Next Article in Journal
Encephalartos villosus Lem. Displays a Strong In Vivo and In Vitro Antifungal Potential against Candida glabrata Clinical Isolates
Next Article in Special Issue
Mucosal-Associated Invariant T Cells Accumulate in the Lungs during Murine Pneumocystis Infection but Are Not Required for Clearance
Previous Article in Journal
Metabolomics Analysis and Antioxidant Potential of Endophytic Diaporthe fraxini ED2 Grown in Different Culture Media
Previous Article in Special Issue
The Multifaceted Role of Mating Type of the Fungus and Sex of the Host in Studies of Fungal Infections in Humans
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Taxonomy, Phylogenetic and Ancestral Area Reconstruction in Phyllachora, with Four Novel Species from Northwestern China

1
International Fungal Research and Development Centre, Institute of Highland Forest Science, Chinese Academy of Forestry, Kunming 650224, China
2
Key Laboratory of Breeding and Utilization of Resource Insects, National Forestry and Grassl and Administration, Kunming 650224, China
3
College of Grassland Science, Shanxi Agricultural University, Mingxian South Road, Taigu District, Jinzhong 030801, China
4
Research Center of Microbial Diversity and Sustainable Utilization, Chiang Mai University, Chiang Mai 50200, Thailand
5
Centre for Yunnan Plateau Biological Resources Protection and Utilization, College of Biological Resource and Food Engineering, Qujing Normal University, Qujing 655011, China
6
Section of Genetics, Institute for Research and Development in Health and Social Care, No. 393/3, Lily Avenue, Off Robert Gunawardane Mawatha, Battaramulla 10120, Sri Lanka
*
Author to whom correspondence should be addressed.
Submission received: 15 April 2022 / Revised: 11 May 2022 / Accepted: 14 May 2022 / Published: 18 May 2022
(This article belongs to the Special Issue Women in Mycology)

Abstract

:
The members of Phyllachora are biotrophic, obligate plant parasitic fungi featuring a high degree of host specificity. This genus also features a high degree of species richness and worldwide distribution. In this study, four species occurring on leaf and stem of two different species of grass were collected from Shanxi and Shaanxi Provinces, China. Based on morphological analysis, multigene (combined data set of LSU, SSU, and ITS) phylogenetic analyses (maximum likelihood and Bayesian analysis), and host relationship, we introduce herein four new taxa of Phyllachora. Ancestral area reconstruction analysis showed that the ancestral area of Phyllachora occurred in Latin America about 194 Mya. Novel taxa are compared with the related Phyllachora species. Detailed descriptions, illustrations, and notes are provided for each species.

1. Introduction

Phyllachorales is an ascomycetous order (in Sordariomycetes) introduced by Barr [1]. This order comprises biotrophic, obligate plant parasitic fungi that infect mostly plant leaves and stems [2,3,4]. Species of Phyllachorales are mainly distributed in tropical and subtropical regions [5,6,7,8]. Members of Phyllachorales present with shiny black stromata, leading to the common name ‘tropical tar spot fungi’. Most of the members of Phyllachorales are parasitic on angiosperms with a few notable exceptions including the lichenicolous Lichenochora species, the marine algicolous genus Phycomelaina, as well as ferns and gymnosperms [8,9,10,11].
Currently, Phyllachorales has four families, including Phaeochoraceae, Phaeochorellaceae, Phyllachoraceae, and Telimenaceae [12,13]. They are morphologically characterized by black stromata of various shapes in the host plant; having paraphyses; unitunicate asci cylindrical to clavate in shape, with an inconspicuous apical ring, usually 8-spored; and aseptate ascospores, which in most species are hyaline and 1-celled, appearing as brown in a few species (e.g., Phyllachora stenostoma) [1,4,6,8,14]. The asexual morph of Phyllachorales has been reported as a coelomycetous morph [15]. Large-scale phylogenetic studies comprising many representative species have confirmed the position of Phyllachorales in the subclass Sordariomycetidae with high support (100% MLBP) as well as the monophyly of the order [4,8,16]. Mardones et al. [8] used three morphological characteristics and one ecological characteristic to reconstruct the ancestral state of genera in Phyllachorales based on the Likelihood Ancestral States method, reasoning that these characteristics had evolved independently numerous times. The ancestral state of members of Phyllachorales were monocotyledonous host plants with immersed perithecia, which was lost in the family Phaeochoraceae and evolved into erumpent or superficial perithecia in some species of Phyllachoraceae. The presence of clypeus as a morphological characteristic was lost only once in Phaeochoraceae. Therefore, it is thought that the presence of clypeus in these fungi is an evolutionarily stable characteristic. The ancestor of the Phyllachorales species had a black stroma, and the presence of bright black stromata may have evolved at least twice.
The family Phyllachoraceae was introduced by Theissen and Sydow [17] with Phyllachora as the type genus [3,18]. It is the largest family in Phyllachorales and currently comprises 54 genera [13]. Members of the family are characterized by forming leaf spots on the host that are abundant but scattered, raised, mostly rounded to oblong or elongated, sometimes parallel with leaf venation, surrounded by a light-brown necrotic region; lacking periphyses; having numerous paraphyses, branched or unbranched; 8-spored asci, persistent, cylindrical to fusiform, often present with an apical ring; ascospores fusiform to narrowly oval, hyaline, often with a mucilaginous sheath [4]. The type genus, Phyllachora, was introduced based on P. agrostis, which is a single species on the herbarium label in Fuckels exsiccate series ‘Fungi Rhenani’ [5]. Phyllachoraceae is similar to Phaeochoraceae, but Phaeochoraceae species are characterized by 6-8-spored asci, usually without apical structure, yellow to olivaceous ascospores or in various shades of brown, thick-walled; conversely, Phyllachoraceae species are characterized by 8-spored asci, an often-present an apical ring, usually hyaline ascospores, rarely pale brown, thin and smooth-walled [4,8]. These morphological characteristics can be used to distinguish the two families, and they form two independent branches in the phylogenetic tree [8].
Phyllachora is the type genus of Phyllachoraceae. Clements [19] designated the lectotype as Phyllachora graminis. Currently, Phyllachora is the largest genus within Phyllachoraceae, and about 1513 epithets are listed in the Index Fungorum (Index Fungorum 2022; accession date: 28.03.2022). Nevertheless, only 1382 species are accepted in the Species Fungorum (accession date: 28.03.2022). Species of the genus are morphologically characterized by clypeate pseudostroma in leaf tissues; generalized infection of the entire section of the mesophyll forming leaf spots on the host, mostly rounded to oblong or elongated, surrounded by a light-brown necrotic region; perithecium globose; numerous paraphyses, branched, slightly longer than asci; asci 8-spored, persistent, cylindrical to fusiform, short pedicellate, an apical ring often present; and ascospores 1–3 seriate, fusiform to narrowly oval, hyaline, sometimes with a gelatinous sheath [4,18,20]. Some members of the genus can inflict crop diseases, leading to yield loss. Phyllachora maydis is an example occurring in the United States, which can seriously impact quality and corn yield [21,22,23,24]. Owing to its biotrophic habit and high degree of host specificity, most Phyllachora species are given names based on host association and coevolution with the host [5,8,20,25]. Phyllachora species cannot grow on agar media since they are biotrophic [8]. Phyllachora species have been reported as pathogenic species on more than 1000 plant species (belonging to 121 families, including Cyperaceae, Fabaceae, Lauraceae, Moraceae, Myrtaceae, Poaceae, Proteaceae, and Rosaceae), and they are commonly found with Poaceae [20,26,27].
In this study, several specimens with tar spot diseases were collected. Based on polyphasic approaches (e.g., morphological analyses, information of host plant, and phylogenetic analyses), four novel species of Phyllachora are introduced herein. Based on paleontological evidence and paleoclimate records, we also reconstructed the ancestral area of Phyllachora. The analysis was restricted to members of Phyllachora, considering the history of their biogeographic diversity and dispersal route as well as estimating the divergence time and ancestral location of this genus.

2. Materials and Methods

2.1. Collecting, Morphological Study, and Depositing Specimens

Phyllachora-like fungi were collected from living leaves of Cenchrus flaccidus (Poaceae) and Chloris virgata (Poaceae) during field surveys in 2019 in Shanxi and Shaanxi Provinces, China. Specimens were taken to the laboratory in paper envelopes. Specimens were processed and examined with microscopes, and photos of ascomata and host were taken using a compound stereomicroscope (KEYENCE CORPORATION V.1.10 with camera VH-Z20R) following Wu et al. [28]. Hand sections were made under a stereomicroscope (OLYMPUS SZ61) and mounted in water and blue cotton, and photomicrographs of fungal structures were taken with a compound microscope (Nikon ECLIPSE 80i).
Images used for figures were processed with Adobe Photoshop CC v. 2015.5.0 software (Adobe Systems, San Jose, CA, USA).
Holotype collections were deposited at the herbarium of IFRD (International Fungal Research & Development Centre; Institute of Highland Forest Science, Chinese Academy of Forestry, Kunming, China). Registration numbers for new species were obtained in MycoBank Database (https://www.mycobank.org/, accession date: 10 April 2022).
For our specimens, no culture was obtained by multiple single-spore isolation or tissue isolation.

2.2. DNA Isolation, Amplification, and Sequencing

In accordance with the manufacturer’s instructions, genomic DNA was extracted from ascomata at room temperature using the Forensic DNA Kit (OMEGA, New York, NY, USA). The primers LR0R and LR5 were used to amplify the large subunit (LSU) rDNA [29]. The internal transcribed spacer (ITS) rDNA was amplified and sequenced with the primers ITS5 and ITS4 [30]. The partial small subunit (SSU) rDNA was amplified using primers NS1 and NS4 [30]. PCR reactions were in accordance with instructions from Golden Mix, Beijing TsingKe Biotech Co. Ltd, Beijing, China: initial denaturation at 98 °C for 2 min, followed by 30 cycles of 98 °C denaturation for 10 s, 56 °C annealing for 10 s and 72 °C extensions for 10 s (ITS and SSU) or 20 s (LSU), and a final extension at 72 °C for 1 min. All PCR products were sequenced by Biomed (Beijing Biomed Gen Technology Co., Ltd., Beijing, China). PCR products were sequenced by Biomed using the same primers as before.

2.3. Sequence Alignment and Phylogenetic Analyses

BioEdit version 7.0.5.3 [31] was used to re-assemble the sequences generated from forward and reverse primers for obtaining integrated sequences. Sequences of Phyllachoraceae species were downloaded from GenBank (Table 1) following the relevant publications [8,20,32,33]. All sequences were adjusted manually with Bioedit 7.0.5.3 [31] and aligned using the default setting of MAFFT version 7 online [34] (https://mafft.cbrc.jp/alignment/server/, accession date: 10 April 2022).
Maximum Likelihood (ML) analysis using the aligned sequences as input was conducted with the help of RAxNLGUI v. 2.0 [35]. Telimena bicincta (MM-108) and T. bicincta (MM-133) were selected as an outgroup. One thousand nonparametric bootstrap iterations were employed with the “ML + rapid bootstrap” tools and “GTRGAMMA” arithmetic.
For Bayesian analysis, MrModeltest 2.3 [36] was used to estimate the best-fitting model for the combined LSU, SSU, and ITS loci, and model GTR+G was the best fit. In MrBayes v.3.2 [37], six simultaneous Markov chains were run for 2,000,000 generations; trees were sampled and printed every 100 generations. The first 5000 trees were submitted to the burn-in phase and discarded, while the remaining trees were used for calculating posterior probabilities in the majority rule consensus tree [38,39,40,41].

2.4. Reconstruction of Ancestral State

Members of Phyllachora were coded based on their collection locality according to field notes and references. Six areas were delimited based on the distribution data of Phyllachora: A = East Asia, B = Southeast Asia, C = North America, D = South America, E = Latin America, F = Central Europe, G = Unknown, using species from Asia, Europe, North America, South America, Latin America, and Central Europe. In MrBayes v.3.2, chains were run for 1 00000 generations; trees were sampled and printed every 100 generations. RASP 4.2 (Reconstruct Ancestral State in Phylogenies, http://mnh.scu.edu.cn/soft/blog/RASP, accession date: 10 April 2022) was used to reconstruct the ancestral state, and the most-optimal model was DEC [42].

2.5. Calibration Procedure

The second calibration time referenced the results of Dayarathne et al. [33] and Hongsanan et al. [43]. We followed the conclusion that the family Phyllachoraceae divergence time was about 217 Mya as a calibration point (root) for ancestral distribution reconstruction.

3. Results

3.1. Molecular Phylogenetic Results

We analyzed a three-loci (LSU, SSU, ITS) data set of Phyllachora. Based on the combined data of LSU, SSU, and ITS sequences. It was found that the two topological trees obtained by maximum likelihood (ML) and Bayesian were similar, and the best scoring RAxML tree was used as the representative tree (Figure 1). We generated a total of 161 sequences from 74 taxa of Phyllachorales, 57 sequences of LSU, 34 sequences of SSU, 70 sequences of ITS, and concatenated sequences of three genes, with 3341 characters including gaps. Bootstrap values of ML higher than 50% are shown on the phylogenetic tree, while values of Bayesian posterior probabilities above 0.5 are shown on the tree (Figure 1). Phylogenetic analysis showed that all four new taxa belonging to Phyllachora cluster together with Phyllachora panicicola with bootstrap values of 68% (in ML analysis) and Bayesian posterior probability of 0.92. Phyllachora panicicola and four new taxa form two clades independent from each other with bootstrap values of 100% and Bayesian posterior probabilities of 1.00.

3.2. Ancestral Area Reconstruction Analysis for Phyllachora

Ancestral area reconstruction analysis revealed that Phyllachora species originated from Latin America about 194 Mya (Figure 2, node 145). Dispersal, vicariance, extinction, and other historical events affected the biogeographical distribution of the species. The evolutionary history of ancestors from the genus Phyllachora reveals that the species of this genus underwent 20 dispersals, 13 vicariances, and 1 extinction (Figure 2, blue coils represent dispersal, green coils represent vicariance, and orange coils represent extinction). Species of Phyllachora migrated from Latin America to Southeast Asia during the Jurassic period, with two dispersal events noted (Figure 2, node 127). In approximately 60–155 Mya, there were frequent dispersal and vicariance events, and moreover, vicariances were always accompanied by dispersal events. There is only low support suggesting that species belonging to Phyllachora may have migrated from Latin America to Southeast Asia 119 Mya (Figure 2, node 108). About 100 Mya, species migrated from East Asia or North America to central Europe, with one dispersal and one vicariance (Figure 2, node 77).

3.3. Taxonomy of Fungi

Phyllachora flaccidudis H. X. Wu & J. C. Li. sp. nov. (Figure 3 and Figure 4).
MycoBank: MB843395.
Etymology: Epithet derived from host species Cenchrus flaccidus.
Holotype: IFRD9445.
Parasitic on leaves and stems of Cenchrus flaccidus (Poaceae). Sexual morph: Stroma 1618–1900 × 641–764 μm ( x ¯ = 1769 × 716 μm, n = 10) (Figure 3a–c), fusiform or cymbaeform, domed above the leaf surface, amphigenous, scattered, sometimes gregarious, like black nevus, no edge, black, carbonaceous. Section of stroma 307–379 μm high, multilocular, peridium 35–60 μm wide, composed of brown to dark brown cells of textura angularis (Figure 3d). Paraphyses 2–3 μm wide ( x ¯ = 2.9 μm, n = 20), numerous, persistent, filiform, unbranched, aseptate, many guttules, slightly longer than asci (Figure 3e). Asci 82–110 × 7–10 μm ( x ¯ = 91.6 × 9.0 μm, n = 20), thin-walled, 8-spored, persistent, cylindrical to clavate, apex obtuse, with pedicel (Figure 3f–i). Ascospores 11–13 × 4–7 μm ( x ¯ = 11.4 × 5.9 μm, n = 20), 1-seriate, fusiform to oval, both ends obtuse, hyaline, aseptate, verrucose, with many guttules (Figure 3j–m) and a mucilaginous sheath (Figure 4a–d). Asexual morph: Not observed.
Material examined: China, Shanxi Province, Xinzhou, Wutai County, 38.71917699 N, 113.25921 E, on stems and leaves of Cenchrus flaccidus (Poaceae), 4 October 2019, Yuying Li IFRD9445, holotype. GenBank accession numbers: ITS: ON075524, LSU: ON072101, SSU: ON072097.
Notes: Phyllachora flaccidudis was collected from Cenchrus flaccidus (Poaceae) in Shanxi Province of China. According to the phylogenetic analysis, P. flaccidudis and P. sandiensis are closely related to P. panicicola (Figure 1). However, P. panicola was reported from Panicum sp., and its asci were significantly longer and wider than in P. flaccidudis and P. sandiensis (110–130 × 10–14 μm vs. 82–110 × 7–10 μm, 92–126 × 7–10 μm, respectively). Furthermore, the ascospores of P. panicola are significantly larger than in P. flaccidudis and P. sandiensis (14–16 × 6–8 μm vs. 11–13 × 4–7 μm, 10–14 × 6–7 μm, respectively). The ascospores of P. panicicola are rounded at both ends, with a central concave depression and lacking guttules different from P. flaccidudis and P. sandiensis. Differences in morphological characteristics are also supported by the phylogenetic tree, as the two species cluster independently (100% bootstrap support) in a subclade of Phyllachora (Figure 1).
The hosts of Phyllachora sphaerosperma (= Phyllachora cenchricola), P. flaccidudis and P. sandiensis belong to species of Cenchrus. However, the host of P. cenchricola is Cenchrus echinatus, which has been found in Brazil, the southern United States, South America, and the West Indies. The ascospores are nearly spherical, wider than both P. flaccidudis and P. sandiensis (Table 2).
Phyllachora flaccidudis and P. chloridis [33] have similar morphological characteristics, but their host plants are different. The host of P. flaccidudis and P. sandiensis was reported from Cenchrus flaccidus (Poaceae), while the host of P. chloridis is Chloris sp. Morphologically, the asci of P. flaccidudis and P. sandiensis are significantly longer than those of P. chloridis (Table 2) and feature pedicels, but they are absent in P. chloridis. Ascospores of P. flaccidudis and P. sandiensis have 1–2 or more guttules, while P. chloridis has only one central guttule, which can serve as an important characteristic for species delimitation.
Phyllachora sandiensis H. X. Wu & J. C. Li. sp. nov. (Figure 4 and Figure 5).
MycoBank: MB 843396.
Etymology: Epithet derived from the type locality, a sandy forest park roadside (Shaanxi Province, Yulin City, Yuyang District) in China.
Holotype: IFRD9446.
Parasitic on leaves and stems of Cenchrus flaccidus (Poaceae). Sexual morph: Stroma 993–2742 × 371–438 μm diam. ( x ¯ = 1378 × 415 μm, n = 10) (Figure 5a–c), domed above the leaf surface, amphigenous, fusiform, cymbaeform or irregular shape, like black nevus, scattered, sometimes gregarious, no edge, black, carbonaceous. Section of stroma 344–421 μm high, oval, multilocular, peridium 36–58 μm wide, composed of brown to dark-brown cells of textura angularis (Figure 5d). Paraphyses 2–4 μm wide ( x ¯ = 3 μm, n = 20), numerous, persistent, filiform, unbranched, aseptate, many guttules, slightly longer than asci (Figure 5e). Asci 92–126 × 7–10 μm ( x ¯ =106.6 × 9.7 μm, n = 20), thin–walled, 8-spored, persistent, cylindrical to clavate, apex obtuse, with pedicel (Figure 5f–i). Ascospores 10–14 × 6–7 μm ( x ¯ =12.3 × 6.3 μm, n = 20), 1-seriate, droplet or oval to ellipse, acute at the end, narrowly rounded at another end, hyaline, aseptate, verrucose, with 1–3 guttules (Figure 5j–m) and a thin mucilaginous sheath (Figure 4e–h). Asexual morph: Not observed.
Material examined: China, Shaanxi Province, Yulin City, Yuyang District, sandy forest park roadside, 38.29470799 N, 109.75387 E, on stems and leaves of Cenchrus flaccidus (Poaceae), 7 October 2019, Yuying Li, IFRD9446, holotype. GenBank accession numbers: LSU: ON075528, SSU: ON072098, ITS: ON075525.
Notes: Phyllachora sandiensis was collected from Cenchrus flaccidus in Shaanxi Province of China. According to the phylogenetic analysis, P. sandiensis is closely related to P. flaccidudis, but the samples were collected from different locales. Morphologically, the size of asci and stromata of P. flaccidudis is significantly longer than in the case of P. sandiensis (Table 2). In addition, the ascospores of P. sandiensis are longer than those of P. flaccidudis (average 106.6 × 9.7 μm vs. 91.6 × 9.0 μm). According to sequence alignment results, LSU, SSU, and ITS sequences differed by 9 bases, 5 bases, and 13 bases between both taxa, respectively. Therefore, P. sandiensis is considered to be a new species of Phyllachora.
Phyllachora virgatae H. X. Wu & J. C. Li. sp. nov. (Figure 4 and Figure 6).
MycoBank: MB843397.
Etymology: Epithet derived from host species Chloris virgata.
Holotype: IFRD9447.
Parasitic on leaves and stems of Chloris virgata (Poaceae). Sexual morph: Stroma 738–2678 × 490–701 μm ( x ¯ = 1762 × 642 μm, n = 10) (Figure 6a–c), domed above the leaf surface, amphigenous, fusiform, cymbaeform or irregular shape, black spots, scattered, sometimes gregarious, no edge, shiny black, carbonaceous. Section of stroma 215–232 μm high, oval, multilocular, peridium 25–32 μm wide, composed of brown to dark-brown cells of textura angularis (Figure 6d). Paraphyses 2–3.5 μm wide ( x ¯ = 3.3 μm, n = 20), numerous, persistent, filiform, unbranched, aseptate, slightly longer than asci (Figure 6e). Asci 84–120 × 7–11 μm ( x ¯ = 102.6 × 9.1 μm, n = 20), thin-walled 8-spored, persistent, clavate, apex obtuse, with pedicel (Figure 6f–i). Ascospores 10–15 × 6–9 μm ( x ¯ = 12.9 × 6.9 μm, n = 20), 1-seriate, ovoid to oblong, acute at both ends, hyaline, aseptate, verrucose, with 1–2 or more guttules (Figure 6j–m) and a glutinous mucilaginous sheath (Figure 4i–l). Asexual morph: Not observed.
Material examined: China, Shanxi Province, Xinzhou, Dingxiang County, 38.49269099 N, 112.94377 E, on stems and leaves of Chloris virgata (Poaceae), 4 October 2019, Yuying Li, IFRD9447, holotype. GenBank accession numbers: LSU: ON075439, SSU: ON072099, ITS: ON075526.
Notes: In the course of investigating the grass resources of northern China, two fungal species were collected from the Chloris virgata in Shanxi and Shaanxi Provinces. According to phylogenetic analysis, these two species are related to P. panicicola. They can be easily distinguished from P. panicicola based on a different host plant, larger asci (110–130 × 10–14 μm vs. 84–120 × 7–11 μm, 77–114 × 8–12 μm), and ascospore size (14–16 × 6–8 μm vs. 10–15 × 6–9 μm). Furthermore, the ascospores of P. panicicola are lacking guttules and have one central concave depression, differing from P. virgatae and P. jiaensis.
Phyllachora cynodontis and P. koondrookensis have been reported from the same host (i.e., Chloris) [14,46]. However, P. virgatae and P. jiaensis are clearly distinguishable from the two species (Table 2). Phyllachora chloridis-virgatae (MHYAU 20136), P. chloridis-virgatae (MHYAU 20137), and P. chloridis-virgatae (MHYAU 20058) all have Chloris virgata as a host species, and there are no references about their morphological characteristics. However, they did not cluster with P. virgatae and P. jiaensis in the phylogenetic analysis.
We also searched for Phyllachora species reported from the same host genus in Farr et al. [49]. The results showed that Phyllachora africana (= P. oblongospora), P. graminis, and P. minutissima (=P. bonariensis) can also be parasitic on Chloris species. Phyllachora chloridis, P. graminis, P. africana, P. minutissima, P. virgatae, and P. jiaensis have been reported from the same host genus. However, morphologically they are easily distinguishable: the asci and ascospores of P. virgatae and P. jiaensis were significantly longer than in P. chloridis, P. graminis, and P. minutissima (Table 2). However, the asci of P. chloridis lack pedicels, and the asci of P. graminis have an ascus crown at the apex. Phyllachora minutissima has no paraphyses, while P. virgatae and P. jiaensis have pedicels and paraphyses and lack an ascus crown at the apex. The ascospores of P. virgatae and P. jiaensis have 1–2 or more guttules, but P. chloridis has only one central guttule. The asci of P. africana Parbery and P. minutissima were significantly longer than in P. virgatae and P. jiaensis. Hence, P. virgatae is distinguishable by its different morphological characteristics, which qualify it as a new species of Phyllachora.
Phyllachora jiaensis H. X. Wu & J. C. Li. sp. nov. (Figure 4 and Figure 7).
MycoBank: MB 843398.
Etymology: Epithet derived from the type locality, Jia County (Shaanxi Province, Yulin City) in China.
Holotype: IFRD9448.
Parasitic on leaves and stems of Chloris virgata (Poaceae). Sexual morph: Stroma 825–2321 × 347–640 μm diam. ( x ¯ = 1372 × 501 μm, n = 10) (Figure 7a–c), domed above the leaf surface, amphigenous, fusiform, cymbaeform or of irregular shape, black spots, scattered, sometimes gregarious, without an edge, shiny black, carbonaceous. Section of stroma 143–170 μm high, oval, multilocular, peridium 18–27 μm wide, composed of brown to dark-brown cells of textura angularis (Figure 7d). Paraphyses 3–4 μm wide ( x ¯ = 3.1 μm, n = 20), numerous, persistent, filiform, unbranched, aseptate, slightly longer than asci (Figure 7e). Asci 77–114 × 8–12 μm ( x ¯ = 93.4 × 9.4 μm, n = 20), thin-walled, 8-spored, persistent, clavate, apex obtuse, with pedicels (Figure 7f–i). Ascospores 9–17 × 8–9 μm ( x ¯ =13.4 × 6.5 μm, n = 20), 1-seriate, ovoid to oblong, acute at both ends, hyaline, aseptate, verrucose, with 1–2 or more guttules (Figure 7j–m) and a glutinous mucilaginous sheath (Figure 4m–p). Asexual morph: Not observed.
Material examined: China, Shaanxi Province, Yulin City, Jia County, 38.02329299 N, 110.4956 E, on stems and leaves of Chloris virgata (Poaceae), 6 October 2019, Yuying Li, IFRD9448, holotype. GenBank accession numbers: LSU: ON075440, SSU: ON072100, ITS: ON075527.
Notes: In the course of investigating the grass resources of northern China, Phyllachora jiaensis was collected from Shaanxi Province. According to phylogenetic analysis, P. virgatae and P. jiaensis are closely related; however, they were collected from different locales. Morphologically, the stromata color of P. virgatae is bright black, and in P. jiaensis it is black. In addition, the asci of P. virgatae were longer than in P. jiaensis (102.6 × 9.1 μm vs. 93.4 × 9.4 μm). The ascospores of P. jiaensis were also longer than in P. virgatae (13.4 × 6.5 μm vs. 12.9 × 6.9 μm) (Table 2).
Phylogenetically, P. virgatae and P. jiaensis clustered together with high bootstrap support and probability value (100/1.0), with P. jiaensis forming a long branch. The LSU, SSU, and ITS loci differ by 8 bases, 109 bases, and 3 bases, respectively. Phylogenetically, P. virgatae grouped with P. jiaensis to form one clade, and P. flaccidudis with P. sandiensis to form another clade, with high bootstrap and probability values (100/1.0), but they occur on different hosts. P. virgatae and P. jiaensis both occur on Chloris virgata, and the host of P. flaccidudis and P. sandiensis is Cenchrus flaccidus. The ascospores of P. flaccidudis and P. sandiensis are acute at one end and blunt at the opposite end, while the ascospores of P. virgatae and P. jiaensis are blunt at both ends.
The four new species described herein have ascospores with gelatinous sheaths that differ in black ink (Figure 4). The gelatinous sheaths of P. flaccidudis and P. sandiensis are larger than in P. virgatae and P. jiaensis. Hence, based on both morphological and phylogenetic evidence, we introduce the novel species, P. jiaensis.

4. Discussion

In this study, we introduced four new taxa of Phyllachora (P. flaccidudis, P. sandiensis, P. virgatae, and P. jiaensis) that have morphological characteristics typical of Phyllachora: black leaf spots, peridium clypeate, multilocular, asci cylindrical, an unobvious apical ring, shortly pedicellate, numerous paraphyses and slightly longer than asci, and aseptate ascospores with guttules [4,33,50]. All novel taxa were introduced based on morphological characteristics and novel phylogenetic lineages in Phyllachora (Figure 1). We compared the morphological characteristics of the four new species and similar Phyllachora taxa (Table 2).
In recent years, several Phyllachora species have been introduced in many places of China, such as Phyllachora heterocladae (Sichuan, China), P. panicicola (Yunnan, China), P. eriochloae var. colombiensis (Yunnan, China), P. graminis var. cynodonticola Speg. (Yunnan, China), and P. eriochloae Speg. var. eriochloae (Yunnan, China) [20,33,43,44,45,46,47,48]. However, relevant molecular data only exists for a few of these species (e.g., P. heterocladae and P. panicicola). The majority of these species were reported from Yunnan Province (e.g., P. panicicola). Host specificity plays an important role when introducing novel Phyllachora species [3,4]. Yang et al. [20] proposed that phyllachora-like species that are parasitic on Poaceae should be treated as Phyllachora, and our study also provides strong evidence that supports this hypothesis.
Yang et al. [20] introduced Phyllachora heterocladae from Sichuan Province, and the phylogenetic tree was artificially divided into five lineages based on the host plants. Most species of Phyllachora that cluster within lineage I are graminicolous (Poaceae), but P. qualeae grows on Qualea multiflora (Vochysiaceae). They formed a distinct subclade with P. arundinellae (MHYAU:108), P. cynodontis (MHYAU:20043), and P. imperatae (MHYAU:014). Species within lineage II and lineage IV are bambusicolous fungi. Lineage III is solely composed of P. thysanolaenae (MFLU 16-2071), which is an unstable species in the phylogeny. Lineage V contains only P. pomigena, associated with an unknown host plant. Li et al. [32] introduced two new species, P. dendrocalami-membranacei and P. dendrocalami-hamiltonii, and phylogenetic analysis was generated four main clades. Lineage I consisted of all Phyllachora species obtained from the subfamily Agrostidoideae of the Poaceae, except for Polystigma pusillum (MM-19), which was found growing on Fabaceae. Neophyllachora species occurred in the family Myrtaceae within Lineage II. Lineage III and Lineage IV are Phyllachora species collected from the subfamily Bambusoideae of the Poaceae. However, it is important to note that Yang et al. [20] did not include all Phyllachora species in their analysis.
In this study, the generated phylogenetic tree comprises 74 species belonging to six genera (viz., Ascovaginospora, Camarotella, Coccodiella, Neophyllachora, Phyllachora, and Polystigma). We found that the Phyllachora genus is paraphyletic. Because the host of P. pomigena remains unknown, the species Phyllachora pomigena formed a single clade [20,51]. In the phylogenetic analysis, the new species described herein are included within the Phyllachora genus and separated from other taxa with a single subclade. Their hosts are Cenchrus flaccidus and Chloris virgata, both belonging to Poaceae (graminicolous).
The study revealed that the ancestor of Phyllachora species originated from Latin America. Phyllachora species ancestors initially spread from Latin America to North America, East Asia, South America, and eventually to Central Europe. The characteristic of Phyllachora species in Latin America are consistent with the ancestral characteristics of Phyllachora genus found in Mardones et al. [8]. For example, existing species P. maydis and P. graminis still retain ancestral characteristics, such as growing on monocotyledonous hosts, immersed perithecia, black stromata, and the presence of clypeus [8]. Reconstruction analysis of ancestral location indicates that a vicariance event (i.e., the splitting of the range of a taxon or biota into two or more geographical subdivisions by the formation of natural barriers, for example, mountain building, glaciation, plate tectonics or climate change) affected speciation allowing some species to retain ancestral morphological characteristics [52]. The appearance of Polystigma could have resulted from the extinction event (Figure 2 node 143). The extinction event may have resulted in the host of Polystigma species shifting from monocotyledons to dicotyledons (Fabaceae).
During the Cretaceous geological upheaval, orogeny, continental drift as well as the emergence of the Atlantic and the Indian Ocean led to dramatic terrestrial climate changes across the earth’s surface [53]. These led to the mass extinction of the dominant Mesozoic gymnosperm and ferns in the tropics, subtropical plains, and low mountains areas, which were replaced by angiosperms (the origin of Poaceae) that flourished in the Paleogene [54]. The emergence of angiosperms may have triggered the evolution and migration of the ancestors of the Phyllachora fungi.
There are few studies examining the co-evolution and ancestral state reconstruction of Phyllachora species; this is because of the scarcity of existing species with high-quality molecular data, which adds uncertainty to the process of ancestral state reconstruction. Extensive sampling and high-quality molecular data will reveal more accurate changes in the ancestral status of species in this group. Ancestor state reconstruction currently requires inferring phenotypes of ancestral species using observations from present-day species [55,56]. As new classical and molecular methods for identifying fungi continue to develop [57], ancestor state reconstruction analysis of fungal taxonomy is at the forefront of a new trend [8,58,59,60]. Future studies on species diversity and evolution of Phyllachora species require more extensive sampling and high-quality molecular data.

Author Contributions

Methodology, H.-X.W. and J.-C.L.; formal analysis, J.-C.L., H.-X.W.; resources, Y.L., J.-C.L. and H.-X.W.; data curation, J.-C.L., X.-H.L., J.-Y.S. and H.-X.W., writing—original draft preparation, J.-C.L., H.-X.W., N.N.W.; writing—review and editing, J.-C.L., H.-X.W., N.N.W. and N.S.; supervision, H.-X.W.; project administration, H.-X.W.; funding acquisition, H.-X.W., N.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Grant for Essential Scientific Research of National Nonprofit Institute (no. CAFYBB2019QB005), the Yunnan Province Ten Thousand Plan of Youth Top Talent Project (no. YNWR-QNBJ-2018-267), the National Natural Science Foundation of China (grant No. 32170024) and the Yunnan Fundamental Research Projects (grant No. 202001AT070014). Nalin N. Wijayawardene thanks the High-Level Talent Recruitment Plan of Yunnan Provinces (“Young Talents” Program and “High-End Foreign Experts” Program). Nakarin Suwannarach thanks Chiang Mai University, Thailand for partial support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All sequence data are available in NCBI GenBank following the accession numbers in the manuscript. All species data are available in MycoBank.

Acknowledgments

We acknowledge Austin smith for providing helpful suggestions to improve this paper.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. Barr, M.E. The Ascomycete Connection. Mycologia 1983, 75, 1–13. [Google Scholar] [CrossRef]
  2. Hyde, K.D.; Cannon, P.F.; Barr, M.E. Phaeochoraceae, a new family from palms. Syst. Ascomycetum. 1997, 15, 117–120. [Google Scholar]
  3. Kirk, P.M.; Cannon, P.F.; Minter, D.W.; Stalpers, J.A. Dictionary of Fungi, 10th ed.; CABI: London, UK, 2008; p. 527. [Google Scholar]
  4. Maharachchikumbura, S.; Hyde, K.D.; Jones, E.; Mckenzie, E.; Wijayawardene, N.N. Families of Sordariomycetes. Fungal Divers. 2016, 79, 1–317. [Google Scholar] [CrossRef]
  5. Cannon, P.F. A Revision of Phyllachora and Some Similar Genera on the Host Family Leguminosae; CABI: Wallingford, UK, 1991; pp. 1–302. [Google Scholar]
  6. Pearce, C.A.; Reddell, P.; Hyde, K.D. A revision of Phyllachora (Ascomycotina) on hosts in the angiosperm family Asclepiadaceae, including P. gloriana sp. nov. on Tylophora benthamii from Australia. Fungal Divers. 1999, 3, 123–138. Available online: https://www.fungaldiversity.org/fdp/sfdp/FD_3_123-138.pdf (accessed on 13 May 2022).
  7. Piepenbring, M.; Hofmann, T.A.; Kirschner, R.; Mangelsdorff, R.; Trampe, T. Diversity patterns of Neotropical plant parasitic microfungi. Ecotropica 2011, 17, 27–40. [Google Scholar]
  8. Mardones, M.; Trampe-Jaschik, T.; Oster, S.; Elliott, M.; Piepenbring, M. Phylogeny of the order Phyllachorales (Ascomycota; Sordariomycetes): Among and within order relationships based on five molecular loci. Persoonia 2017, 39, 74–90. [Google Scholar] [CrossRef]
  9. Sutton, B.C.; Hodges, C.S. Hawaiian Forest Fungi. III. A New Species, Gloeocoryneum hawaiiense, on Acacia koa. Mycologia 1983, 75, 280–284. [Google Scholar] [CrossRef]
  10. Calatayud, V.; Navarro-Rosinés, P.; Calvo, E. Lichenochora Mediterraneae (phyllacorales, Ascomycota) a new Lichenicolous Fungus from Spain. Lichenologist 2000, 32, 225–231. [Google Scholar] [CrossRef]
  11. Pearce, C.A.; Reddell, P.; Hyde, K.D. Revision of the Phyllachoraceae (Ascomycota) on hosts in the angiosperm family, proteaceae. Aust. Syst. Bot. 2001, 14, 283–328. [Google Scholar] [CrossRef]
  12. Mardones, M.; Trampe-Jaschik, T.; Piepenbring, M. Phylogenetics and taxonomy of Telimenaceae (Phyllachorales) from Central America. Mycol. Prog. 2020, 19, 1587–1599. [Google Scholar] [CrossRef]
  13. Wijayawardene, N.N.; Hyde, K.D.; Dai, D.Q.; Sánchez-García, M.; Goto, B.T. Outline of Fungi and fungus-like taxa 2021. Mycosphere 2022, 13, 53–453. [Google Scholar] [CrossRef]
  14. Parbery, D.G. Studies on graminicolous species of Phyllachora Nke. in Fckl. V. A taxonomic monograph. Aust. J. Bot. 1967, 15, 271–375. [Google Scholar] [CrossRef]
  15. Hawksworth, D.L.; Kirk, P.M.; Sutton, B.C.; Pegler, D.N. Ainsworth and Bisby’s Dictionary of the Fungi, 8th ed.; CABI: Oxford, UK, 1995; p. 616. [Google Scholar]
  16. Konta, S.; Maharachchikumbura, S.S.N.; Senanayake, I.C.; McKenzie, E.H.C.; Stadler, M.; Boonmee, S.; Phookamsak, R.; Jayawardena, R.S.; Senwanna, C.; Hyde, K.D.; et al. Refined families of Sordariomycetes. Mycosphere 2020, 11, 305–1059. [Google Scholar] [CrossRef]
  17. Theissen, F.; Sydow, H. Die Dothideales. Annales Mycologici 1915, 13, 147–746. [Google Scholar]
  18. Santos, M.D.M.D.; Fonseca-Boiteux, M.E.; Boiteux, L.S.; Câmara, P.E.A.S.; Dianese, J.C. ITS-phylogeny and taxonomy of Phyllachora species on native Myrtacae from the Brazilian Cerrado. Mycologia 2016, 108, 1141–1164. [Google Scholar] [CrossRef]
  19. Clements, F.E.; Shear, C.L. The Genera of Fungi, 2nd ed.; Wilson Co.: New York, NY, USA, 1931. [Google Scholar]
  20. Yang, C.L.; Xu, X.L.; Liu, Y.G.; Hyde, K.D.; Mckenzie, E. A new species of Phyllachora (Phyllachoraceae, Phyllachorales) on Phyllostachys heteroclada from Sichuan, China. Phytotaxa 2019, 392, 186–196. [Google Scholar] [CrossRef]
  21. McCoy, A.G.; Romberg, M.K.; Zaworsk, E.R.; Robertson, A.E.; Phibbs, A.; Hudelson, B.D.; Smith, D.L.; Beiriger, R.L.; Raid, R.N.; Byrne, J.M.; et al. First Report of Tar Spot on Corn (Zea mays) Caused by Phyllachora maydis in Florida, Iowa, Michigan, and Wisconsin. Plant Dis. 2018, 102, 6. [Google Scholar] [CrossRef]
  22. Lana, F.D.; Plewa, D.E.; Phillippi, E.S.; Garzonio, D.; Hesterman, R.; Kleczewski, N.M.; Paul, P.A. First report of tar spot of maize (Zea mays), caused by Phyllachora maydis, in Ohio. Plant Dis. 2019, 103, 1780. [Google Scholar] [CrossRef]
  23. Mottaleb, K.A.; Loladze, A.; Sonder, K.; Kruseman, G.; Vicente, F.S. Threats of Tar Spot Complex disease of maize in the United States of America and its global consequences. Mitig Adapt Strateg Glob Chang 2019, 24, 281–300. [Google Scholar] [CrossRef] [Green Version]
  24. Telenko, D.E.P.; Ross, T.J.; Shim, S.; Wang, Q.; Singh, R. Draft Genome Sequence Resource for Phyllachora maydis—An Obligate Pathogen That Causes Tar Spot of Corn with Recent Economic Impacts in the United States. Mol. Plant Microbe Interact. 2020, 33, 884–887. [Google Scholar] [CrossRef] [Green Version]
  25. Cannon, P.F. Proposal to merge the Phyllachorales with the Diaporthales, with a new family structure. Syst. Ascomycetum 1988, 7, 23–43. [Google Scholar]
  26. Zhang, Z.Y.; Zhang, T. Flora Fungorum Sinicorum, Vol. 46, Phyllachora; Science Press: Beijing, China, 2014; p. 530. (In Chinese) [Google Scholar]
  27. Tamakaew, N.; Maharachchikumbura, S.; Hyde, K.D.; Cheewangkoon, R. Tar spot fungi from Thailand. Mycosphere 2017, 8, 1054–1058. [Google Scholar] [CrossRef]
  28. Wu, H.X.; Li, Y.M.; Ariyawansa, H.A.; Li, W.J.; Hyde, K.D. A new species of Microthyrium from Yunnan, China. Phytotaxa 2014, 176, 213–218. [Google Scholar] [CrossRef] [Green Version]
  29. Vilgalys, R.; Hester, M. Rapid genetic identification and mapping of enzymatically amplified ribosomal DNA from several Cryptococcus species. J. Bacteriol. 1990, 172, 4238–4246. [Google Scholar] [CrossRef] [Green Version]
  30. White, T.J.; Bruns, S.; Lee, S.; Taylor, J. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In PCR Protocols; Michael, A., David, H., Eds.; Academic Press: London, UK, 1990; pp. 315–322. [Google Scholar]
  31. Hall, T.A. BioEdit: A user-friendly biological sequence alignment program for Windows 95/98/NT. Nucl. Acids Symp. Ser. 1999, 41, 95–98. [Google Scholar] [CrossRef]
  32. Li, X.; Wu, S.; Wang, C.; Feng, Y.; Zhao, C.; Chen, Z.-Q.; Yu, J.-F.; Luo, R.; Promputtha, I.; Sun, D.-F. Two new species of Phyllachora (Phyllachoraceae; Phyllachorales) on bamboo from China. Phytotaxa 2019, 425, 78–86. [Google Scholar] [CrossRef]
  33. Dayarathne, M.; Maharachchikumbura, S.; Jones, E.B.G.; Goonasekara, I.D.; Bulgakov, T.S.; Al-Sadi, A.M.; Lumyong, S.; McKenzie, E.H.C. Neophyllachora gen. nov. (Phyllachorales), three new species of Phyllachora from Poaceae and resurrection of Polystigmataceae (Xylariales). Mycosphere 2017, 8, 1598–1625. [Google Scholar] [CrossRef]
  34. Katoh, K.; Rozewicki, J.; Yamada, K.D. MAFFT online service: Multiple sequence alignment, interactive sequence choice and visualization. Brief. Bioinformatics 2019, 20, 1160–1166. [Google Scholar] [CrossRef] [Green Version]
  35. Edler, D.; Klein, J.; Antonelli, A.; Silvestro, D. raxmlGUI 2.0: A graphical interface and toolkit for phylogenetic analyses using RAxML. Methods Ecol. Evol. 2021, 12, 373–377. [Google Scholar] [CrossRef]
  36. Nylander, J.A.A. MrModeltest v2. Program distributed by the author. Bioinformatics 2004, 24, 533–581. [Google Scholar] [CrossRef] [Green Version]
  37. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Cai, L.; Jeewon, R.; Hyde, K.D. Phylogenetic investigations of Sordariaceae based on multiple gene sequences and morphology. Mycol. Res. 2006, 110, 137–150. [Google Scholar] [CrossRef] [PubMed]
  39. Cai, L.; Guo, X.Y.; Hyde, K.D. Morphological and molecular characterization of a new anamorphic genus Cheirosporium; from freshwater in China. Persoonia 2008, 20, 53–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Liu, J.K.; Phookamsak, R.; Doilom, M.; Wikee, S.; Li, Y.M.; Ariyawansha, H.; Boonmee, S.; Chomnunti, P.; Dai, D.-Q.; Bhat, J.D.; et al. Towards a natural classification of Botryosphaeriales. Fungal Divers. 2012, 57, 149–210. [Google Scholar] [CrossRef]
  41. Wu, H.X.; Schoch, C.L.; Boonmee, S.; Bahkali, A.H.; Chomnunti, P.; Hyde, K.D. A reappraisal of Microthyriaceae. Fungal Divers. 2011, 51, 189–248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Yu, Y.; Harris, A.J.; Blair, C.; He, X.J. RASP (Reconstruct Ancestral State in Phylogenies): A tool for historical biogeography. Mol. Phylogenet. Evol. 2015, 87, 46–49. [Google Scholar] [CrossRef]
  43. Hongsanan, S.; Maharachchikumbura, S.S.; Hyde, K.D.; Samarakoon, M.C.; Jeewon, R.; Zhao, Q.; Al-Sadi, A.; Bahkali, A.H. An updated phylogeny of Sordariomycetes based on phylogenetic and molecular clock evidence. Fungal Divers. 2017, 84, 25–41. [Google Scholar] [CrossRef]
  44. Liu, N.; Guo, L. Two new records of Phyllachora (Phyllachorales) from China. Mycotaxon 2009, 107, 303–306. [Google Scholar] [CrossRef]
  45. Lan, J.Q.; Lu, H.J.; Wu, J. Three new records of Phyllachora in China. Mycosystema 2012, 31, 639–641. [Google Scholar]
  46. Parbery, D.G. Studies on Graminicolous species, of Phyllachora Nke. in Fckl. VI. Additions and corrections to part V. Aust. J. Bot. 1971, 19, 207–235. [Google Scholar] [CrossRef]
  47. Wu, J. Additional three new Chinese records of Phyllachora. Mycosystema 2012, 31, 642–644. [Google Scholar]
  48. Teng, X.Q. Additional three new records of Phyllachora in China. Mycosystema 2011, 30, 782–784. [Google Scholar]
  49. Farr, D.F.; Rossman, A.Y. Fungal Databases, U.S. National Fungus Collections, ARS, USDA. Available online: https://nt.ars-grin.gov/fungaldatabases/ (accessed on 12 April 2022).
  50. Silva-Hanlin, D.M.W.; Halin, R.T. The order Phyllachorales: Taxonomic review. Mycoscience 1998, 39, 97–104. [Google Scholar] [CrossRef]
  51. Vu, D.; Groenewald, M.; de Vries, M.; Gehrmann, T.; Stielow, B.; Eberhardt, U.; Al-Hatmi, A.; Groenewald, J.Z.; Cardinali, G.; Houbraken, J.; et al. Large-scale generation and analysis of filamentous fungal DNA barcodes boosts coverage for kingdom fungi and reveals thresholds for fungal species and higher taxon delimitation. Stud. Mycol. 2019, 92, 135–154. [Google Scholar] [CrossRef] [PubMed]
  52. Humphries, C.J.; Escudero, M.; Martín-Bravo, S. Vicariance Biogeography. In Reference Module in Life Sciences, 3rd ed.; Roitberg, B., Ed.; Elsevier: Cambridge, UK, 2017; pp. 1–13. [Google Scholar]
  53. Feng, J.G.; Fan, F.C.; Li, F. Origin, Evolution and Distribution of the Gramineae. Chinese J. Bot. 1996, 13, 9–13. (In Chinese) [Google Scholar]
  54. Tzvelev, N.N. The system of grasses (Poaceae) and their evolution. Bot. Rev. 1989, 55, 141–203. [Google Scholar] [CrossRef]
  55. Schneider, J.V.; Bissiengou, P.; Amaral, M.; Tahir, A.; Fay, M.F.; Thines, M.; Sosef, M.S.; Zizka, G.; Chatrou, L.W. Phylogenetics, ancestral state reconstruction, and a new infrafamilial classification of the pantropical Ochnaceae (Medusagynaceae, Ochnaceae s.str. Quiinaceae) based on five dna regions. Mol. Phylogenet. Evol. 2014, 78, 199–214. [Google Scholar] [CrossRef]
  56. Ho, L.; Dinh, V.; Nguyen, C.V. Multi-task learning improves ancestral state reconstruction. Theor. Popul. Biol. 2019, 126, 33–39. [Google Scholar] [CrossRef]
  57. Gautam, A.K.; Verma, R.K.; Avasthi, S.; Sushma; Bohra, Y.; Devadatha, B.; Niranjan, M.; Suwannarach, N. Current Insight into Traditional and Modern Methods in Fungal Diversity Estimates. J. Fungi 2022, 8, 226. [Google Scholar] [CrossRef]
  58. Li, J.; Han, L.H.; Liu, X.B.; Zhao, Z.W.; Yang, Z.L. The saprotrophic Pleurotus ostreatus species complex: Late Eocene origin in East Asia; multiple dispersal, and complex speciation. IMA Fungus 2020, 11, e10. [Google Scholar] [CrossRef]
  59. Quan, Y.; Muggia, L.; Moreno, L.F.; Wang, M.; Hoog, S.D. A re-evaluation of the Chaetothyriales using criteria of comparative biology. Fungal Divers. 2020, 103, 47–85. [Google Scholar] [CrossRef]
  60. Thiyagaraja, V.; Lücking, R.; Ertz, D.; Karunarathna, S.C.; Wanasinghe, D.N.; Lumyong, S.; Hyde, K. The Evolution of Life Modes in Stictidaceae, with Three Novel Taxa. J. Fungi 2021, 7, 105. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Phylogenetic tree of maximum likelihood showing the relationships of Phyllachoraceae based on combined LSU, SSU, and ITS data set analysis. Bootstrap values of maximum likelihood higher than 50% are shown on the left, while values of Bayesian posterior probabilities above 0.5 are shown on the right. New species are given in bold, followed by the host of the species behind its strain number.
Figure 1. Phylogenetic tree of maximum likelihood showing the relationships of Phyllachoraceae based on combined LSU, SSU, and ITS data set analysis. Bootstrap values of maximum likelihood higher than 50% are shown on the left, while values of Bayesian posterior probabilities above 0.5 are shown on the right. New species are given in bold, followed by the host of the species behind its strain number.
Jof 08 00520 g001
Figure 2. Ancestral character state reconstruction based on the Bayesian tree. Each event is represented with a number at the nodes. Bayesian posterior probabilities are presented (≥0.5). The colored circle near the number at the nodes indicate that blue represents Dispersal, green represents Vicariance, orange represents Extinction. New species are given in bold.
Figure 2. Ancestral character state reconstruction based on the Bayesian tree. Each event is represented with a number at the nodes. Bayesian posterior probabilities are presented (≥0.5). The colored circle near the number at the nodes indicate that blue represents Dispersal, green represents Vicariance, orange represents Extinction. New species are given in bold.
Jof 08 00520 g002
Figure 3. Phyllachora flaccidudis (IFRD9445, holotype). (a) Black spots on Cenchrus flaccidus (Poaceae); (b,c) Stromata; (d) Vertical section of ascomata in cotton blue; (e) Paraphyses; (fi) Asci; (jm) Ascospores. Scale bars, (b) 1 mm; (c)0.5 mm; (d) 200 μm; (e) 50 μm; (fi) 10 μm; (jm) 5 μm. Microscopic techniques: DIC.
Figure 3. Phyllachora flaccidudis (IFRD9445, holotype). (a) Black spots on Cenchrus flaccidus (Poaceae); (b,c) Stromata; (d) Vertical section of ascomata in cotton blue; (e) Paraphyses; (fi) Asci; (jm) Ascospores. Scale bars, (b) 1 mm; (c)0.5 mm; (d) 200 μm; (e) 50 μm; (fi) 10 μm; (jm) 5 μm. Microscopic techniques: DIC.
Jof 08 00520 g003
Figure 4. (a) P. flaccidudis ascospore in DIW (Deionized Water); (b,c) P. flaccidudis ascospores with gelatinous sheath in ink; (e) P. sandiensis ascospore in DIW; (fh) P. sandiensis ascospores with gelatinous sheath in ink; (i) P. virgatae ascospore in DIW; (j,k) P. virgatae ascospores with gelatinous sheath in ink; (m) P. jiaensis ascospore in DIW; (np) P. jiaensis with gelatinous sheath in ink. Scale bars, (ap) 5 μm. Microscopic techniques: DIC.
Figure 4. (a) P. flaccidudis ascospore in DIW (Deionized Water); (b,c) P. flaccidudis ascospores with gelatinous sheath in ink; (e) P. sandiensis ascospore in DIW; (fh) P. sandiensis ascospores with gelatinous sheath in ink; (i) P. virgatae ascospore in DIW; (j,k) P. virgatae ascospores with gelatinous sheath in ink; (m) P. jiaensis ascospore in DIW; (np) P. jiaensis with gelatinous sheath in ink. Scale bars, (ap) 5 μm. Microscopic techniques: DIC.
Jof 08 00520 g004
Figure 5. Phyllachora sandiensis (IFRD9446, holotype). (a) Black spots on Cenchrus flaccidus (Poaceae); (b,c) Stromata; (d) Vertical section of ascomata; (e) Paraphyses; (f) Ascus in cotton blue; (gi) Asci; (j) Ascospore in cotton blue; (km) Ascospores. Scale bars, (c) 0.5 mm; (d) 100 μm; (e) 50 μm; (fi) 10 μm; (jm) 5 μm. Microscopic techniques: DIC.
Figure 5. Phyllachora sandiensis (IFRD9446, holotype). (a) Black spots on Cenchrus flaccidus (Poaceae); (b,c) Stromata; (d) Vertical section of ascomata; (e) Paraphyses; (f) Ascus in cotton blue; (gi) Asci; (j) Ascospore in cotton blue; (km) Ascospores. Scale bars, (c) 0.5 mm; (d) 100 μm; (e) 50 μm; (fi) 10 μm; (jm) 5 μm. Microscopic techniques: DIC.
Jof 08 00520 g005
Figure 6. Phyllachora virgatae (IFRD9447, holotype). (a) Black spots on Chloris virgata (Poaceae); (b,c) Stromata; (d) Vertical section of ascomata; (e) Paraphyses; (f) Ascus in cotton blue; (gi) Asci; (jm) Ascospores. Scale bars, (b) 1 mm, (c) 0.5 mm; (d) 100 μm; (e) 20 μm; (fi) 10 μm; (jm) 5 μm. Microscopic techniques: DIC.
Figure 6. Phyllachora virgatae (IFRD9447, holotype). (a) Black spots on Chloris virgata (Poaceae); (b,c) Stromata; (d) Vertical section of ascomata; (e) Paraphyses; (f) Ascus in cotton blue; (gi) Asci; (jm) Ascospores. Scale bars, (b) 1 mm, (c) 0.5 mm; (d) 100 μm; (e) 20 μm; (fi) 10 μm; (jm) 5 μm. Microscopic techniques: DIC.
Jof 08 00520 g006
Figure 7. Phyllachora jiaensis (IFRD9448, holotype). (a) Black spots on Chloris virgata (Poaceae); (b,c) Stromata; (d) Vertical section of ascomata; (e) Paraphyses; (f) Ascus in cotton blue; (gi) Asci; (jm) Ascospores. Scale bars, (c) 0.5 mm; (d) 100 μm; (e) 20 μm; (fi) 10 μm; (jm) 5 μm. Microscopic techniques: DIC.
Figure 7. Phyllachora jiaensis (IFRD9448, holotype). (a) Black spots on Chloris virgata (Poaceae); (b,c) Stromata; (d) Vertical section of ascomata; (e) Paraphyses; (f) Ascus in cotton blue; (gi) Asci; (jm) Ascospores. Scale bars, (c) 0.5 mm; (d) 100 μm; (e) 20 μm; (fi) 10 μm; (jm) 5 μm. Microscopic techniques: DIC.
Jof 08 00520 g007
Table 1. List of source and GenBank accession numbers used in this study. Sequences generated in this study are written in blue. Type collections are marked with ‘T’.
Table 1. List of source and GenBank accession numbers used in this study. Sequences generated in this study are written in blue. Type collections are marked with ‘T’.
SpeciesLocationSourceHost FamilyGenBank Accession NumbersReference
LSUSSUITS
Ascovaginospora stellipalaTNorth America (northern Wisconsin)P5-13ACyperaceaeU85088U85087-[33]
Camarotella costaricensisLatin America (Panama)MM-21ArecaceaeKX430490KX451851KX451900[8]
Camarotella costaricensisLatin America (Panama)MM-149ArecaceaeKX430484KX451863KX451913[8]
Camarotella sp.Latin America (Panama)MM-27ArecaceaeKX430492KX451852KX451901[8]
Coccodiella calatheaeTLatin America (Panama)MP5133MarantaceaeMF460370MF460376MF460366[8]
Coccodiella melastomatumNorth America (Venezuela)CMU78543Melastomataceae-U78543-[8]
Coccodiella miconiaeLatin America (Panama)ppMP1342MelastomataceaeKX430506KX451871MF460365[8]
Coccodiella miconiicolaLatin America (Panama)TH-571MelastomataceaeKX430512KX451880-[8]
Coccodiella miconiicolaLatin America (Panama)CBMAP-H290AMelastomataceaeMF460373MF460379MF460368[32]
Coccodiella miconiicolaLatin America (Ecuador)SO-15MelastomataceaeMF460374MF460380MF460369[32]
Coccodiella toledoiLatin America (Ecuador)MM-165MelastomataceaeKX430488KX451865KX451917[8]
Neophyllachora cerradensisLatin America (Brazil)UB21823Myrtaceae--KC683470[18]
Neophyllachora cerradensisTLatin America (Brazil)UB21908Myrtaceae--KC683471[18]
Neophyllachora myrciaeLatin America (Brazil)UB21292Myrtaceae--KC683463[18]
Neophyllachora myrciaeLatin America (Brazil)UB22192Myrtaceae--KC683476[18]
Neophyllachora myrciariaeTLatin America (Brazil)UB21781Myrtaceae--KC683469[18]
Neophyllachora subcircinansLatin America (Brazil)UB09748Myrtaceae--KC683441 [18]
Neophyllachora subcircinansLatin America (Brazil),
South America (Paraguay)
UB21347Myrtaceae--KC683466[18]
Neophyllachora subcircinansLatin America (Brazil),
South America (Paraguay)
UB21747Myrtaceae-KC902622KC683467[18]
Neophyllachora truncatisporaLatin America (Brazil)UB14083Myrtaceae-KC902614KC683448[18]
Phyllachora arthraxonisEast Asia (China)MHYAU:072PoaceaeMG269803-MG269749[20]
Phyllachora arundinellaeEast Asia (China)MHYAU:108PoaceaeMG269815-MG269761[20]
Phyllachora capillipediicolaEast Asia (China)MHYAU 20089PoaceaeMG356698-KY498084[32]
Phyllachora capillipediicolaEast Asia (China)MHYAU:20090PoaceaeMG356699-KY498115[20]
Phyllachora chloridisTSoutheast Asia (Thailand)MFLU 15-0173PoaceaeMF197499MF197505KY594026[33]
Phyllachora chloridisSoutheast Asia (Thailand)MFLU 16-2980PoaceaeMF197500MF197506KY594027[33]
Phyllachora chloridis-virgataeEast Asia (China)MHYAU 20136PoaceaeMG356685-KY498122[32]
Phyllachora chloridis-virgataeEast Asia (China)MHYAU:20058PoaceaeMG356683-KY498102[32]
Phyllachora chloridis-virgataeEast Asia (China)MHYAU 20137PoaceaeMG356686-KY498092[32]
Phyllachora chrysopogonicolaTSoutheast Asia (Thailand)MFLU 16-2096PoaceaeMF372146-MF372145[20]
Phyllachora cynodonticolaTSoutheast Asia (Thailand)MFLU 16-2977PoaceaeMF197501MF197507KY594024[33]
Phyllachora cynodonticolaSoutheast Asia (Thailand)MFLU 16-2978PoaceaeMF197502MF197508KY594025[33]
Phyllachora cynodontisEast Asia (China)MHYAU 20042PoaceaeKY498080-KY471328[32]
Phyllachora cynodontisEast Asia (China)MHYAU:20043PoaceaeKY498081-KY471329[20]
Phyllachora cynodontisEast Asia (China)MHYAU 20131PoaceaeKY498079-KY471327[32]
Phyllachora dendrocalami-hamiltoniicolaEast Asia (China)MHYAU 221PoaceaeMK614118--[32]
Phyllachora dendrocalami-membranaceiEast Asia (China)MHYAU 220PoaceaeMK614117-MK614102[32]
Phyllachora dendrocalami-membranaceiEast Asia (China)MHYAU 222PoaceaeMK614119-MK614103[32]
Phyllachora flaccidudisTEast Asia (China)IFRD9445PoaceaeON072101ON072097ON075524This study
Phyllachora graminisNorth America (Canada)DAOM 240981Poaceae--HQ317550[33]
Phyllachora graminisCentral Europe (Germany)MM-166Poaceae-KX451869KX451920[8]
Phyllachora heterocladaeTEast Asia (China)MFLU 18-1221PoaceaeMK296472MK296468MK305902[20]
Phyllachora imperataeEast Asia (China)MHYAU:014PoaceaeMG269800-MG269746[20]
Phyllachora indosasaeEast Asia (China)MHYAU 125PoaceaeMG195662-MG195637[32]
Phyllachora isachnicolaTEast Asia (China)MHYAU:179PoaceaeMH018563-MH018561[20]
Phyllachora isachnicolaEast Asia (China)MHYAU:180PoaceaeMH018564-MH018562[20]
Phyllachora jiaensisTEast Asia (China)IFRD9448PoaceaeON075440ON072100ON075527This study
Phyllachora keralensisEast Asia (China)MHYAU:20082PoaceaeMG269792-KY498106[20]
Phyllachora maydisNorth America (USA) BPI 893231Poaceae--KU184459[33]
Phyllachora maydisNorth America (Wisconsin)BPI 910560Poaceae--MG881846[20]
Phyllachora miscanthiEast Asia (China)MHYAU:167PoaceaeMG195669-MG195644[20]
Phyllachora miscanthiEast Asia (China)MHYAU:157PoaceaeMG195668-MG195643[20]
Phyllachora panicicolaTEast Asia (China)MFLU 16-2979PoaceaeMF197503MF197504KY594028[33]
Phyllachora pogonatheriEast Asia (China)MHYAU:071PoaceaeMG269802-MG269748[20]
Phyllachora pogonatheriEast Asia (China)MHYAU:070PoaceaeMG269801-MG269747[20]
Phyllachora pomigenaunknownCBS 194.33UnknownMH866861-MH855410[20]
Phyllachora pomigenaunknownCBS 193.33UnknownMH866860-MH855409[20]
Phyllachora qualeaeunknownUB 21159Vochysiaceae--KU682781[33]
Phyllachora qualeaeunknownUB 21771Vochysiaceae--KU682780[33]
Phyllachora sandiensisTEast Asia (China)IFRD9446PoaceaeON075528ON072098ON075525This study
Phyllachora sinobambusaeEast Asia (China)MHYAU 085 PoaceaeMG195655-MG195630 [32]
Phyllachora sphaerocaryiTEast Asia (China)MHYAU 178PoaceaeMK614114-MK614100[32]
Phyllachora sphaerocaryiEast Asia (China)MHYAU:178Poaceae--MH018560[20]
Phyllachora thysanolaenaeTSoutheast Asia (Thailand)MFLU 16-2071Poaceae -MF372147-[20]
Phyllachora virgataesTEast Asia (China)IFRD9447PoaceaeON075439ON072099ON075526This study
Phyllachora yushaniae-falcatiauritaeEast Asia (China)MHYAU 123PoaceaeMG195656-MG195631 [32]
Phyllachora yushaniae-polytrichaeEast Asia (China)MHYAU 122 PoaceaeMG195657MH992455MG195632 [32]
Phyllachora yushaniae-polytrichaeEast Asia (China)MHYAU 158PoaceaeMG195658-MG195633 [32]
Polystigma pusillumLatin America (Costa Rica)MM-113FabaceaeKX430474KX451858KX451907[8]
Polystigma pusillumLatin America (Costa Rica)MM-147FabaceaeKX430483KX451862-[8]
Polystigma pusillumLatin America (Panama)MM-19FabaceaeKX430489KX451850KX451899[8]
Polystigma sp. Latin America (Ecuador)MM-163PoaceaeKX430487KX451864KX451916[8]
Telimena bicinctaLatin America (Costa Rica)MM-108PicramniaceaeKX430473KX451857KX451906[8]
Telimena bicinctaLatin America (Costa Rica)MM-133PicramniaceaeKX430478KX451861KX451910[8]
Table 2. Morphological comparison of four new species (in bold) and related species in Phyllachora reported from Poaceae.
Table 2. Morphological comparison of four new species (in bold) and related species in Phyllachora reported from Poaceae.
Fungal TaxaHostsColor of the StromataAsci (μm)Ascospores (μm)References
SizeNo. of
Septa
Shape
P. africana
(P. oblongospora)
Eremopogon delavayi,
Leea elata,
Chloris sp.
Black100–140 × 9–12.510–17 × 5–9Aseptateovoid[44]
P. sphaerosperma
(P. cenchricola)
Cenchrus echinatusBlack65–100 × 10–13 8–11 × 7–9Aseptatenearly spherical[14]
P. centothecaeCentotheca lappaceaBright46.3 × 9.0, pedicel
11.6 × 2.6
7.7–9.0 × 4.6–5.1Aseptateoval[45]
P. coorgianaCoix lachryma-jobiBright59–100 ×18–2610.3–20.6 × 7.2–12.9Aseptateellipsoid or ovoid, rarely subglobose[45]
P. chloridisChloris sp.Bright50–72 × 6–88–12 × 3.5–4.8Aseptatefusiform to oval[33]
P. cynodontisChloris sp.Unknown45–50 × 12–15
with a stipe 20–25 long, sometimes short
8–15 × 5–60–1ovoid[14,46]
P. digitariicolaDigitaria sanguinalisBright62–103 × 12–1510–18 × 6–8Aseptateellipsoid, rounded at both ends, rarely subglobose or oval[45]
P. eriochloae var.
colombiensis
GramineaeUnknown37–71 × 102–134–13.9 × 5.4–7.2Aseptateovate, rarely subglobose, rarely irregularly[47]
P. eriochloae var. eriochloaeEragrostis sp.Bright43.7–69.4 × 10.3–12.97.7–12.9 × 5.1–5.1Aseptateovoid or tear-like[47]
P. flaccidudisCenchrus flaccidusBlack82–110 × 7–1011–13 × 4–7Aseptatedrop shape, oval to ellipse, rounded at the endsThis study
P. graminisChloris sp., Elymus sp., Agvopyron sp., Arrenathevum sp., Asperella sp., Agrostis sp., Brachyelytvum sp., Bvomus sp., Cinna sp.Dark brown to black60–70× 8–107–14× 4–7Aseptateoval to ovoid or ovoid with obtuse end flattened or blunted[14]
P. graminis var.
cynodonticola
Cynodon dactylonBright82–87 × 7.7–8.1, with short peduncle 26 × 2.17.5–14 × 5.1–6.5Aseptateusually oblique, rarely irregularly biseriate,
ellipsoid or subglobose
[47]
P. jiaensisChloris virgataBlack77–114 × 8–129–17 × 8–9Aseptateoval to ellipse, rounded at the endsThis study
P. koondrookensisChloris truncataeBlack75-87 × 12-1614–16.5 × 5–5.5Aseptateuniseriatae vel inordinatae, anguste ellipsoideae usque and
oblongae
[14]
P. minutissimaPennisetum flaccidum, Chloris sp.,
Panicum sp.,
Pennisetum sp., Pseudoechinochlaena sp., Setaria sp.
Black51.2–54.0 × 11.6–14.915.7–19.1 × 6.3–8.2Aseptateovoid or ovate acuminately[14,48]
P. platyellipticaThemeda giguntiaBright64.1–92.1 × 10.1–12.113.6–16.5 × 3.8–6.5Aseptatenarrow-ellipsoid[48]
P. panicicolaPanicum sp.Bright110–130 × 10–1414–16 × 6–8Aseptateellipsoidal, rounded at the ends[33]
P. sandiensisCenchrus flaccidusBlack92–126 × 7–1010–14 × 6–7Aseptatedrop shape, oval to
ellipse
This study
P. virgataeChloris virgataBright84–120 × 7–1110–15 × 6–9Aseptateoval to ellipse, rounded at the endsThis study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, J.-C.; Wu, H.-X.; Li, Y.; Li, X.-H.; Song, J.-Y.; Suwannarach, N.; Wijayawardene, N.N. Taxonomy, Phylogenetic and Ancestral Area Reconstruction in Phyllachora, with Four Novel Species from Northwestern China. J. Fungi 2022, 8, 520. https://0-doi-org.brum.beds.ac.uk/10.3390/jof8050520

AMA Style

Li J-C, Wu H-X, Li Y, Li X-H, Song J-Y, Suwannarach N, Wijayawardene NN. Taxonomy, Phylogenetic and Ancestral Area Reconstruction in Phyllachora, with Four Novel Species from Northwestern China. Journal of Fungi. 2022; 8(5):520. https://0-doi-org.brum.beds.ac.uk/10.3390/jof8050520

Chicago/Turabian Style

Li, Jin-Chen, Hai-Xia Wu, Yuying Li, Xin-Hao Li, Jia-Yu Song, Nakarin Suwannarach, and Nalin N. Wijayawardene. 2022. "Taxonomy, Phylogenetic and Ancestral Area Reconstruction in Phyllachora, with Four Novel Species from Northwestern China" Journal of Fungi 8, no. 5: 520. https://0-doi-org.brum.beds.ac.uk/10.3390/jof8050520

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop