardtia 8: 537, 1974 Note: Mycelia form poorly developed stromata; conidiophores erect, straight or flexuous, septate, elongating Fig. 40 Stachybotrys sansevieriae (Misra 1975) by proliferation, up to 275 μm, thick-walled, dark reddish below; phialides smooth, obovate-cylindrical to curved, colourless, 9–12.5×4.5–5.5 μm; conidia ellipsoid or subfusiform, colourless or salmon coloured, smooth, 13.8–18.4 × 4.5–5.5 μm, guttulate. It is similar to S. bambusicola, but has much longer conidia and smaller phialides. No ex-type sequence data in GenBank (2014). Fungal Diversity Fig. 41 Stachybotrys sphaerospora (Morgan-Jones and Sinclair 1980) 41. Stachybotrys parvispora S. Hughes, Mycol. Pap. 48: 74, 1952 (Fig. 33) Note: Conidia are dark brown, ovoid, smooth, 3.5– 6×3–3.5 μm (Hughes 1952). This taxon resembles S. chartarum in morphology but its smaller and smooth conidia differentiate it from the latter. No ex-type sequence data in GenBank (2014). 42. Stachybotrys proliferata K.G. Karand., S.M. Kulk. & Patw., Biovigyanam 18: 79, 1992 Note: The conidia of S. proliferata are small (5.5–7× 4–5 μm), smooth, reniform; conidiophores indeterminate, extending by proliferation (2–4 times) via apex of conidiophores at the place of phialides (Karandikar et al. 1992). Its proliferation and reniform conidia differentiate Fungal Diversity Fig. 42 Stachybotrys subcylindrospora (Jie et al. 2013) it from other species. No ex-type sequence data in GenBank (2014). 43. Stachybotrys punctata (Dulym., P.F. Cannon, K.D. Hyde & Peerally) Yong Wang bis, K.D. Hyde, McKenzie, Y.L. Jiang & D.W. Li, comb. nov. ≡ Ornatispora punctata Dulym., P.F. Cannon, K.D. Hyde & Peerally, Fungal Diversity 8: 95 (2001) MycoBank MB 809107 Note. No ex-type sequence data in GenBank (2014). 44. Stachybotrys queenslandica Matsush., Matsush. Mycol. Mem. 6: 40, 1989 (Fig. 34) Note. Conidiophores scattered, solitary, cylindrical, erect, straight, simple, 1–3-septate, rigid, smooth, bearing 8–12 phialides at the apex, 40–70 μm in length (including phialides), swollen at the base, 5–6 μm wide above the base. Phialides obovoid, subhyaline, 7–10× 3.5–5.5 μm, collapsing when aged. Conidia oblong, verruculose not striated, biguttulate, 6–8×3–4 μm. No ex-type sequence data in GenBank (2014). 45. Stachybotrys ramosa Dorai & Vittal, Trans. Br. mycol. Soc. 87: 642, 1987 [1986] (Fig. 35) Note: Conidiophores differentiated, erect, but loosely intertwined, flexuous, sympodially branched, smooth, dark brown, up to 400 μm in length and 4–5 μm wide; conidiophore branches very short, colourless, smooth, 6–8 μm long, 3–4 μm wide, bearing a group of 3–7 phialides at their apices. Phialides 4–5.5 μm long and 2.5–3.5 μm diam. Conidia subsphaerical to ellipsoidal, smooth, dark brown to black, 7–9×6–7 μm (Vittal and Dorai 1986). This is a valid species. However, its type no longer has any fungal structures for examination. A neotype should be designated in the future. Compare it with S. proliferata for difference. The latter grows by proliferation at the location of phialide, without short branches. No ex-type sequence data in GenBank (2014). 46. Stachybotrys reniformis Tubaki, Trans. Mycol. Soc. Japan 4: 86, 1963 (Fig. 36) = Stachybotrys nephrospora Hansf. , Proc. Linn. Soc. London 155: 45, 1943 [1942–43] Note: Conidiophores erect, but sometimes slightly curved, unbranched, solitary or in groups, rough at the upper parts, 100–130 μm long or more, 3–4.5 μm wide, Fungal Diversity Fig. 43 Stachybotrys subreniformis (Li and Jiang 2011) slightly enlarged at apices, colourless, bearing terminal phialides in whorl of 3–5. Phialides obovate or clavate, smooth, colourless, 10–13×4–6 μm. Conidia reniform or comma shaped, biguttulate, smooth at first, then markedly warted when mature, 10–12×5–6 μm, dark olive green to nearly black (Tubaki 1963). It is similar to S. nephrospora. The only difference between the two taxa is whether the conidia are smooth, as in S. nephrospora, or warted as in S. reniformis. See S. nephrospora for additional discussion. Our opinion is that they should be separate species. No ex-type sequence data in GenBank (2014). 47. Stachybotrys renispora P.C. Misra, Mycotaxon 4: 161, 1976 (Fig. 37) Note: Conidiophores sympodially branched, 20– 50 μm in length, 2.3–3.2 μm wide, subhyaline to pale greyish brown, minutely verrucose; phialides subhyaline to pale greyish brown, minutely verrucose, 7.5–9.3×3– 4.5 μm; conidia reniform, black, smooth, 5.2–7×3.5– 5.2 μm (Misra 1976). See S. renisporoides for additional discussion. No ex-type sequence data in GenBank (2014). 48. Stachybotrys renisporoides K.G. Karand., S.M. Kulk. & Patw., Biovigyanam 18: 79, 1992 ?=Stachybotrys renispora P.C. Misra, Mycotaxon 4(1): 161, 1976 Note: Conidiophores single or 2–3 in group, unbranched, pale brown, colourless or subhyaline near apex, smooth, 65–121.5×5.5–8.5 μm; phialides olivaceous, obovoid, smooth, 7–11×4.5–6 μm; conidia reniform, olivaceous to black, smooth, 4.5–7×3–4.5 μm (Karandikar et al. 1992). This taxon is morphologically similar to S. renispora. The phialides of the latter are minutely verruculose. When Karandikar et al. (1992) discussed the differences between the two taxa, they mistakenly referred conidia of S. renispora as minutely verruculose. The conidia of S. renispora are actually smooth (Misra 1976). Any difference between these two taxa needs to be further studied. 49. Stachybotrys reniverrucosa Whitton, McKenzie & K.D. Hyde, N.Z. Jl Bot. 39: 496, 2001 (Fig. 38) Note: Conidia reniform, ellipsoidal when viewed from above or below, both ends rounded, 10–13.5×6– 9.5 μm, light grey, but the coarse ornamentation of the conidia is black, coarsely granulate. S. reniverrucosa is not conspecific with previously described species which have reniform conidia. In S. nephrospora s.l. (including S. reniformis and S. sinuatophora) the conidia are smaller (8–12× 4.5–6 μm), black, smooth or verrucose (Ellis 1971). The conidia of S. oenanthes are reniform or ellipsoidal, 9–12 × 4.5–8 μm, smooth or verrucose Fungal Diversity Fig. 44 Stachybotrys subsimplex (Jong and Davis 1976) (Ellis 1976). The conidia of S. proliferata are small (5.5– 7×4–5 μm). S. renispora has small (5–7×3.5–5 μm), black, smooth conidia (Misra 1976). The conidiophores of S. proliferata continue growing by proliferation via the apex to produce whorls of phialides (Karandikar et al. 1992). The conidia of S. nephrodes are tightly curled, versicoloured, with a pale inner zone and a black outer zone, verrucose, and 11–15.5 × 10.5–14 μm (McKenzie 1991). No ex-type sequence data in GenBank (2014). 50. Stachybotrys ruwenzoriensis Matsush., Matsush. Mycol. Mem. 4: 17, 1985 (Fig. 39) Note. Conidiophores solitary, simple, erect, 70– 90 μm long, 4–6 μm wide at the base; phialides obovoid 10–15×4.6–6 μm, smooth, pale olivaceous; conidia globose, papillate at the base end, 6–8 μm in diam., inconspicuously verruculose, dark brown (Matsushima 1985). Its papillate conidia separate it from other species with globose conidia. No ex-type sequence data in GenBank (2014). 51. Stachybotrys sansevieriae [as sansevierii] G.P. Agarwal & N.D. Sharma, in Sharma & Agarwal, J. Indian Bot. Soc. 53: 78, 1974 (Fig. 40) = Stachybotrys indica P.C. Misra, Mycotaxon 2: 107, 1975 Note. Conidia ellipsoid or boat-shaped, straight, truncate at the base, dark brown, smooth, 6–9×3–4 μm (Ellis 1976). No ex-type sequence data in GenBank (2014). 52. Stachybotrys sinuatophora Matsush., in Kobayasi et al., Bull. natn. Sci. Mus., Tokyo 14(3): 476, 1971 ≡ Cephalosporiopsis sinuatophora (Matsush.) C. Booth, ined. Note. Conidiophores repeatedly alternately branched, undulating or even circinate, up to 360 μm in height, 3– 4.5 μm wide near the base, tapered toward the apex, 2.2– 2.8 μm wide just below verticillate phialides, smooth or finely verrucose, colourless to pale brown; phialides verticillate, terminate, 2–4 phialides in whorls, ovate, smooth, 8–11×5–7 μm, brown, paler towards the base; conidia reniform, dark brown, verrucose, 8–12×6– Fungal Diversity Fig. 45 Stachybotrys terrestris (Kong, Zhang, and Zhang 2007) 7 μm. Jong and Davis (1976) demoted it to a synonym of S. nephrospora. However, in our opinion, after examining the type material, its very characteristic undulating conidiophores make it a distinct species. No ex-type sequence data in GenBank (2014). 53. Stachybotrys sphaerospora Morgan-Jones & R.C. Sinclair, Mycotaxon 10: 372, 1980 (Fig. 41) Note. The conidia of this species are ellipsoid to ovoid when young, spherical at maturity, dark brown to black, 11–12 μm in diam., ornamented by irregular and prominent ridges (Morgan-Jones and Sinclair 1980). See S. crassa for detailed discussion on morphologically similar species. No ex-type sequence data in GenBank (2014). 54. Stachybotrys stilboidea Munjal & J.N. Kapoor, Mycopath. Mycol. appl. 39: 121, 1969 ≡ Memnoniella stilboidea (Munjal & J.N. Kapoor) M.B. Ellis, More Dematiaceous Hyphomycetes: 464, 1976 Note. This is one of two synnematous species in Stachybotrys including Memnoniella. Its conidia are dark brown to black, globose to slightly angular, 3– 5 μm diam. (Munjal and Kapoor 1969). No ex-type sequence data in GenBank (2014). 55. Stachybotrys subcylindrospora C.Y. Jie, Y.L. Jiang, D.W. Li, McKenzie & Yong Wang bis, Mycol. Prog. 12: 695, 2013 (Fig. 42) Conidiophores solitary or in groups, simple or irregularly branched, smooth, hyaline, 52–88×2.4–4.3 μm; phialides borne in groups of 3–8 clavate, smooth, 8.4– 14.3×4.0–6.1 μm; conidia cylindrical or subcylindrical, truncate at the base, rounded at the apex, surface of both young and mature spores show delicate and irregular striations under oil lens, 9.7–14.7×2.9–5.0 μm, usually Fungal Diversity Fig. 46 Stachybotrys thaxteri (Li 2011) containing one to three oil drops, especially when young. It is similar to S. longispora and S. eucylindrospora in producing cylindrical conidia (Matsushima 1971, 1975; Li 2007), but differs in conidial size and ornamentation. Conidia of S. subcylindrospora have irregular striations, while those of S. eucylindrospora have longitudinal striations, and those of S. longispora are smooth. Ex-type sequence data = KC305354. 56. Stachybotrys subreniformis Q.R. Li & Y.L. Jiang, Mycotaxon 115: 171, 2011 (Fig. 43) Note: Conidiophores solitary, erect or flexuous, branched, septate, hyaline to dark brown, 48–98×5.0– 7.5 μm, tapering toward the apex which bears terminal phialides in a whorl of 2–7 around a central phialide. Phialides brown, obovate, smooth, 8.0–11.5 × 4.5– 6.0 μm. Conidia aggregated in slimy masses, spherical or slightly reniform, dark brown, verruculose, 6.0–9.5× Fungal Diversity Fig. 47 Stachybotrys theobromae (Hansford 1943) 4.5–7.5 μm (Li and Jiang 2011). It is similar to S. microspora. The ITS phylogenetic result confirms it should be a synonym of S. chartarum. Ex-type sequence data = KC305344. 57. Stachybotrys subsimplex Cooke, Grevillea 12: 33, 1883 (Fig. 44) ≡ Memnoniella subsimplex (Cooke) Deighton, Mycol. Pap. 78: 5, 1960 = Haplographium musae Sawada, Natn. Taiwan Univ., Coll. Agric., Spec. Publ., 8: 193, 1959. Note: This species is very similar to S. echinata but its larger conidia (6–9 μm diam.) differentiate it from S. echinata (4–5 μm diam.). It is not as common as S. echinata, especially in indoor environments. S. subsimplex is a well defined species. There is no extype culture available. Thus, Haugland et al. (2001) proposed ATCC 32888 as an epitype strain for S. subsimplex. However, this proposal is questionable. ATCC 32888 was deposited as M. echinata. Its conidia are 5–7 μm diam., which is in between those of S. echinata and S. subsimplex. The morphological characters of its conidia do not fit those of type specimen IMI 10941. No ex-type sequence data in GenBank (2014). 58. Stachybotrys suthepensis Photita, P. Lumyong, K.D. Hyde & McKenzie, in Photita, Lumyong, McKenzie, Hyde & Lumyong, Cryptog. Mycol. 24: 149, 2003 Note: Conidiophores smooth or verrucose, pale brown, darker towards apex, 48–94 μm in length; Fungal Diversity Fig. 48 Stachybotrys variabilis (Wang and Zhang 2009) phialides 7.5–11×3 μm; conidia ellipsoid or cylindrical, rounded at both ends, olivaceous-brown, smooth when young, verruculose at maturity, 7–9×3–6 μm (Photita et al. 2003). It shares similarities with S. albipes, S. chartarum, and S. zeae in morphology. See Photita et al. (2003) for detailed comparisons. No ex-type sequence data in GenBank (2014). 59. Stachybotrys taiwanensis (Sivan. & W.H. Hsieh) Yong Wang bis, K.D. Hyde, McKenzie, Y.L. Jiang & D.W. Li, comb. nov. ≡ Niesslia taiwanensis Sivan. & W.H. Hsieh, Mycol. Res. 93(3): 342 (1989) MycoBank MB 809108 Note. No ex-type sequence data in GenBank (2014). 60. Stachybotrys terrestris J.H. Kong & T.Y. Zhang, in Kong, Zhang & Zhang, Mycosystema 26(2): 200, 2007 (Fig. 45) Note: Conidiophores straight or slightly curved, unbranched or rarely branched, subhyaline near the base, greyish brown above, verrucose, occasionally covered with large granules, 23–75 × 2–3 μm; phialides pale olivaceous-grey, smooth, 8.5–14.5× 3–4.5 μm; conidia clavate, or oblong, pale grey, smooth, 7–10 ×3–5 μm, biguttulate (Kong et al. 2007). It is similar to S. guttulispora in morphology and guttulation. The conidia of the latter are ellipsoid, olivaceous or greenish brown, smooth, biguttulate, 9–12 × 3.5–5 μm (Muhsin and AlHelfi 1981). No ex-type sequence data in GenBank (2014). 61. Stachybotrys thaxteri D.W. Li, Mycotaxon 115: 240, 2011 (Fig. 46) Note: Its conidiophores are colourless, up to 360 μm long and conidia 15.0–17.5×6.5–7.5 μm, oblong or Fungal Diversity Fig. 49 Stachybotrys waitakere (Whitton et al. 2001) ellipsoid, with a median constriction, and bearing undulating diagonal striations (Li 2011). Its phylogenetic relationship with other taxa of Stachybotrys remains unknown. No ex-type sequence data in GenBank (2014). 62. Stachybotrys theobromae Hansf., Proc. Linn. Soc. London 155: 45, 1943 [1942–43] (Fig. 47) Note: It is a well defined species and easy to differentiate from other species. Its conidia are black, ovate or limoniform, apiculate at the base, smooth, 20–28×15– 18 μm (Hansford 1943). No ex-type sequence data in GenBank (2014), but in this paper we selected ATCC 18877 as its epitype, so AF081479 (ITS) is the epitype sequence of S. theobromae. 63. Stachybotrys thermotolerans McKenzie, in Pinruan, McKenzie, Jones & Hyde, Fungal Diversity 17: 149, 2004 ≡ Stachybotrys ramosa Udaiyan, J. Econ. Taxon. Bot. 15: 641, 1992 [1991] [nom. inval., Arts 37.1 (see Art. 9.5), 37.3; nom. illegit., Art 53.1], [non Dorai & Vittal, Trans. Br. mycol. Soc. 87: 642, 1986[1987]]. Note: Udaiyan (1991) indicated that this species of Stachybotrys has an optimum temperature for growth of 43 °C and referred to it as a ‘thermophile’, but gave insufficient data to equate to the definition of a thermophile (Mouchacca 1997). Since S. ramosa Udaiyan (1991) is a later homonym of S. ramosa Dorai & Vittal (1987), the epithet of former species is invalid. Thus, McKenzie proposed the new name, S. thermotolerans McKenzie. No ex-type sequence data in GenBank (2014). Fungal Diversity Fig. 50 Stachybotrys xigazenensis (Wu and Zhang 2011) 64. Stachybotrys variabilis H.F. Wang & T.Y. Zhang, Mycosystema 28: 23, 2009 (Fig. 48) Note: Conidia variable in shape, ovoid, oblong, globose, subsphaerical to ellipsoidal, pale brown to dark brown, smooth or coarsely roughened, variable in size, 4–20×3–13 μm (Wang and Zhang 2009). No ex-type sequence data in GenBank (2014). 65. Stachybotrys verrucispora Matsush., Matsush. Mycol. Mem. 4: 18, 1985 Note. Conidia ellipsoid or obovoid, tuberculate, 10– 15×9.5–11 μm, dark brown (Matsushima 1985). See S. nielamuensis for additional comments and comparison. No ex-type sequence data in GenBank (2014). 66. Stachybotrys virgata Krzemien. & Badura, Acta Soc. Bot. Pol. 23: 759, 1954 Note. Conidiophores erect, 100 × 4–4.5 μm, unbranched, smooth at the lower half, brown and rough at the upper half; phialides ovoid, colourless, 13× 6.5 μm; conidia ellipsoid, brown with colourless bands, 12–13 × 5–5.5 μm. The character of conidial Fungal Diversity Fig. 51 Stachybotrys yunnanensis (Kong 1997) ornamentation can differentiate it from other species. No ex-type sequence data in GenBank (2014). 67. Stachybotrys waitakere Whitton, McKenzie & K.D. Hyde, N.Z. Jl Bot. 39: 497, 2001 (Fig. 49) Note. Conidia ellipsoidal to broadly ellipsoidal, apex rounded, base rounded or with a broad truncate papilla, black, verrucose, 14.5–19×8–11.5 μm (Whitton et al. 2012). The conidia of S. freycinetiae are smaller (10– 15 × 3.5–5 μm), more coarsely verrucose than S . w ai t ak e re ( M c K en z i e 1 9 9 1 ) . C o n i d i a o f S. verrucispora are similar in size (11–16.5 × 8– 11 μm), but are tuberculate, and conidiophores (80– 235 × 7–12.5 μm) are much longer and wider (Matsushima 1985). No ex-type sequence data in GenBank (2014). 68. Stachybotrys xanthosomatis [as xanthosomae] Mercado & J. Mena, Acta Bot. Cubana 55: 4, 1988 Note. This species is considered illegitimate (MycoBank 2014). The database lists Stachybotrys xanthosomae B. Huguenin as a homonym that has priority. However, literature information on Stachybotrys xanthosomae B. Huguenin was not available in MycoBank. The publication list of (Huguenin 2013) was used to check Stachybotrys xanthosomae B. Huguenin in his papers published in the 1960’s and 1970’s, but publication of this name could not be verified. Instead, he published Stemphylium xanthosomatis B. Huguenin [as ‘xanthosomae’] and Bipolaris xanthosomatis B. Huguenin [as ‘xanthosomae’] in 1966, not Stachybotrys xanthosomae B. Huguenin. Thus, S. xanthosomatis Mercado & J. Mena is a valid species. It has ellipsoid or cylindrical, olivaceous brown or dark brown, verrucose, 10–19×5–6.5 μm conidia and rough, branched conidiophores (Mercado-Sierra and Fungal Diversity Fig. 52 Stachybotrys zeae (Morgan-Jones and Karr 1976) Mena-Portales 1988). No ex-type sequence data in GenBank (2014). 69. Stachybotrys xigazenensis Y.M. Wu & T.Y. Zhang, Mycotaxon 114: 461, 2011 (Fig. 50) Note. Conidia ovoid, ellipsoid or oblong, tuberculate, brown to dark brown, 9–12.5×7.5–10 μm (Wu and Zhang 2011). The conidial ornamentation is rather prominent. See S. nielamuensis for additional comments and comparison. No ex-type sequence data in GenBank (2014). 70. Stachybotrys yunnanensis H.Z. Kong, Mycotaxon 62: 427, 1997 (Fig. 51) Note: Conidia are cylindrical or subcylindrical, black, smooth or sometimes rough, 7–11×3.5–5 μm. The original illustration showed that some conidia are ellipsoid (Kong 1997). The delineation of this taxon from Fungal Diversity Fig. 53 Stachybotrys zhangmuensis (Wu and Zhang 2009) S. chartarum needs to be further studied. No ex-type sequence data in GenBank (2014). 71. Stachybotrys zeae Morgan-Jones & Karr, Mycotaxon 4: 510, 1976 (Fig. 52) Note. Conidia are ellipsoidal, brown to dark brown, verruculose, 7–9×3.5–4.5 μm (Morgan-Jones and Karr 1976). Any difference from S. chartarum is subtle and further study is necessary. The authors indicated that the type was deposited to BPI without an accession number (Morgan-Jones and Karr 1976). However, BPI has no record of it in its collection. 72. Stachybotrys zhangmuensis Y.M. Wu & T.Y. Zhang, Mycotaxon 109: 463, 2009 (Fig. 53) Note: Conidiophores straight or slightly curved, unbranched or rarely branched, 1–3-septate, subhyaline near the base, greyish brown above, verrucose, covered with large granules, 100–120 μm long, 6–9 μm wide near the base. Phialides borne in groups of 6–7, pale olive-grey, verrucose, 8–12×5–6 μm. Conidia ovoid, ellipsoid or oblong, tuberculate, greyish brown, 8–9.5× 6–7 μm (Wu and Zhang 2009). See S. nielamuensis for additional comments and comparison. No ex-type sequence data in GenBank (2014). 73. Stachybotrys zingiberis (V. Rao) Yong Wang bis, K.D. Hyde, McKenzie, Y.L. Jiang & D.W. Li, comb. nov. (Fig. 54) ≡ Memnoniella zingiberis V. Rao, Sydowia, 16(1–6): 43, 1963 Fungal Diversity Fig. 54 Stachybotrys zingiberis (Rao 1962) MycoBank MB 809109 Note. This species is probably a synonym of S. echinata. Its conidia are 4.4–6.8 μm and it was proposed based on its short conidiophores (38–50 μm) and size of phialides (6.3–14.7×4.2–5.2 μm) (Rao 1962). However, no conclusion should be drawn prior to examining the type. No ex-type sequence data in GenBank (2014). 74. Stachybotrys zuckii K. Matsush. & Matsush., Matsush. Mycol. Mem. 8: 53, 1995 Note: This is a species that develops conidia in both Fungal Diversity Fig. 55 Stachybotrys clitoriae. Reproduced from Bastista et al. (1960) slimy masses and in chains. Its phylogenetic relationship with species or isolates which develop similar dimorphic conidia, such as M. longistipitata D.W. Li et al. and the isolates of S. echinata should be studied in the future. No ex-type sequence data in GenBank (2014). Doubtful species 1. Stachybotrys alternans var. atoxica Pidopl., 1946 [nom. inval., Art. 36.1] Note. This is an invalid variety. It was also invalidly published in 1953 (Pidoplichko 1953) due to no Latin diagnosis. Most mycologists do not accept any varieties or other taxonomical names at rank below species proposed in the genus Stachybotrys. 2. Stachybotrys alternans var. jateli Pidopl., 1946 [nom. inval., Art. 36.1] Note. As with S. alternans var. atoxica, this is an invalid variety due to no Latin diagnosis. Most mycologists do not accept any varieties or other taxonomical names at rank below species proposed in the genus Stachybotrys. 3. Stachybotrys clitoriae Bat. & Peres, Publções Inst. Micol. Recife 298: 19, 1960 (Fig. 55) Note: It is not a species of Stachybotrys. According to the illustration of this species, it seems to belong to Periconiella. Its placement should be determined, when its type is available for examination. This is another species, described by Batista et al. (1960), which should be excluded from Stachybotrys due to the presence of ramoconidia. Its original description and illustration Fungal Diversity Fig. 56 Stachybotrys elongata. Reproduced from Peck (1890) (Fig. 56) would suggest that this fungus likely belongs to Periconiella. Since the type specimen (IMUR 19393) is not available for examination, its placement remains uncertain. 4. Stachybotrys complementi W. Miyazaki, H. Tamaoka, M. Shinohara, H. Kaise, T. Izawa, Y. Nakano, T. Kinoshita, K. Hong & K. Inoue, Microbiol. Immunol. 24(11): 1097, 1980 [nom. inval., Art. 34.1(c)] Note. It is an invalid species due to no Latin diagnosis. 5. Stachybotrys elongata Peck 1890 This is a problematic species. Pound and Clements ( 1 89 6 ) tr a n sf e rr e d S t a c h y bo t r y s e l o n g a t a t o Sterigmatobotrys. However, Sterigmatobotrys develops penicillate conidiophores and none phialidic, multicelled conidia. Réblová and Seifert (2011) opinioned that this species does not likely relate to Sterigmatobotrys and it is perhaps better placed in Aspergillus or retained in Memnoniella, or Stachybotrys. Examination of type specimen of this species showed that its conidiogenous cells were not phialidic. Instead, they were sympodial with extended necks without scars or denticles (Fig. 3). This fungus fits Civisubramaniania and thus, a new combination is proposed below: 6. Civisubramaniania elongata (Peck) D.W. Li & W.B. Kendr. comb. nov. (Fig. 56) ≡ Stachybotrys elongata Peck, Ann. Rep. N.Y. St. Mus. 43: 75 (1890) ≡ Sterigmatobotrys elongata (Peck) Pound & Clem., Minn. Bot. Stud. 1 (Bulletin 9): 667 (1896) MycoBank MB 809110 Specimen Examined: USA, NEW YORK, Manor, on dead twigs of Acer rubrum L., September 1889, Charles H. Peck, (NYS# 1081, holotype). Notes: Civisubramaniania Vittal & Dorai is monotypic and typified with Civisubramaniania eucalypti Vittal & Dorai (Vittal and Dorai 1986). This new combination adds the second species to the genus. 7. Stachybotrys lunzinensis [as lunzinense] Svilv., Zentbl. Bakt. ParasitKde, Abt. II 103: 182, 1941 Note: Verona and Mazzucchetti (1968) considered it a doubtful species. The conidia are hyaline or dark coloured, cylindrical, 5.9–10.2×2.2–5.2 μm and produced from a whorl of phialides (von Szilvinyi 1941). 8. Stachybotrys parva [as parvum] R.S. Dwivedi & B.P. Singh, Proc. 56th Indian Sci. Congr. 3: 305, 1969 [?1970–1971] [nom. inval., Art. 34.1(c)] Note: It is an invalid species due to no Latin diagnosis. Conidiophores 36–57.6 × 3.6 μm; conidia 3.6–7.2 × 3.6 μm. 9. Stachybotrys setosa M. Sierra, M. Calduch & Gené Note: MycoBank (2014) listed this species, but the name appears to be unpublished. A literature search failed to find it. Thus, the validity of this species cannot be determined at present. Fungal Diversity Acknowledgments Authors are appreciative to Dr. John Haines for allowing us access to type specimen of Stachybotrys elongata. The authors are also very grateful to Drs. Walter Gams, Rafael F. Castañeda Ruíz, Hongxin Mao, Weidong Wu, and Ariunaa Jalsrai for their assistance in finding literatures, especially the ones in Japan, Russia, and Ukraine. Without their assistance some references would not be available to us. This project was supported by the International Scientific Cooperated Project of Guizhou Province (No [2013]7004). References Abbas S, Katelaris CH, Singh AB et al (2012) World allergy organization study on aerobiology for creating first pollen and mold calendar with clinical significance in Islamabad, Pakistan: a project of World Allergy Organization and Pakistan Allergy, Asthma & Clinical Immunology Centre of Islamabad. World Allergy Organ J 5(9):103–110 Abdeljalil S, Hmad IB, Saiba W, Amouri B, Maalej, Kaaniche M, Koubaa A, Gargouri A (2014) Investigations on hydrolytic activities from Stachybotrys microspora and their use as an alternative in yeast DNA extraction. Appl Microbiol Biotechnol 172(3): 1599–1611 Akamatsu N, Fueta Y, Endo Y et al (2001) Decreased susceptibility topentylenetetrazol-induced seizures after low-frequency transcranial magneticstimulation in rats. Neurosci Lett 310:153–156 Al-Ahmad M, Manno M, Ng V, Ribeiro M, Liss G, Tarlo S (2010) Symptoms after mould exposure including Stachybotrys chartarum, and comparison with darkroom disease. Allergy 65(2):245–255 Amuzie CJ, Islam Z, Kim JK, Seo JH, Pestka JJ (2010) Kinetics of satratoxin g tissue distribution and excretion following intranasal exposure in the mouse. Toxicol Sci 116:433–440 Andersen B, Nielsen KF, Jarvis BB (2002) Characterization of Stachybotrys from water-damaged buildings based on morphology, growth, and metabolite production. Mycologia 94(3):392–403 Andersen B, Nielsen KF, Thrane U, Szaro T, Taylor JW, Jarvis BB (2003) Molecular and phenotypic descriptions of Stachybotrys chlorohalonata sp. nov. and two chemotypes of Stachybotrys chartarum found in water-damaged buildings. Mycologia 95(6): 1227–1238 Archambault P, Roff JC, Bourget E, Bang B, Ingram GR (1998) Nearshore abundance of zooplankton in relation to coastal topographical heterogeneity and mechanisms involved. J Plankton Res 20:671–690 Barr M (1990) Prodromus to nonlichenized, pyrenomycetous members of class Hymenoascomycetes. Mycotaxon 39:43–184 Barron G (1968) The genera of hyphomycetes from soil. Williams & Wilkins Company, Baltimore Batista A (1961) Um pugilo do gêneros novos de líquens imperfeitos. Publicações Instituto de Micologia da Universidade do Recife & Instituto Nacional de Pesquisas da Amazonia, Conselho Nacional de Pesquisas 320:1–31, 35 figs Batista AC, Bezerra JL, Peres GEP (1960) Singera n. gen. e outros fungos Moniliales. Publicações Inst Micol Univ Rsecife 298:1–33 Begerow D, Nilsson H, Unterseher M, Maier W (2010) Current state and perspectives of fungal DNA barcoding and rapid identification procedures. Appl Microbiol Biotechnol 87(1):99–108 Benyagoub M, Jabaji-Hare SH (1992) Parasitism of hyphae and sclerotia of Rhizoctonia solani by Stachybotrys elegans. Phytopathology 82: 1119 Bhan U, Newstead MJ, Zeng X, Ballinger MN, Standiford LR, Standiford TJ (2011) Stachybotrys chartarum-induced hypersensitivity pneumonitis is TLR9 dependent. Am J Pathol 179(6):2779– 2787 Bidochka MJ, De Koning J (2001) Are teleomorphs really necessary?: modelling the potential effects of Muller’s Ratchet on deuteromycetous entomopathogenic fungi. Mycol Res 105(08): 1014–1019 Bisby G (1943) Stachybotrys. Trans Br Mycol Soc 26(3):133–143 Bisby G, Ellis M (1949) Stachybotrys dichroa Grove. Trans Br Mycol Soc 32(2):158–161 Booth C (1957) Studies on Pyrenomycetes. I. Four species of Chaetosphaeria, two with Catenularia conidia. II. Melanopsamma pomiformis and its Stachybotrys conidia. Mycol Pap 68:1–27 Brasel TL, Martin JM, Carriker CG, Wilson SC, Straus DC (2005) Detection of airborne Stachybotrys chartarum macrocyclic trichothecene mycotoxins in the indoor environment. Appl Environ Microbiol 71(11):7376–7388 Cannon PF, Kirk PM (2007) Fungal families of the world. CABI Publishing, Oxfordshire Carey SA, Plopper CG, Hyde DM, Islam Z, Pestka JJ, Harkema JR (2012) Satratoxin-G from the black mold Stachybotrys chartarum induces rhinitis and apoptosis of olfactory sensory neurons in the nasal airways of rhesus monkeys. Toxicol Pathol 40:887–898 Carmichael JW, Kendrick WB, Conners I, Sigler L (1980) Genera of hyphomycetes. University Alberta Press, Edmonton Castañeda RF (1986) Fungi cubenses rev. Jard. Bot. Nac., Univ. Habana Castlebury LA, Rossman AY, Sung GH, Hyten AS, Spatafora JW (2004) Multigene phylogeny reveals new lineage for Stachybotrys chartarum, the indoor air fungus. Mycol Res 108(8):864–872 Centers for Disease Control and Prevention (2012) Facts about Stachybotrys chartarum and Other Molds. Retrieved August 04, 2012, from http://www.cdc.gov/mold/stachy.htm Chamoun R, Jabaji S (2011) Expression of genes of Rhizoctonia solani and the biocontrol Stachybotrys elegans during mycoparasitism of hyphae and sclerotia. Mycologia 103(3):483–493 Chamoun R, Aliferis KA, Jabaji SH (2013) Characterization and transcriptional regulation of mitogenactivated-protein kinase gene s following mycoparasitism and starvation conditions. Curr Genet 59(1–2):43–54 Chaverri P, Samuels GJ, van Wetenschappen KNA (2003) Hypocrea/ Trichoderma (Ascomycota, Hypocreales, Hypocreaceae): Species with green ascospores. Centraalbureau voor Schimmelcultures Utrecht Chung YJ, Copeland LB, Doerfler DL, Ward MDW (2010) The relative allergenicity of Stachybotrys chartarum compared to house dust mite extracts in a mouse model. Inhal Toxicol 22(6):460–468 Corda A (1837) Icones fungorum hucusque cognitorum, vol 1. J.G. Calve, Prague Crous PW, Shivas RG, Quaedvlieg W, van der Bank M, Zhang Y, Summerell BA, Guarro J, Wingfield MJ, Wood AR, Alfenas AC, Braun U, Cano-Lira JF, García D, Marin-Felix Y, Alvarado P, Andrade JP, Armengol J, Assefa A, den Breeÿen A, Camele I, Cheewangkoon R, De Souza JT, Duong TA, Esteve-Raventós F, Fournier J, Frisullo S, García-Jiménez J, Gardiennet A, Gené J, Hernández-Restrepo M, Hirooka Y, Hospenthal DR, King A, Lechat C, Lombard L, Mang SM, Marbach PA, Marincowitz S, Marin-Felix Y, Montaño-Mata NJ, Moreno G, Perez CA, Pérez Sierra AM, Robertson JL, Roux J, Rubio E, Schumacher RK, Stchigel AM, Sutton DA, Tan YP, Thompson EH, van der Linde E, Walker AK, Walker DM, Wickes BL, Wong PT, Groenewald JZ (2014) Fungal planet description sheets: 214–280. Persoonia - Mol Phylogenet Evol Fungi 32:184–306 Cruse M, Telerant R, Gallagher T, Lee T, Taylor JW (2002) Cryptic species in Stachybotrys chartarum. Mycologia 94(5):814–822 Deighton FC (1960) African fungi. I. Mycol Pap 78:1–43 Dorai M, Vittal BPR (1987) A new Stachybotrys from eucalyptus litter. Trans Br Mycol Soc 87(4):642–644 Elidemir O, Colasurdo GN, Rossmann SN, Fan LL (1999) Isolation of Stachybotrys from the lung of a child with pulmonary hemosiderosis. Pediatrics 104(4):964–966 Fungal Diversity Ellis MB (1971) Dematiaceous hyphomycetes. Commonwealth Mycological Institute, Kew Ellis MB (1976) More dematiaceous hyphomycetes. CABI Publishing, Kew, Surrey Etzel RA, Montana E, Sorenson W, Kullman GJ, Allan TM, Dearborn DG (1998) Acute pulmonary hemorrhage in infants associated with exposure to Stachybotrys atra and other fungi. Arch Pediatr Adolesc Med 152(8):757–762 Foarde KK, Menetrez MY (2002) Factors relating to the release of Stachybotrys chartarum spores from contaminated sources. In: Levin H (ed) Indoor Air 2002. Proceedings of International Conference on Indoor Air Quality and Climate, Santa Cruz Forgacs J (1972) Stachybotryotoxicosis. In: Kadis S, Ciegler A, Ajl S (eds) Microbial toxins, vol 8, Fungal toxins. Academic, New York, pp 95–128 Forgacs J, Carll WT, Herring AS, Hinshaw WR (1958) Toxicity of Stachybotrys atra for animals. N Y Acad Sci 20:787–808 Frazer S, Pestka J, Kim JK, Medina A, Aldred D, Magan N (2012) Impact of environmental factors on growth and satratoxin G production by strains of Stachybotrys chartarum. World Mycotoxin J 5(1):37–43 Fu X, Lindgren T, Guo M, Cai GH, Lundgren H, Norbäck D (2013) Furry pet allergens, fungal DNA and microbial volatile organic compounds (MVOCs) in the commercial aircraft cabin environment. Environ Sci Processes Impacts 15:1228–1234 Geiser DM, Jiménez-Gasco MM, Kang S, Makalowska I, Veeraraghavan N, Ward TJ, Zhang N, Kuldau GA, O’Donnell K (2004) FUSARIUM-ID v. 1.0: A DNA sequence database for identifying Fusarium. Eur J Plant Pathol 110:473–479 GenBank (2014) http://www.ncbi.nlm.nih.gov/GenBank/ Gilly M, Benson NR et al (1985) Affinity labeling the ribosome with eukaryotic-specific antibiotics: (bromoacetyl) trichodermin. Biochemistry 24(21):5787–5792 Grove WB (1886) New or noteworthy fungi. Part III. J Bot Br Foreign 24: 197–207 Hansford CG (1943) Cotributions towards the fungus flora of Uganda, V. Fungi imperfecti. Proc Linn Soc 1:34–67 Haugland RA, Heckman JL (1998) Identification of putative sequence specific PCR primers for detection of the toxigenic fungal species Stachybotrys chartarum. Mol Cell Probes 12:387–396 Haugland RA, Vesper SJ, Harmon SM (2001) Phylogenetic relationships of Memnoniella and Stachybotrys species and evaluation of morphological features for Memnoniella species identification. Mycologia 54:54–65 Hawksworth DL (2011) A new dawn for the naming of fungi: impacts of decisions made in Melbourne in July 2011 on the future publication and regulation of fungal names. IMA Fungus 2(2):155–162 Hebert PDN, Cywinska A, Ball SL, deWaard JR (2003) Biological identifications through DNA barcodes. Proc R Soc B 270:313–321 Hibbett DS, Taylor JW (2013) Fungal systematics: is a new age of enlightenment at hand? Nat Rev Microbiol 11(2):129–133 Hinkley SF, Jiang J, Mazzola EP, Jarvis BB (1999) Atranones, novel diterpenoids from the toxigenic mold, Stachybotrys atra. Tetrahedron Lett 40:2725–2728 Hinkley SF, Mazzola EP, Fettinger JC, Lam YKT, Jarvis BB (2000) Atranones A-G, from the toxigenic mold Stachybotrys chartarum. Phytochemistry 55:663–673 Hinkley SF, Moore JA, Squillari J, Tak H, Oleszewski R, Mazzola EP, Jarvis BB (2003) New atranones from the fungus, Stachybotrys chartarum. Magn Reson Chem 47:337–343 Hodgson MJ, Morey P, Leung WY, Morrow L, Miller D, Jarvis BB, Robbins H, Halsey JF, Storey E (1998) Building-associated pulmonary disease from exposure to Stachybotrys chartarum and Aspergillus versicolor. J Occup Environ Med 40(3):241–249 Hossain MA, Ahmed MS, Ghannoum MA (2004) Attributes of Stachybotrys chartarum and its association with human disease. J Allergy Clin Immunol 113:200–208 Hu FB, Willett WC, Li T, Stampfer MJ, Colditz GA, Manson JE (2004) Adiposity as compared with physical activity in predicting mortality among women. N Engl J Med 351(26):2694–2703 Hughes SJ (1952) Fungi from the Gold Coast. I. Mycol Pap 48:1–91 Hughes SJ (1958) Revisiones hyphomycetum aliquot cum appendice de nominibus rejiciendis. Can J Bot 36(6):727–836 Huguenin B (2013) List of publications and reports of Bernard Huguenin. Hyde KD, Goh TK, Taylor JE, FrÖHlich J (1999) Byssosphaeria, Chaetosphaeria, Niesslia and Ornatispora gen. nov., from palms. Mycol Res 103(11):1423–1439 Hyde KD, Jones EG, Liu J-K, Ariyawansa H, Boehm E, Boonmee S, Braun U, Chomnunti P, Crous PW, Dai D-Q (2013) Families of Dothideomycetes. Fungal Divers 63(1):1–313 Index Fungorum (2014) Index Fungorum. http://www.indexfungorum. org. Accessed 9 April 2014 Islam Z, Harkema JR, Pestka JJ (2006) Satratoxin G from the black mold Stachybotrys chartarum evokes olfactory sensory neuron loss and inflammation in the murine nose and brain. Environ Health Perspect 114:1099–1107 Islam Z, Shinozuka J, Harkema JR, Pestka JJ (2009) Purification and comparative neurotoxicity of the trichothecenes satratoxin G and roridin L2 from Stachybotrys chartarum. J Toxicol Environ Health A 72(20):1242–1251 Jarvis BB (1986) Potential indoor air pollution problems associated with macrocyclic trichothecene producing fungi. In: Health and Welfare Canada Working Group on Fungi and Indoor Air (ed) Significance of fungi in indoor air: report of a Working Group. Health and Welfare Canada, Ottawa Jarvis BB (1991) Macrocyclic trichothecenes. In: Sharma RP, Salunkhe DK (eds) Mycotoxins and phytoalexins. CRC Press, Boca Raton, pp 361–421 Jarvis BB (2002) Chemistry and toxicology of molds isolated from waterdamaged buildings. Adv Exp Med Biol 504:43–52 Jarvis BB (2003) Analysis for mycotoxins: the chemist’s perspective. Arch Environ Health 58(8):479–483 Jeewon R, Liew E, Hyde K (2004) Phylogenetic evaluation of species nomenclature of Pestalotiopsis in relation to host association. Fungal Divers 17:39–55 Jiang Y, Zhang T (2009) Notes on soil dematiaceous hyphomycetes from Sichuan Province, China. Mycosystema 28(5):644–647 Jie C-Y, Geng K, Jiang Y-L, Xu J-J, Hyde K, McKenzie EC, Zhang T-Y, Bahkali A, Li D-W, Wang Y (2013) Stachybotrys from soil in China, identified by morphology and molecular phylogeny. Mycol Prog 12(4):693–698 Jong S, Davis E (1976) Contribution to the knowledge of Stachybotrys and Memnoniella in culture. Mycotaxon 3(3):409–485 Judd WS, Campbell CS, Kellogg EA, Stevens PF, Donoghue MJ (2002) Plant systematics: a phylogenetic approach, 2nd edn. Sinauer Associates, Sunderland, pp 1–39 Karandikar K, Kulkarni S, Patwardhan P (1992) Some new and interesting Hyphomycetes from India. Biovigyanam 18(2):78–81 Keshava Prasad T, Asha L, Bhat D (2003) A new species of Memnoniella from India. Mycotaxon 85:341–344 Kirk P, Cannon PF, Minter D, Stalpers J (2008) Dictionary of the fungi, 10th edn. CAB International, Wallingford Kong H-Z (1997) Stachybotrys yunnanensis sp. nov. and Neosartorya delicata sp. nov. isolated from Yunnan, China. Mycotaxon 62:427– 434 Kong J, Zhang T, Zhang W (2007) Notes on soil dematiaceous hyphomycetes from Hexi Corridor, Gansu Province. Mycosystema 26(2): 196–201 Kordula T, Banbula A, Macomson J, Travis J (2002) Isolation and properties of stachyrase A, a chymotrypsin-like serine proteinase from Stachybotrys chartarum. Infect Immun 70(1):419–421 Fungal Diversity Koster B, Scott J, Wong B, Malloch D, Straus N (2003) A geographically diverse set of isolates indicates two phylogenetic lineages within Stachybotrys chartarum. Can J Bot 81(6):633–643 Koster B, Wong B, Straus N, Malloch D (2009) A multi-gene phylogeny for Stachybotrys evidences lack of trichodiene synthase (tri5) gene for isolates of one of three intrageneric lineages. Mycol Res 113: 877–886 Kuhn DM, Ghannoum MA (2003) Indoor mold, toxigenic fungi, and Stachybotrys chartarum: infectious disease perspective. Clin Microbiol Rev 16:144–172 Lechat C, Hairaud M, Lesage-Meessen L (2013) Stachybotrys oleronensis Lechat, Hairaud & Lesage-Messen, sp. nov., Fungal Planet 208. Persoonia 31:282–283 Li DW (2007) Stachybotrys eucylindrospora, sp. nov. resulting from a reexamination of Stachybotrys cylindrospora. Mycologia 99(2):332– 339 Li DW (2011) Stachybotrys thaxteri sp. nov. and the nomenclatural status of three Stachybotrys species. Mycotaxon 115(1):239–250 Li QR, Jiang YL (2011) Stachybotrys subreniformis, new from soil in China. Mycotaxon 115(1):171–173 Li DW, Yang CS (2004a) Fungal contamination as a major contributor to sick building syndrome. Adv Appl Microbiol 55:31–112 Li DW, Yang CS (2004b) Notes on indoor fungi I: new records and noteworthy fungi from indoor environments. Mycotaxon 89(2): 473–488 Li D-W, Yang CS (2005) Taxonomic history and current status of Stachybotrys chartarum and related species. Indoor Air 15(S9):5–10 Li DW, Yang CS, Haugiand R, Vesper S (2003) A new species of Memnoniella. Mycotaxon 85:253–257 Linnik YV (1949) On the theory of nonuniform Markov chains. Izv Akad Nauk SSSR 13(1):65–94 Liu YJ, Hall BD (2004) Body plan evolution of ascomycetes, as inferred from an RNA polymerase II, phylogeny. Proc Natl Acad Sci U S A 101:4507–4512 Mena-Portales J, Heredia-Abarca G, Mercado-Sierra A, Hernández CIB, Mota RMA, Cornelio SAG (2009) Species of Stachybotrys Corda (anamorphic fungi) from tropical and subtropical regions of Mexico. Bol Soc Micol Madrid 33:7–23 Manoharachary C, Agarwal DK, Sureshkumar G, Kunwar IK, Sharath BK (2006) Memnoniella mohanramii sp. nov. and Zygosporium anupamvarmae sp. nov. from India. Indian Phytopathol 59(4): 489–491 Marchal E (1895) Champignons copropiles de Belgique. VII. Bull Soc R Bot Belg 34:125–149 Matsushima T (1971) Microfungi of the Solomon Islands & Papua - New Guinea. Matsushima, Kobe Matsushima T (1975) Icones Microfungorum: a Matsushima lectorum. Matsushima, Kobe Matsushima T (1985) Matsushima mycological memoirs 4. Matsushima, Kobe Matsushima T (1989) Matsushima mycological memoirs 6. Matsushima, Kobe Matsushima T (1995) Matsushima mycological memoirs 8. Matsushima, Kobe Mayden RL (1997) A hierarchy of species concepts: the denoument in the saga of the species problem. In: Claridge MF, Dawah HA, Wilson MR (eds) Species: the units of diversity. Chapman and Hall, London, pp 381–423 McKenzie EHC (1991) Dematiaceous hyphomycetes on Freycinetia (Pandanaceae). I, Stachybotrys. Mycotaxon 41(1):179–188 Mercado-Sierra A, Mena-Portales J (1988) Nuevos o raros hifomicetes de Cuba. V. Especies de Stachybotrys. Acta Bot Cuba 55:1–8 Minagawa K, Kouzuki S, Yoshimoto J et al (2002a) Stachyflin and acetylstachyflin, novel anti-influenza A virus substances, produced by Stachybotrys sp. RF-7260. I. Isolation, structure elucidation and biological activities. J Antibiot 55:155–164 Minagawa K, Kouzuki S, Kamigauchi T (2002b) Stachyflin and acetylstachyflin, novel anti-influenza A virus substances, produced by Stachybotrys sp. RF-7260. II. Synthesis and preliminary structure-activity relationships of stachyflin derivatives. J Antibiot 55:165–171 Misra PC (1975) A new species of Stachybotrys. Mycotaxon 2(1):107– 108 Misra P (1976) Stachybotrys renispora sp. nov. Mycotaxon 4:161–162 Misra P, Srivastava S (1982) Two undescribed Stachybotrys species from Índia. Trans Br Mycol Soc 78(3):556–559 Morgan-Jones G, Karr G Jr (1976) Notes on hyphomycetes. XVI. A new species of Stachybotrys. Mycotaxon 4(2):510–512 Morgan-Jones G, Sinclair R (1980) Notes on hyphomycetes: XXXIII. Stachybotrys sphaerospora sp. nov. from South Africa. Mycotaxon 10:372–374 Morissette DC, Driscoll BT, Jabaji-Hare S (2003) Molecular cloning, characterisation, and expression of a cDNA encoding an endochitinase gene from the mycoparasite Stachybotrys elegans. Fungal Genet Biol 39:276–285 Mouchacca J (1997) Thermophilic fungi: biodiversity and taxonomic status. Cryptogam Mycol 18(1):19–69 Motohashi Y, Igarashi M, Okamatsu M, Noshi T, Sakoda Y, Yamamoto N, Ito K, Yoshida R, Kida H (2013) Antiviral activity of stachyflin on influenza A viruses of different hemagglutinin subtypes. Virol J 10:118 Muhsin TM, Al-Helfi M (1981) Hyphomycetes of Iraq—the genus Stachybotrys. Sydowia 34:130–134 Munjal R, Kapoor J (1969) Some hyphomycetes from the Himalayas. Mycopathologia 39(2):121–128 Munk A (1957) Danish pyrenomycetes. Dansk Bot Ark 17(1):1–491 MycoBank (2014) MycoBank (Fungal databases nomenclature and species banks). www.mycobank.org. Accessed 8 April 2014 Nagayoshi M, Tada Y, West J, Ochiai E, Watanabe A, Toyotome T, Tanabe N, Takiguchi Y, Shigeta A, Yasuda T (2011) Inhalation of Stachybotrys chartarum evokes pulmonary arterial remodeling in mice, attenuated by rho-kinase inhibitor. Mycopathologia 172:5–15 Nielsen K (2003) Mycotoxin production by indoor molds. Fungal Genet Biol 39:103–117 Nielsen KF, Huttunen K, Hyvärinen A, Andersen B, Jarvis BB, Hirvonen MR (2002) Metabolite profiles of Stachybotrys isolates from waterdamaged buildings and their induction of inflammatory mediators and cytotoxicity in macrophages. Mycopathologia 154(4):201–206 Nilsson RH, Hyde KD, Pawlowska J et al (2014) Improving ITS sequence data for identification of plant pathogenic fungi. Fungal Divers 67:11–19 Norback D, Cai G-H (2011) Fungal DNA in hotel rooms in Europe and Asia-associations with latitude, precipitation, building data, room characteristics and hotel ranking. J Environ Monitor 13(10):2895– 2903 Ochiai E, Kamei K, Hiroshima K, Watanabe A, Hashimoto Y, Sato A, Ando A (2005) The pathogenicity of Stachybotrys chartarum. Nippon Ishink Gakkai Zassi 46:109–117 O’Donnell K, Cigelnik E, Nirenberg HI (1998) Molecular systematics and phylogeography of the Gibberella fujikuroi species complex. Mycologia:465–493 Orpurt P (1954) Studies on the soil microfungi of Wisconsin prairies. Ph. D. Disseration, University of Wisconsin, Madison Peay KG, Kennedy PG, Bruns TD (2008) Fungal community ecology: a hybrid beast with a molecular master. Bioscience 58:799–810 Peck CH (1890) Report of the Botanist. Ann Rep N Y State Mus Nat Hist (1889) 43:51–97 Pei R, Gunsch CK (2013) Inflammatory cytokine gene expression in THP-1 cells exposed to Stachybotrys chartarum and Aspergillus versicolor. Environ Toxicol 28:51–60 Penttinen P, Pelkonen J, Huttunen K, Hirvonen MR (2006) Co-cultivation of Streptomyces californicus and Stachybotrys chartarum stimulates Fungal Diversity the production of cytostatic compound(s) with immunotoxic properties. Toxicol Appl Pharmacol 217:342–351 Pestka JJ, Yike I, Dearborn DG, Ward MDW, Harkema JR (2007) Stachybotrys chartarum, trichothecene mycotoxins, and damp building-related illness: new insights into a public health enigma. Toxicol Sci 104(1):4–26 Pestka JJ, Yike I, Dearborn DG, Ward MD, Harkema JR (2008) Stachybotrys chartarum, trichothecene mycotoxins, and damp building–related Illness: new insights into a public health enigma. Toxicol Sci 104(1):4–26 Photita W, Lumyong P, McKenzie EHC, Hyde KD, Lumyong S (2003) Memnoniella and Stachybotrys species from Musa acuminata. Crypt Mycol 24(2):147–152 Pidoplichko NM (1953) Fungus flora of coarse fodders. Izdatel’stvo Akademii Nauk Ukrainskoj SSR, Kiev Piecková E, Hurbánková M, Černá S, Lišková A, Kováčiková Z, Kolláriková Z, Wimmerová S (2009) Inflammatory and haematotoxic potential of indoor Stachybotrys chartarum (Ehrenb.) hughes metabolites. Arch Ind Hyg Toxicol 60(4):401–409 Pinruan U, McKenzie EHC, Jones EBG, Hyde KD (2004) Two new species of Stachybotrys, and a key to the genus. Fungal Divers 17: 145–157 Pound R, Clements FE (1896) A re-arrangement of the North American hyphomycetes. Minn Bot Stud 9:644–673 Rakkestad KE, Skaar I, Ansteinsson VE, Solhaug A, Holme JA, Pestka JJ, Samuelsen JT, Dahlman HJ, Hongslo JK, Becher R (2010) DNA damage and DNA damage responses in THP-1 monocytes after exposure to spores of either Stachybotrys chartarum or Aspergillus versicolor or to T-2 toxin. Toxicol Sci 115(1):140–155 Rao VC (1962) Some new records of fungi Imperfeeti from India. Sydowia 16(1–6):41–45 Réblová M, Seifert K (2011) Discovery of the teleomorph of the hyphomycete, Sterigmatobotrys macrocarpa, and epitypification of the genus to holomorphic status. Stud Mycol 68(1):193–202 Réblová M, Barr M, Samuels G (1999) Chaetosphaeriaceae, a new family for Chaetosphaeria and its relatives. Sydowia 51:49–70 Reeb V, Lutzoni F, Roux C (2004) Contribution of RPB2 to multilocus phylogenetic studies of the euascomycetes (Pezizomycotina, Fungi) with special emphasis on the lichen-forming acarosporaceae and evolution of polyspory. Mol Phylogenet Evol 32:1036–1060 Rifai M (1964) Stachybotrys bambusicola sp. nov. Trans Br Mycol Soc 47(2):269–272 Rossman AY (2007) Report of the planning workshop for all fungi DNA barcoding. Inoculum 58:1–5 Rossman AY, Mckemy JM, Pardo-Schultheiss RA, Schroers HJ (2001) Molecular studies of the Bionectriaceae using large subunit rDNA sequences. Mycologia 93:100–110 Saibi W, Amouri B, Gargouri A (2007) Purification and biochemical characterization of a transglucosilating β-glucosidase of Stachybotrys strain. Appl Microbiol Biotechnol 77(2):293–300 Saito K (1939) Studies on the fungi in the Orient (V) (東洋産菌類知見(第5 報)). J Ferment Technol (Osaka, Japan) 17(1):1–6 Sakurai J, Kikuchi T, Takahashi O, Watanabe K, Katoh T (2011) Enantioselective total synthesis of (+)-stachyflin: a potential antiinfluenza A virus agent isolated from a microorganism. Eur J Org Chem 2011(16):2948–2957 Samuels GJ, Barr ME (1997) Notes on and additions to the niessliaceae (hypocreales). Can J Bot 75(12):2165–2176 Schoch CL, Seifert KA, Huhndorf S, Robert V, Spouge JL, Levesque CA, Chen W, Fungal Barcoding Consortium (2012) Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi. Proc Natl Acad Sci U S A 109:6241–6246 Schroers HJ, Geldenhuis MM, Wingfield MJ, Schoeman MH, Yen YF, Shen WC, Wingfield BD (2005) Classification of the guava wilt fungus Myxosporium psidii, the palm pathogen Gliocladium vermoesenii and the persimmon wilt fungus Acremonium diospyri in Nalanthamala. Mycologia 97:375–395 Seifert KA, Morgan-Jones G, Gams W, Kendrick B (2011) The genera of hyphomycetes. Centraalbureau voor Schimmelcultures (CBSKNAW Fungal Diversity Centre), Utrecht Shanks OC, Kelty CA, Archibeque S, Jenkins M, Newton RJ, McLellan SL, Huse SM, Sogin ML (2011) Community structures of fecal bacteria in cattle from different animal feeding operations. Appl Environ Microbiol 77:2992–3001 Shelton BG, Kirkland KH, Flanders WD, Morris GK (2002) Profiles of airborne fungi in buildings and outdoor environments in the United States. Appl Environ Microbiol 68(4):1743–1753 Shenoy BD, Jeewon R, Hyde KD (2007) Impact of DNA sequence-data on the taxonomy of anamorphic fungi. Fungal Divers 26:1–54 Shenoy BD, Jeewon R, Wang H, Amandeep K, Ho WH, Bhat DJ, Crous PW, Hyde KD (2010) Sequence data reveals phylogenetic affinities of fungal anamorphs Bahusutrabeeja, Diplococcium, Natarajania, Paliphora, Polyschema, Rattania and Spadicoides. Fungal Divers 44:161–169 Smith G (1962) Some new and interesting species of micro-fungi. III. Trans Br Mycol Soc 45(3):387–394 Sorenson W, Frazer DG, Jarvis BB, Simpson J, Robinson V (1987) Trichothecene mycotoxins in aerosolized conidia of Stachybotrys atra. Appl Environ Microbiol 53(6):1370–1375 Straus D (2011) The possible role of fungal contamination in sick building syndrome. Front Biosci (Elite Ed) 3:562–580 Subramanian C (1954) Fungi Imperfecti from Madras-VI. Indian Bot Soc 33:36–42 Subramanian C (1957) Hyphomycetes-IV. Proc Plant Sci 46(5):324–335 Summerbell RC, Gueidan C, Schroers H-J, de Hoog GS, Starink M, van Iperen A, Arocha Rosete Y, Guarro J, Scott JA (2011) Acremonium phylogenetic overview and revision of Gliomastix, Sarocladium, and Trichothecium. Stud Mycol 68:139–162 Talbot PHB (1971) Principles of fungal taxonomy. The Macmillan Press, London, pp 1–274 Tang AMC, Jeewon R, Hyde KD (2007) Phylogenetic utility of protein (RPB2, β-tubulin) and ribosomal (LSU, SSU) gene sequences in the systematics of sordariomycetes (ascomycota, fungi). Anton Leeuw Int J G 91(4):327–349 Taylor JW (2011) One fungus= one name: DNA and fungal nomenclature twenty years after PCR. IMA Fungus Glob Mycol J 2(2):113–120 Taylor J, Jacobson D, Fisher M (1999) The evolution of asexual fungi: reproduction, speciation and classification. Annu Rev Phytopathol 37(1):197–246 Taylor JW, Geiser DM, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, Hibbett DS, Fisher MC (2000) Phylogenetic species recognition and species concepts in fungi. Fungal Genet Biol 31:21–32 Taylor G, Suha J, Pierre M, Wagahatullam K (2002) Purification and characterization of an extracellular exochitinase, B-Nacetylhexosaminidase, from the fungal mycoparasite Stachybotrys elegans. Can J Microbiol 48:311–319 Tripi PA, Modlin S, Sorenson W, Dearborn DG (2000) Acute pulmonary haemorrhage in an infant during induction of general anaesthesia. Pediatr Anesth 10(1):92–94 Tubaki K (1963) Notes on the Japanese Hyphomycetes. I. Chloridium, Clonostachys, Isthmospora, Pseudobotrytis, Stachybotrys and Stephanoma. Trans Mycol Soc Jpn 4:83–90 Tweddell RJ, Jabaji-Hare SH, Charest PM (1994) Production of chitinases and β-1,3-glucanases by Stachybotrys elegans, a mycoparasite of Rhizoctonia solani. Appl Environ Microbiol 60(2):489–495 Tweddell RJ, Marchall J, Jabaji-Hare SH (1996) Endo-1,3 b-glucanase production by Stachybotrys elegans, a mycoparasite of Rhizoctonia solani. Mycologia 88:410–415 Udaiyan K (1991) Some interesting hyphomycetes from the industrial water cooling towers of Madras. J Econ Taxon Bot 15(3):627–647 Fungal Diversity Verona O, Mazzucchetti G (1968) I Generi “Stachybotrys” E “Memnoniella”. Publicazioni Dell’Ente Nazionale, Per La Cellulosa E Per La Carta, Laboratorio Di Cartotecnica Speciale, Roma Vesper SJ, Vesper MJ (2002) Stachylysin may be a cause of hemorrhaging in humans exposed to Stachybotrys chartarum. Infect Immun 70(4):2065–2069 Vesper SJ, Magnuson ML, Dearborn DG, Yike I, Haugland RA (2001) Initial characterization of the hemolysin stachylysin from Stachybotrys chartarum. Infect Immun 69(2):912–916 Vittal BPR, Dorai M (1986) Civisubramaniania eucalypti gen. et sp. nov. Trans Br Mycol Soc 87(3):482–485 von Höhnel F (1923) Studien über Hphomyzeten. Centralb f Bakt Abt II LX 2:1–26 von Szilvinyi A (1941) Mikrobiologische Boden untersuchungen im Lunzer Gebiet. Zentr Bakteriol Parasit Abt II 103:133–189 Wang Y-Z (2011) Three species excluded from Melanopsamma (Ascomycetes). Mycotaxon 115(1):365–368 Wang HF, Zhang TY (2009) Notes on soil dematiaceous hyphomycetes from the Qaidam Basin, Qinghai Province, China. Mycosystema 28: 20–24 Whitton SR, McKenzie EHC, Hyde KD (2001) Microfungi on the pandanaceae: Stachybotrys, with three new species. N Z J Bot 39(3):489–499 Whitton SR, McKenzie EHC, Hyde KD (2012) Teleomorphic microfungi associated with Pandanaceae. In: Fungi associated with pandanaceae. Fungal diversity research series 21. Springer Netherlands, pp 23–124 Wu Y, Zhang T (2009) Two new species of Stachybotrys from soil. Mycotaxon 109:461–464 Wu Y-M, Zhang T-Y (2011) [2010] Two new species of Stachybotrys from soil. Mycotaxon 114(1):459–462 Wu H-X, Tian Q, Li WJ, Hyde KD (2014) A reappraisal of Microthyriaceaea. Phytotaxa 176(1):201–212 Ye TT, Huang JX, Shen YE, Lu PL, Christiani DC (1998) Respiratory symptoms and pulmonary function among Chinese rice-granary workers. Int J Occup Med Environ Health 4:155–159 Yike I, Dearborn D (2011) Guest editorial — novel insights into the pathology of Stachybotrys chartarum. Mycopathologia 172:1–3 Zuck RK (1946) Isolates intermediate between Stachybotrys and Memnoniella. Mycologia 38(1):69–76