Next Article in Journal
The Therapeutic Monoclonal Antibody Bamlanivimab Does Not Enhance SARS-CoV-2 Infection by FcR-Mediated Mechanisms
Next Article in Special Issue
First Report of Colletotrichum fructicola, C. rhizophorae sp. nov. and C. thailandica sp. nov. on Mangrove in Thailand
Previous Article in Journal
Mosquitoes, Lymphatic Filariasis, and Public Health: A Systematic Review of Anopheles and Aedes Surveillance Strategies
Previous Article in Special Issue
Comprehensive Review of Fungi on Coffee
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Appressoria-Producing Sordariomycetes Taxa Associated with Jasminum Species

by
Deecksha Gomdola
1,2,
Eric H. C. McKenzie
3,
Kevin D. Hyde
1,2,4,
Digvijayini Bundhun
1,2 and
Ruvishika S. Jayawardena
1,2,5,*
1
School of Science, Mae Fah Luang University, Chiang Rai 57100, Thailand
2
Center of Excellence in Fungal Research, Mae Fah Luang University, Chiang Rai 57100, Thailand
3
Manaaki Whenua-Landcare Research, Auckland 1072, New Zealand
4
Innovative Institute for Plant Health, Zhongkai University of Agriculture and Engineering, Guangzhou 510225, China
5
Kyung Hee University, Seoul 02447, Republic of Korea
*
Author to whom correspondence should be addressed.
Pathogens 2023, 12(12), 1407; https://doi.org/10.3390/pathogens12121407
Submission received: 3 October 2023 / Revised: 22 November 2023 / Accepted: 23 November 2023 / Published: 29 November 2023
(This article belongs to the Special Issue Filamentous Fungal Pathogens: 2nd Edition)

Abstract

:
Appressoria are specialized structures formed by certain phytopathogenic fungi during the early stages of the infection process. Over the years, significant advancements have been made in understanding the formation, types, and functions of appressoria. Besides being formed primarily by fungal pathogens, many studies have reported their occurrence in other life modes such as endophytes, epiphytes, and saprobes. In this study, we observed the formation of appressoria in fungal genera that have been found associated with leaf spots and, interestingly, by a saprobic species. We used morphological descriptions and illustrations, molecular phylogeny, coalescent-based Poisson tree processes (PTP) model, inter- and intra-species genetic distances based on their respective DNA markers, and Genealogical Concordance Phylogenetic Species Recognition Analysis (GCPSR) to establish a new species (Pseudoplagiostoma jasmini), a Ciliochorella sp., and a new host record (Coniella malaysiana). The Ciliochorella sp. is reported as a saprobe, while Pseudoplagiostoma jasmini and Coniella malaysiana were found to be associated with leaf spots of Jasminum species. All three taxa produce appressoria, and this is the first study that reports the formation of appressoria by a Ciliochorella sp. and a Pseudoplagiostoma sp.

Graphical Abstract

1. Introduction

Appressoria are infection pegs, mostly produced by pathogenic fungi [1]. However, since these structures are also produced by endophytes, epiphytes, and saprobes, Chethana et al. [1] proposed a general definition of appressoria as “specialized cells or adhesion structures produced by fungi from which a penetration peg emerges that pierces or enters the host tissues”. Frank [2] discovered appressoria and came up with this term when he isolated the pathogen, Colletotrichum lindemuthanium, that causes diseases of beans. Based on the various shapes and sizes, appressoria can be grouped either as single-celled or multi-cellular/compound structures [3]. Single-celled appressoria are sub-divided into proto-appressoria, hyaline, and melanized appressoria. Compound appressoria are further classified as expressoria, infection cushion, and infection plaques [1,3].
Overall, in pathogenesis, appressoria are important for the successful invasion of host plants by certain pathogenic fungi. By attaching to the host, generating turgor pressure, and facilitating penetration, these structures ensure that the pathogen can overcome physical barriers and initiate infection of the plant [1,4]. The most frequently observed appressoria among several fungal species are single-celled, occurring mostly at the tip of germ tubes, sometimes formed laterally or intercalary on hyphae [1,3]. In this study, we identified three taxa isolated from Jasminum spp. that produce appressoria.
Jasminum (Oleaceae), native to tropical and warm temperate regions in Asia, Africa, and Europe, comprises around 200 species [5], several of which are also ecologically and economically important worldwide [6]. They are cultivated as ornamental plants, but they also have traditional and horticultural significance [7,8]. The leaves, stems, bark, roots, and flowers possess beneficial properties, including aphrodisiac, antiseptic, and diuretic [9]. Leaves of J. grandiflorum are used to cure odontalgia, otorrhea, otalgia, dysmenorrhea, leprosy, ulcerative stomatitis, ulcers, and wounds, among other disorders [10,11].
Several studies have reported fungi from Jasminum species. These studies include Colletotrichum jasminigenum and C. siamense on living leaves and flowers of J. sambac in Vietnam [12]; Curvularia prasadii isolated from leaves of J. sambac [13]; Dothidea kunmingensis reported from J. nudiflorum in southwestern China [14]; and Puccinia aizazii, a rust fungus, reported on J. humile from the foothills of the Himalayan ranges, Pakistan [15].
In this study, we employed a polyphasic approach to identify the three species, which resulted in one new taxon (Pseudoplagiostoma jasmini), a Ciliochorella sp., and a new host record (Coniella malaysiana). We used morphological descriptions and illustrations, molecular phylogeny, a coalescent-based Poisson tree processes (PTP) model, inter- and intra-species genetic distances based on their respective DNA markers, and Genealogical Concordance Phylogenetic Species Recognition Analysis (GCPSR). Updated phylogenetic trees comprising all species with molecular data are provided for the three genera. We also present drawings to show the variation in conidial shapes of Pseudoplagiostoma species. All three taxa belong to Sordariomycetes and were isolated from Jasminum spp. in northern Thailand. The Ciliochorella sp. is reported from dead leaves as a saprobe, while Pseudoplagiostoma jasmini and Coniella malaysiana were found to be associated with leaf spots. Single-celled hyaline appressoria were observed in these taxa. This is the first study that reports the formation of appressoria in Ciliochorella spp. and Pseudoplagiostoma spp.

2. Materials and Methods

2.1. Collection, Isolation and Morphological Analysis

Fallen leaf specimens with leaf spots and dead leaves of Jasminum spp. were collected from different sites in Chiang Mai, Thailand, in October 2019 and 2021, during the wet season. These were carried to the laboratory in paper bags. Single-spore isolation was performed as outlined by Senanayake et al. [16]. Axenic cultures were grown on malt extract agar (MEA, 50 g/L) and incubated for three to four weeks at 25 °C. Appressoria were observed after 24–48 h, forming at the tip of the germ tubes of the conidia. Free-hand sections of conidiomata were performed to examine the morphological characters of each species. Sterilized water was used as the mounting reagent to observe the different fungal features. A Motic SMZ 168 Series stereomicroscope was used to observe their morphology. Micro-morphological characters were photographed with a Canon 750D camera (Canon, Tokyo, Japan) attached to a Nikon ECLIPSE E600 compound microscope (Nikon, Tokyo, Japan). The photo-plates were assembled in Adobe Photoshop CS6 version 2020 (Adobe Systems Inc., San Jose, CA, USA), and measurements were made using Tarosoft® Image Frame Work software (v.0.97).

2.2. Material Deposition and Reference Numbers

The holotype specimens and ex-type living cultures were deposited in the Mae Fah Luang University herbarium (MFLU) and Mae Fah Luang University Culture Collection (MFLUCC), respectively. FacesofFungi “https://www.facesoffungi.org/ (accessed on 20 November 2023)” and Index Fungorum numbers are given [17,18], with the species description updated in the GMS microfungi database “https://gmsmicrofungi.org/ (accessed on 20 November 2023)” [19]. Species identification and the establishment of the new taxon were based on Chethana et al. [20], Jayawardena et al. [21], and Maharachchikumbura et al. [22].

2.3. DNA Extraction, PCR Amplification, and Sequencing

Fresh mycelia, grown and incubated at 25 °C on MEA plates for four weeks, were scraped from the margins of colonies. Genomic DNA was extracted from these mycelia by using the Forensic DNA Kit (D3396-01, OMEGA bio-tek, Inc., Winooski, VT, USA), following the guidelines of the manufacturer. The loci of internal transcribed spacer (ITS, nuclear rDNA consisting of ITS1-5.8S-ITS2) and large subunit (28S, D1–D2 domains of nuclear 28S rDNA), and the genes for beta-tubulin (β-tub), RNA polymerase 2 (Rpb2), and translation elongation factor 1α (tef-1α) were amplified using the following primers: ITS5/ITS4 for ITS [23]; LR0R/LR5 for 28S [23]; Bt2a/Bt2b for β-tub [24]; Rpb2-5F2/7CR for Rpb2 [25,26]; and EF1-728F/EF2 for tef-1α [27,28]. The polymerase chain reaction (PCR) was carried out in an Applied Biosystems C1000 TouchTM Thermal Cycler under the following conditions: Initial denaturation at 95 °C for 3 min; denaturation at 95 °C for 45 s; annealing at 55 °C for 50 s (ITS), 52 °C for 50 s (28S), 58 °C for 1 min 30 s (β-tub, Rpb2, and tef-1α); extension at 72 °C for 1 min; and final extension at 72 °C for 10 min (number of cycles = 40). The PCR mixture, totaling 25 µL, comprised 12.5 µL of Taq mix (PROMEGA GoTaq®, Green master mix, Madison, WI, USA), 1.5 µL of genomic DNA, 1 µL of the forward and reverse primer each, and 9 µL of double-distilled water.
The results of the amplification procedure were visualized using gel electrophoresis (1.7% agarose gel) by loading the resulting amplicons and DNA fluorescent loading dye (FluoroDyeTM, SMOBIO, Seoul, Republic of Korea) in the sample wells. These amplicons were purified, and DNA was sequenced at SolGent Co. (Daejeon, Republic of Korea). Consensus sequences of the forward and reverse DNA sequence data were produced using SeqMan (DNAStar, Madison, WI, USA).
Accession numbers for all sequences deposited in the NCBI GenBank database “https://submit.ncbi.nlm.nih.gov/ (accessed on 20 November 2023)” are listed (Table 1).

2.4. Phylogenetic Analyses

A BLAST search in NCBI “https://blast.ncbi.nlm.nih.gov/ (accessed on 20 November 2023)” was conducted for our sequences, and sequence data of ITS, 28S, β-tub, Rpb2, and tef-1α from related species were retrieved from GenBank (https://www.ncbi.nlm.nih.gov/ accessed on 20 November 2023) (Table 1). Sequences were aligned using MAFFT v.7 by applying the default settings (https://mafft.cbrc.jp/alignment/server/ accessed on 20 November 2023) [29] and trimmed using trimAl [30]. Individual loci were combined using BioEdit v.7.0.5.2 [31]. Phylogenetic trees were constructed using maximum likelihood (ML), maximum parsimony (MP), and Bayesian inference (BI) methods. Both single and combined gene trees were constructed to compare the topology and taxonomic placement of each taxon.
Maximum likelihood analyses (ML-IQ) were performed in the webserver (https://iqtree.cibiv.univie.ac.at/ accessed on 20 November 2023), by selecting the default parameters and 1000 ultrafast bootstrap replicates [32]. Phylogenetic Analysis Using Parsimony (PAUP) v.4.0b10 was used to compute MP analyses [33]. A heuristic search option with the addition of 1000 random sequence additions was applied. Maxtrees and bootstrap replicates were set up to 1000. Bayesian inference analysis (MrBayes on XSEDE v.3.2.7a) was performed in the CIPRES Science Gateway v.3.3 [34,35], after implementing MrModeltest to estimate the model of evolution of individual gene regions [36]. The partition model for each gene region is given (Table 2). Markov chain Monte Carlo (MCMC) sampling with four Markov chains was used to infer posterior probabilities (PP) for 1,000,000, 5,000,000, and 2,000,000 generations for Ciliochorella, Coniella, and Pseudoplagiostoma, respectively. The tree sample frequencies were set to 100. The first 20% of the total trees were discarded as “burn in” and the remaining 80% was used to calculate posterior probabilities.
FigTree v.1.4.4 was used to display the phylogenetic trees [37], and the phylograms were edited and produced in Microsoft PowerPoint (2016).

2.5. Genealogical Concordance Phylogenetic Species Recognition Analysis (GCPSR)

The GCPSR model was applied to scrutinize any significant recombination event that occurred between the new taxon and other phylogenetically closely related species [38], as inferred by a pairwise homoplasy index (Φw) (PHI) test. The analysis was performed in SplitsTree4 by applying the LogDet transformation and splits decomposition options [39,40]. The final layout of the splitsTree graphs was produced in Microsoft PowerPoint (2016).

2.6. Poisson Tree Processes (PTP)

The coalescent-based PTP model was applied to delineate species in Pseudoplagiostoma. The analysis was computed on the Web Server (https://species.h-its.org/ptp/ accessed on 20 November 2023) [41]. The model assumes that the process of speciation is marked by a branching event in the evolutionary tree of a group of organisms, which separates the ancestral lineage into two or more new lineages. The model further assumes that the number of lineages in a group evolves according to a Poisson process, with the rate of speciation being proportional to the number of lineages. The PTP analysis was based on the concatenated ITS, 28S, β-tub, and tef-1α regions. Maximum likelihood analysis prior to computing PTP was conducted on the IQ-tree Web Server. Genetic distances were calculated in MEGA-X by applying the Kimura 2-parameter substitution model and selecting the gamma distribution and pairwise deletion options.

3. Results

3.1. Sequence Alignment and Phylogenetic Analyses

The number of strains used in the phylogenetic analyses of each genus is given (Table 3). Phylogenetic analyses from single and combined gene regions support the identification of the new species (Pseudoplagiostoma jasmini), a Ciliochorella sp., and the new host record, Coniella malaysiana. The phylogenetic trees generated from ML-IQ, MP, and BI yielded similar taxonomic placements for our isolates.
The tef-1α of Pseudoplagiostoma mangiferae was excluded from the phylogenetic analyses because when we used the BLAST tool for P. mangiferae (accession number: MK084822; 100% identity; 100% query cover; e-value = 0.0), the sequence tallied with Melanconis instead of Pseudoplagiostoma.

3.2. Analysis 1: Ciliochorella

Based on the combined ITS, 28S, and β-tub sequence data of Ciliochorella, our isolate, MFLUCC 23-0239, clusters with other Ciliochorella species and forms a distinct lineage with the larger subclade in which reside C. dipterocarpi, C. mangiferae, and C. phanericola (97% ML-IQ, 96% MP, 0.88 PP) (Figure 1).

Genealogical Concordance Phylogenetic Species Recognition Analysis (GCPSR)

The LogDet transformation and splits decomposition options were selected while configuring the PHI test. The analysis yielded a threshold over 0.05 (Φw = 1.0) for the Ciliochorella sp., indicating no significant recombination event (Figure 2).

3.3. Analysis 2: Coniella

Based on the combined ITS, 28S, Rpb2, and tef-1α sequence data of Coniella, our isolate, MFLUCC 23-0240, forms a sister clade with the ex-type of C. malaysiana with 99% ML-IQ and 100% MP bootstrap support, and 1.00 PP support (Figure 3).

3.4. Analysis 3: Pseudoplagiostoma

Based on the combined ITS, 28S, β-tub, Rpb2, and tef-1α sequences of Pseudoplagiostoma, our isolate, MFLUCC 23-0044, groups with other species of Pseudoplagiostoma and forms a sister clade with P. dipterocarpicola (MFLUCC 21-0142 and MFLUCC 21-0114) with 35% ML-IQ and 32% MP bootstrap support, and 0.95 PP support (Figure 4).

3.4.1. Genealogical Concordance Phylogenetic Species Recognition Analysis (GCPSR)

The LogDet transformation and splits decomposition options were selected while configuring the PHI test. The analysis yielded a threshold over 0.05 (Φw = 0.7314) for the new species, Pseudoplagiostoma jasmini, indicating no significant recombination (Figure 5).

3.4.2. Poisson Tree Processes

The result generated from the PTP analysis (Figure 6) is congruent with the maximum likelihood phylogram that delimits Pseudoplagiostoma jasmini as a new species (Figure 4). Genetic distances of Pseudoplagiostoma jasmini compared with its phylogenetically closely related taxa are summarized in the “note” section under Pseudoplagiostoma in the “Taxonomy” section.

4. Taxonomy

4.1. Sporocadaceae Corda [as “Sporocadeae”], Icon. Fung. (Prague) 5: 34 (1842)

This family comprises saprobic, pathogenic, as well as endophytic genera that are commonly characterized by conidia that have appendages at one or both ends. Sporocadaceae has previously been subjected to multiple taxonomic re-evaluations and classifications [42,43]. Bartaliniaceae, Discosiaceae, Pestalotiopsidaceae, and Robillardaceae were previously treated as synonyms of Sporocadaceae [43,44,45].

4.1.1. Ciliochorella Syd., in Sydow & Mitter, Annls Mycol. 33(1/2): 62 (1935)

Type species—Ciliochorella mangiferae Syd.
Ciliochorella (Sporocadaceae, Amphisphaeriales, Xylariomycetidae) [42,43,46,47] was established by Sydow and Mitter [48]. There are ten species in Index Fungorum [18] and nine species in Species Fungorum [49]. Among these, only five Ciliochorella species have sequence data for one or more gene loci. Ciliochorella species comprise saprobic taxa that have been reported from India, Japan, South America, and Thailand [42,50,51,52]. Our isolate is also reported in its saprobic mode.
The genus is characterized by cylindrical, straight, or slightly curved conidia that are eu-septate, usually bearing two to three or more tubular apical appendages and a single basal appendage.

4.1.2. Ciliochorella sp. Gomdola, K.D. Hyde & Jayaward.

Saprobic on the leaves of Jasminum sp. Sexual morph: Not observed. Asexual morph: Coelomycetous. Conidiomata in cross-section 1000–1100 μm diam., 370–380 μm high ( x ¯ = 1042 × 373 μm, n = 5), acervulus, semi-immersed, carbonaceous, solitary, uniloculate, black. Conidiomata wall 40–53 μm diam. ( x ¯ = 46.7 μm, n = 10), consisting of several layers of pseudoparenchymatous cells of textura angularis, outer layers dark brown, inner layers pale brown to hyaline. Conidiophores indistinct, often reduced to conidiogenous cells. Conidiogenous cells phialidic, (5.2–)6.7–8.5(–9.5) × 1.9–2.8 μm ( x ¯ = 7.4 × 2.4 μm, n = 10), formed from the inner-most layer of the wall, hyaline to pale brown, ampulliform, smooth-walled, proliferating enteroblastically. Conidia 11–15 × 2.4–3.8 μm ( x ¯ = 12.9 × 3.3 μm, n = 50) (excluding basal cell), hyaline to pale brown, guttulate, 1-euseptate, smooth-walled, allantoid to sub-cylindrical, or sub-falcate to reniform, apex sometimes broadly obtuse, tapering towards a slightly curved base with a hyaline obconic basal cell 2.8–4.5 μm long ( x ¯ = 3.6 μm, n = 30); conidia bearing 2 apical and 1 basal appendage. Appendages tubular, filiform, flexuous, apical appendages (6.5–)12.5–18.5 μm long ( x ¯ = 16.5 μm, n = 50), basal appendage (2.5–)4–6.5(–8) μm long ( x ¯ = 5.4 μm, n = 50). Appressorium 20 × 18.5 μm, single-celled, cordate to irregular-shaped, hyaline.
Culture characteristics: Colonies on MEA reaching approximately 20 mm diam. after 14 days of incubation at 25 °C, elevation flat, forming concentric rings with an entire margin, mycelium white.
Material examined: Thailand, Chiang Mai Province, Doi Lo district, on fallen dead leaves of Jasminum sp. (Oleaceae), 15 October 2019, D. Gomdola, DG314 (MFLU 23-0388), living culture MFLUCC 23-0239.
GenBank accession numbers: ITS = OR610581, 28S = OR610582.
Notes: Ciliochorella sp. (MFLUCC 23-0239) groups with other Ciliochorella species and forms a separate lineage with the larger subclade in which reside C. dipterocarpi, C. mangiferae, and C. phanericola (97% ML-IQ, 96% MP, 0.88 PP) (Figure 1). The conidial features match the morphological species concept of Ciliochorella. We compared the morphology of Ciliochorella sp. (MFLUCC 23-0239) with that of its phylogenetically closely related, C. phanericola. The conidial shape, color, and size of Ciliochorella sp. (MFLUCC 23-0239) and C. phanericola are mostly similar (Table 4). However, the conidia of Ciliochorella sp. (MFLUCC 23-0239) are 1-euseptate, while those of C. phanericola are 2-septate. Both the apical and basal appendages of Ciliochorella sp. (MFLUCC 23-0239) are shorter than those of C. phanericola (Table 4). The growth rate of Ciliochorella sp. (MFLUCC 23-0239) (2 cm after 14 days) is slower than that of C. phanericola (2.5 cm after 7 days), both grown on MEA and incubated at 25 °C [51]. In addition, appressoria were not observed in C. phanericola [51].
Excluding gaps in our aligned untrimmed dataset, in comparison of the inter-species genetic distance of Ciliochorella sp. (MFLUCC 23-0239) and C. phanericola, a difference of 0.34% was seen across ITS (533 nucleotides), but no difference was observed across 28S (868 nucleotides). We were unable to compare the differences across β-tub as Ciliochorella sp. (MFLUCC 23-0239) lacks sequence data for the gene region. Despite several trials using different amplification conditions, we were unable to obtain sequence data for β-tub. Therefore, coupled with morphological description and multi-locus phylogenetic analyses, a PHI test was also conducted to support the taxonomic placement of our isolate (MFLUCC 23-0239). The PHI test of the combined ITS and 28S yielded a threshold exceeding 0.05 (Φw = 1.0), suggesting that no recombination event has occurred.
Nevertheless, despite the PHI test result, we suggest establishing our isolate as Ciliochorella sp. instead of identifying it as a new species due to the lack of sequence data. Further studies focusing on the collection of more Ciliochorella taxa and providing sequence data for protein-coding gene regions (β-tub, Rpb2, tef-1α) will yield better resolution in the phylogenetic trees and contribute to proper species identification (Figure 7).

4.2. Schizoparmaceae Rossman, D.F. Farr & Castl. [as “Schizoparmeaceae”], Mycoscience 48(3): 137 (2007)

Schizoparmaceae was introduced to accommodate Schizoparme (sexual morph reported), Coniella, and Pilidiella (asexual morph reported) [53,54]. Alvarez et al. [55] revised the family and synonymized Pilidiella and Schizoparme under Coniella. Species in this family occur in tropical and temperate areas as phytopathogens as well as saprobes and endophytes [43,56].

4.2.1. Coniella Höhn., Ber. Dt. Bot. Ges. 36(7): 316 (1918)

Type species—Coniella pulchella Höhn.
Coniella (Schizoparmaceae, Diaporthales, and Diaporthomycetidae) [43,46,47] was established by Höhnel [57]. There are 64 species in Index Fungorum [18] and 58 species in Species Fungorum [49]. Of these, 42 Coniella species have sequence data for one or more gene regions. The genus is primarily characterized by erumpent, brown to black ascomata or conidiomata, and hyaline conidia that become pigmented upon maturation [58,59].

4.2.2. Coniella malaysiana L.V. Alvarez & Crous, in Alvarez, Groenewald & Crous, Stud. Mycol. 85: 21 (2016)

Index Fungorum number: IF 817823, Facesoffungi number: FoF 14882
Associated with leaf spots of Jasminum sp. Leaf spots irregular or oval to elongated, brown, surrounded by a dark brown to black margin, outermost surrounding reddish brown. Sexual morph: Not observed. Asexual morph: Coelomycetous. Conidiomata 135–140 μm diam., 100–130 μm high ( x ¯ = 139 × 114 μm, n = 5), pycnidial, semi-immersed, sometimes erumpent, solitary, scattered or gregarious, uniloculate, globose to subglobose, black. Conidiomata wall 13.5–24.5(–28) μm diam. ( x ¯ = 18.6 μm, n = 10), consisting of 3–4 layers of thick-walled pseudoparenchymatous cells of textura angularis, outer layers dark brown, inner layer pale brown to hyaline. Conidiophores 6.9–15 μm long ( x ¯ = 11.2 μm, n = 10), straight to flexuous, cylindrical to ampulliform or oblong, hyaline, aseptate, unbranched, sometimes reduced to conidiogenous cells. Conidiogenous cells enteroblastic, phialidic, 6.9–13 × 2.1–3.4 μm ( x ¯ = 10.7 × 2.6 μm, n = 10), hyaline, cylindrical or ampulliform, guttulate, smooth-walled. Conidia (7.5–)8.2–13.1 × 4–5 μm ( x ¯ = 10.6 × 4.1 μm, n = 50), hyaline when immature, becoming pale to dark brown upon maturation, guttulate, aseptate, smooth, thick-walled, 0.4–1.5 μm diam. ( x ¯ = 0.75 μm, n = 30), fusiform to truncate to sub-ellipsoidal, sometimes obovoid, wider in the middle, tapering towards a slightly curved apex and base, often with a prominent protruding basal hilum. Appressoria 19–23 × 9–15 μm ( x ¯ = 20.9 × 12 μm, n = 2), single-celled, sub-ellipsoidal to irregular-shaped, hyaline.
Culture characteristics: Colonies on MEA reaching approximately 20 mm diam. after 7 days of incubation at 25 °C, elevation flat or raised, round with raised margin, forming concentric rings, mycelium dense and aerial, white.
Material examined: Thailand, Chiang Mai Province, Omkoi district, Yang Piang sub-district, associated with leaf spots of Jasminum sp. (Oleaceae), 16 October 2019, D. Gomdola, DG392 (MFLU 23-0389), living culture MFLUCC 23-0240.
Hosts and Distribution: Leaves of Corymbia torelliana in Malaysia [55], leaves of Jasminum sp. in Thailand (this study).
GenBank accession numbers: ITS = OR608286, 28S = OR608334, Rpb2 = OR601568 and tef-1α = OR601569.
Notes: Our collection shares similar morphological characters with those of the ex-type, C. malaysiana (CBS 141598) [55]. Our strain and C. malaysiana (CBS 141598) have hyaline to brown, aseptate conidia with guttules [55]. Conidial sizes are mostly similar (Table 5). The conidial length-to-width ratio of our isolate is 2.6, and that of C. malaysiana (CBS 141598) is 2.5. Other morphological similarities and differences between the two strains of C. malaysiana are given (Table 5).
In the phylogenetic analyses of the combined ITS, 28S, Rpb2, and tef-1α, our isolate is sister to the ex-type of C. malaysiana (99% ML-IQ, 100% MP, 1.00 PP) (Figure 3). Excluding gaps in our aligned untrimmed dataset, upon comparison of the intra-species genetic distance between our strain and the ex-type of C. malaysiana, the following differences were observed: 0.55% across ITS (553 nucleotide base pairs, bp), 0.12% across 28S (827 bp), 0.26% across Rpb2 (767 bp), but 2.4% across tef-1α (295 bp).
Based on morphology and multigene phylogenetic analyses, we identify our strain as a new host record of Coniella malaysiana, associated with leaf spots of Jasminum sp. in northern Thailand (Figure 8).

4.3. Pseudoplagiostomataceae Cheew., M.J. Wingf. & Crous [as “Pseudoplagiostomaceae”], in Cheewangkoon et al., Fungal Diversity 44: 95 (2010)

Pseudoplagiostomataceae, a monotypic family, was introduced to accommodate Pseudoplagiostoma, a genus that is morphologically similar but phylogenetically distinct to Plagiostoma [60].

4.3.1. Pseudoplagiostoma Cheew., M.J. Wingf. & Crous, in Cheewangkoon et al., Fungal Diversity 44: 96 (2010)

Type species—Pseudoplagiostoma eucalypti Cheewangkoon, M.J. Wingf. & Crous
Pseudoplagiostoma (Pseudoplagiostomataceae, Diaporthales, and Diaporthomycetidae) was established by Cheewangkoon et al. [60], with the introduction of three species: P. eucalypti, P. oldii, and P. variabile. There are 13 species in Index Fungorum [18] and nine species in Species Fungorum [49], and all 13 species have sequence data for one or more gene loci. The nomenclature of Pseudoplagiostoma reflects the morphological similarities with Plagiostoma (Gnomoniaceae, Diaporthales). Pseudoplagiostoma species have both sexual and asexual morphs [60]. Their sexual morph is characterized by perithecial, immersed, globose or elliptical ascomata, subcylindrical unitunicate asci (J-), and hyaline, ellipsoidal, and elongated ascospores, usually with a median septum and hyaline appendages at the apex and base (Figure 9). Their asexual morph consists of acervular or pycnidial conidiomata, and hyaline, smooth-walled, aseptate conidia (Figure 9) [60].

4.3.2. Pseudoplagiostoma jasmini Gomdola, K.D. Hyde & Jayaward., sp. nov.

Index Fungorum number: IF 900131, Facesoffungi number: FoF 14104
Etymology: The specific epithet refers to the host genus, Jasminum, from which the species was isolated.
Holotype: MFLU 23-0068
Associated with leaf spots of Jasminum grandiflorum. Leaf spots irregular, pale to medium brown, surrounded by a dark brown to black margin. Sexual morph: Not observed. Asexual morph: Coelomycetous. Conidiomata (145–)150–230(–240) μm diam., (135–)140–200 μm high ( x ¯ = 184 × 171 μm, n = 20), pycnidial, semi-immersed, solitary, scattered, uniloculate, globose to subglobose, pale brown, surrounded with black margin. Conidiomata wall (19–)22–42(–46) μm thick ( x ¯ = 28 μm, n = 15), consisting of 3–4 layers of thick-walled pseudoparenchymatous cells of textura angularis, outer layers dark brown to black, inner layers pale brown to hyaline. Conidiophores indistinct, often reduced to conidiogenous cells. Conidiogenous cells phialidic, (6.8–)7.7–13.7(–15.6) × 1.6–2.4(–3.0) μm ( x ¯ = 10.7 × 2.1 μm, n = 10), hyaline, cylindrical or clavate, guttulate, smooth-walled, proliferating enteroblastically. Conidia (11.8–)14–22 × (5.2–)6.5–11 μm (( x ¯ = 18.5 × 9.5 μm, n = 50), hyaline, guttulate, 0–2-septate, smooth, wall (0.5–)0.6–1.3 μm thick ( x ¯ = 0.8 μm, n = 50), cylindrical to truncate or ellipsoidal, elongated, reniform, pyriform or obovoid, apex broadly obtuse, tapering towards a slightly curved base, often with a prominent protruding hilum. Appressorium 9.6 × 7.2 μm long, single-celled, sub-ellipsoidal to obovoid or subglobose or irregular-shaped, hyaline.
Culture characteristics: Colonies on MEA reaching approximately 20 mm diam. after 14 days of incubation at 25 °C, immature with white mycelium, elevation flat or raised, becoming aerial dense and olivaceous brown, filamentous with an undulate margin when aged.
Material examined: Thailand, Chiang Mai Province, Doi Inthanon National Park, Kew Mae Pan nature trail, on fallen leaves of Jasminum grandiflorum (Oleaceae), 20 October 2021, D. Gomdola, DG-PSEU (MFLU 23-0068, holotype), ex-type living culture MFLUCC 23-0044.
GenBank accession numbers: ITS = OQ786078, 28S = OQ786079, β-tub = OQ850148 and tef-1α = OQ850145.
Notes: Pseudoplagiostoma jasmini groups with other species of Pseudoplagiostoma and forms a sister clade with P. dipterocarpicola (MFLUCC 21-0142 and MFLUCC 21-0114) with 35% ML-IQ, 32% MP, and 0.95 PP support (Figure 4). The features are congruent with the morphological species concept of Pseudoplagiostoma [60].
Pseudoplagiostoma jasmini varies substantially in conidial shape (Figure 9). Conidia of P. jasmini are longer than those of P. dipterocarpicola (MFLUCC 21-0142) (Table 6). The conidial length-to-width ratio of P. jasmini is 2.0, whereas that of P. dipterocarpicola is 2.7.
Excluding gaps in our aligned untrimmed dataset, in pairwise nucleotide comparisons of P. jasmini and P. dipterocarpicola (MFLUCC 21-0142), the following differences were observed: 5.76% across ITS (543 nucleotide base pairs, bp), 1.86% across 28S (818 bp), 21.1% across β-tub (448 bp), and 43.7% across tef-1α (164 bp). The inter-species genetic distances (%) grouped according to the PTP result are provided (Table 7).
Based on the guidelines of Chethana et al. [20], Jayawardena et al. [21], and Maharachchikumbura et al. [22] for introducing new species, we describe P. jasmini as a new species. Despite its support values (35% ML-IQ, 32% MP, and 0.95 PP), we establish P. jasmini as a new taxon, considering the formation of one or two septa in the conidia, a feature lacking in all other Pseudoplagiostoma species; all Pseudoplagiostoma spp. have aseptate conidia. Besides morphology and multigene phylogenetic analyses, we included GCPSR and PTP analyses as further evidence to support the distinct species status of Pseudoplagiostoma jasmini (Figure 10).

5. Discussion

In pathology, appressoria are infection structures generated to invade plant tissues [1,4,69,70]. Basically, they are penetration pegs [1]. Appressoria are not solely confined to fungal pathogens. They also occur in endophytes [71,72], epiphytes [3,73,74], and saprobes [75]. In this study, we establish one new species (Pseudoplagiostoma jasmini), a Ciliochorella sp., and a new host record (Coniella malaysiana) that produce single-celled, irregular-shaped, hyaline appressoria. The Ciliochorella sp. is reported from dead leaves as a saprobe, while P. jasmini and C. malaysiana were found associated with leaf spots. In our study, pathogenicity tests were not performed. Therefore, the occurrence of appressoria in C. malaysiana and P. jasmini reveals their pathogenic and possibly endophytic nature. Certain fungi can switch their lifestyles from endophyte to saprobe and become pathogenic under suitable conditions [76]. We hypothesize that, under favorable circumstances, C. malaysiana, P. jasmini, and the saprobic Ciliochorella sp. may develop phytopathogenic traits and cause diseases. Given that appressoria are produced by fungi in various life modes, as mentioned above, it is of dire need to record their occurrences and diversity from different hosts.
This is the first study that reports the formation of appressoria in a Ciliochorella sp. and a Pseudoplagiostoma sp., but appressoria have previously been observed in Coniella musaiaensis [77]. The primary function of appressoria produced by endophytes is to cross from one cell to another [3]. For saprobes to obtain their nutrients, a living host is not a requisite. Thus, the formation of an appressorium in saprobic fungi is probably a result of adaptation while in their endophytic life mode [4,78,79,80].
Species delimitation is essential to developing a proper comprehension of the biology, geography, host-fungal association, and life modes of individual fungal taxa, as well as their respective roles in the ecosystem [20]. The taxonomy of certain Pseudoplagiostoma species yielded low support values when constructing the phylograms (ML-IQ, MP, PP). Despite the support values for the placement of P. jasmini, we establish the latter as a novel taxon as there are significant differences in the conidial morphology. All Pseudoplagiostoma taxa, except P. jasmini, have aseptate conidia. Apart from P. jasmini, all other species of Pseudoplagiostoma are cryptic, sharing similar morphologies such as shape, color, and size. Therefore, coupled with morphology and phylogenetic analyses, we employed the Genealogical Concordance Phylogenetic Species Recognition Analysis (GCPSR) to infer the species boundaries in Pseudoplagiostoma [38]. Furthermore, we advocate the use of the coalescent-based Poisson tree processes (PTP) model to compare the inter- and intra-species genetic distances in Pseudoplagiostoma [41].
Many Sordariomycetes taxa are demarcated based on ITS, 28S, small subunit (18S, nuclear rDNA), β-tub, tef-1α, and Rpb2 loci [43]. Only five Ciliochorella spp., but all Pseudoplagiostoma spp., have molecular data for one or more gene loci. A few Ciliochorella spp. lack sequence data for β-tub. The collection and examination of more Ciliochorella species, with the addition of more gene regions in the phylogenetic analyses, as applied in the analysis and delineation of other Sordariomycetes taxa, would lead to a better phylogenetic resolution and taxonomic placement of each species. Based on high-throughput sequencing, Baldrian et al. [81] suggested that the fungal diversity is around 6.28 million species worldwide but with only 1.08 million published species. A probable reason for the smaller number of Ciliochorella spp. and Pseudoplagiostoma spp. might be because they occur in poorly studied hosts and countries [82]. Northern Thailand is rich in fungal biodiversity [82]. Undoubtedly, further exploration of the fungal diversity in this area as well as other hotspots worldwide will reveal a higher diversity of these two and other genera [83].

Author Contributions

Conceptualization, D.G., E.H.C.M. and D.B.; formal analysis, D.G.; methodology, D.G.; software, D.G.; supervision, E.H.C.M., K.D.H. and R.S.J.; validation, D.G. and D.B.; writing—original draft, D.G.; writing—review and editing, E.H.C.M., K.D.H., D.B. and R.S.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by MFU student scholarship 2020; The National Science and Technology Development Agency (NSTDA: Project No. P-19-52624), under the National Park Permission No. 0907.4/8218 and No. 0907.4/19647; National Research Council of Thailand (NRCT) (grant no. N42A650547), project entitled “Total fungal diversity in a given forest area with implications towards species numbers, chemical diversity, and biotechnology”; and the Mushroom Research Foundation (MRF).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Acknowledgments

We thank Shaun Pennycook for suggesting and validating the nomenclature of the new species. R. S. Jayawardena thanks the Eminent scholar offered by Kyun Hee University.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Chethana, K.W.T.; Jayawardena, R.S.; Chen, Y.J.; Konta, S.; Tibpromma, S.; Phukhamsakda, C.; Abeywickrama, P.D.; Samarakoon, M.C.; Senwanna, C.; Mapook, A.; et al. Appressorial interactions with host and their evolution. Fungal Divers. 2021, 110, 75–107. [Google Scholar] [CrossRef]
  2. Frank, B. Ueber einige neue und weniger bekannte Pfanzenkrankheiten. Ber. Dtsch. Bot. Ges. 1883, 1, 29–34. [Google Scholar]
  3. Emmett, R.W.; Parbery, D.G. Appressoria. Annu. Rev. Phytopathol. 1975, 13, 147–165. [Google Scholar] [CrossRef]
  4. Chethana, K.T.; Jayawardena, R.S.; Chen, Y.J.; Konta, S.; Tibpromma, S.; Abeywickrama, P.D.; Gomdola, D.; Balasuriya, A.; Xu, J.; Lumyong, S.; et al. Diversity and function of appressoria. Pathogens 2021, 10, 746. [Google Scholar] [CrossRef] [PubMed]
  5. The Plant List. Available online: http://www.theplantlist.org/ (accessed on 18 September 2023).
  6. Madhupriya, R.G.P.; Khurana, S.M.P. Rice yellow dwarf phytoplasma (16Sr XI-B subgroup) infecting Jasminum sambac in India. Phytoparasitica 2015, 43, 77–80. [Google Scholar] [CrossRef]
  7. Wallander, E.; Albert, V.A. Phylogeny and classification of Oleaceae based on rps16 and trnL-F sequence data. Am. J. Bot. 2000, 87, 1827–1841. [Google Scholar] [CrossRef]
  8. Huang, Y.L.; Oppong, M.B.; Guo, Y.; Wang, L.Z.; Fang, S.M.; Deng, Y.R.; Gao, X.M. The Oleaceae family: A source of secoiridoids with multiple biological activities. Fitoterapia 2019, 136, 104155. [Google Scholar] [CrossRef]
  9. Sharma, P.C.; Yelne, M.B.; Dennis, T.J.; Joshi, A.; Billore, K.V. Database on Medicinal Plants Used in Ayurveda; Central Council for Research in Ayurveda and Siddha: Chennai, India, 2000. [Google Scholar]
  10. Umamaheswari, M.; Asokkumar, K.; Rathidevi, R.; Sivashanmugam, A.T.; Subhadradevi, V.; Ravi, T.K. Antiulcer and in vitro antioxidant activities of Jasminum grandiflorum L. J. Ethnopharmacol. 2007, 110, 464–470. [Google Scholar] [CrossRef]
  11. Arun, M.; Satish, S.; Anima, P. Phytopharmacological profile of Jasminum grandiflorum Linn. (Oleaceae). Chin. J. Integr. Med. 2016, 22, 311–320. [Google Scholar] [CrossRef]
  12. Wikee, S.; Cai, L.; Pairin, N.; McKenzie, E.H.C.; Su, Y.Y.; Chukeatirote, E.; Thi, H.N.; Bahkali, A.H.; Moslem, M.A.; Abdelsalam, K.; et al. Colletotrichum species from Jasmine (Jasminum sambac). Fungal Divers. 2011, 46, 171–182. [Google Scholar] [CrossRef]
  13. Mathur, R.L.; Mathur, B.L. A new species of Curvularia from the leaves of Jasminum sambac. Curr. Sci. 1959, 28, 448–449. [Google Scholar]
  14. Gao, Y.; Monkai, J.; Gentekaki, E.; Ren, G.C. Dothidea kunmingensis, a novel asexual species of Dothideaceae on Jasminum nudiflorum (winter jasmine) from Southwestern China. Phytotaxa 2021, 529, 43–56. [Google Scholar] [CrossRef]
  15. Ali, B.; Sohail, Y.; Mumtaz, A.S. Puccinia aizazii, a new rust fungus on Jasminum humile from foothills of Himalayan ranges. Sydowia 2017, 69, 131–134. [Google Scholar] [CrossRef]
  16. Senanayake, I.C.; Rathnayaka, A.R.; Marasinghe, D.S.; Calabon, M.S.; Gentekaki, E.; Lee, H.B.; Hurdeal, V.G.; Pem, D.; Dissanayake, L.S.; Wijesinghe, S.N.; et al. Morphological approaches in studying fungi: Collection, examination, isolation, sporulation and preservation. Mycosphere 2020, 11, 2678–2754. [Google Scholar] [CrossRef]
  17. Jayasiri, S.C.; Hyde, K.D.; Ariyawansa, H.A.; Bhat, J.; Buyck, B.; Cai, L.; Dai, Y.C.; Abd-Elsalam, K.A.; Ertz, D.; Hidayat, I.; et al. The Faces of Fungi database: Fungal names linked with morphology, phylogeny and human impacts. Fungal Divers. 2015, 74, 3–18. [Google Scholar] [CrossRef]
  18. Index Fungorum. Available online: http://www.indexfungorum.org/Names/Names.asp (accessed on 18 September 2023).
  19. Chaiwan, N.; Gomdola, D.; Wang, S.; Monkai, J.; Tibpromma, S.; Doilom, M.; Wanasinghe, D.N.; Mortimer, P.E.; Lumyong, S.; Hyde, K.D. https://gmsmicrofungi.org: An online database providing updated information of microfungi in the Greater Mekong Subregion. Mycosphere 2021, 12, 1513–1526. [Google Scholar] [CrossRef]
  20. Chethana, K.T.; Manawasinghe, I.S.; Hurdeal, V.G.; Bhunjun, C.S.; Appadoo, M.A.; Gentekaki, E.; Raspé, O.; Promputtha, I.; Hyde, K.D. What are fungal species and how to delineate them? Fungal Divers. 2021, 109, 1–25. [Google Scholar] [CrossRef]
  21. Jayawardena, R.S.; Hyde, K.D.; de Farias, A.R.G.; Bhunjun, C.S.; Ferdinandez, H.S.; Manamgoda, D.S.; Udayanga, D.; Herath, I.S.; Thambugala, K.M.; Manawasinghe, I.S.; et al. What is a species in fungal plant pathogens? Fungal Divers. 2021, 109, 239–266. [Google Scholar] [CrossRef]
  22. Maharachchikumbura, S.S.; Chen, Y.; Ariyawansa, H.A.; Hyde, K.D.; Haelewaters, D.; Perera, R.H.; Samarakoon, M.C.; Wanasinghe, D.N.; Bustamante, D.E.; Liu, J.K.; et al. Integrative approaches for species delimitation in Ascomycota. Fungal Divers. 2021, 109, 155–179. [Google Scholar] [CrossRef]
  23. White, T.J.; Bruns, T.; Lee, S.; Taylor, J.W. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. PCR Protoc. 1990, 64, 315–322. [Google Scholar] [CrossRef]
  24. Glass, N.L.; Donaldson, G.C. Development of primer sets designed for use with the PCR to amplify conserved genes from filamentous ascomycetes. Appl. Environ. Microbiol. 1995, 61, 1323–1330. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, Y.J.; Whelen, S.; Hall, B.D. Phylogenetic relationships among ascomycetes: Evidence from an RNA polymerse II subunit. Mol. Biol. Evol. 1999, 16, 1799–1808. [Google Scholar] [CrossRef] [PubMed]
  26. Sung, G.H.; Sung, J.M.; Hywel-Jones, N.L.; Spatafora, J.W. A multi-gene phylogeny of Clavicipitaceae (Ascomycota, Fungi): Identification of localized incongruence using a combinational bootstrap approach. Mol. Phylogenet. Evol. 2007, 44, 1204–1223. [Google Scholar] [CrossRef] [PubMed]
  27. O’Donnell, K.; Kistlerr, H.C.; Cigelnik, E.; Ploetz, R.C. Multiple evolutionary origins of the fungus causing panama disease of banana: Concordant evidence from nuclear and mitochondrial gene genealogies. Proc. Natl. Acad. Sci. USA 1998, 95, 2044–2049. [Google Scholar] [CrossRef] [PubMed]
  28. Carbone, I.; Kohn, L.M. A method for designing primer sets for speciation studies in filamentous ascomycetes. Mycologia 1999, 91, 553–556. [Google Scholar] [CrossRef]
  29. Katoh, K.; Rozewicki, J.; Yamada, K.D. MAFFT online service: Multiple sequence alignment, interactive sequence choice and visualization. Brief. Bioinform. 2019, 20, 1160–1166. [Google Scholar] [CrossRef]
  30. Capella-Gutiérrez, S.; Silla-Martínez, J.M.; Gabaldón, T. trimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 2009, 25, 1972–1973. [Google Scholar] [CrossRef] [PubMed]
  31. Hall, T.A. BioEdit: A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp. Ser. 1999, 41, 95–98. [Google Scholar]
  32. Nguyen, L.T.; Schmidt, H.A.; Von Haeseler, A.; Minh, B.Q. IQ-TREE: A Fast and Effective Stochastic Algorithm for Estimating Maximum-Likelihood Phylogenies. Mol. Biol. Evol. 2015, 32, 268–274. [Google Scholar] [CrossRef]
  33. Swoford, D.L. PAUP: Phylogenetic Analysis Using Parsimony, Version 4.0 b10; Sinauer Associates: Sunderland, MA, USA, 2002. [Google Scholar]
  34. Huelsenbeck, J.; Ronquist, F.; Nielsen, R.; Bollback, J.P. Bayesian inference of phylogeny and its impact on evolutionary biology. Science 2001, 294, 2310–2314. [Google Scholar] [CrossRef]
  35. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef] [PubMed]
  36. Nylander, J.A.A. MrModeltest v2; Evolutionary Biology Center, Uppsala University: Uppsala, Sweden, 2004. [Google Scholar]
  37. Rambaut, A.; Drummond, A. FigTree v1. 3.1; Institute of Evolutionary Biology, University of Edinburgh: Edinburgh, UK, 2014. [Google Scholar]
  38. Taylor, J.W.; Jacobson, D.J.; Kroken, S.; Kasuga, T.; Geiser, D.M.; Hibbett, D.S.; Fisher, M.C. Phylogenetic species recognition and species concepts in fungi. Fungal Genet. Biol. 2000, 31, 21–32. [Google Scholar] [CrossRef] [PubMed]
  39. Bruen, T.C.; Philippe, H.; Bryant, D. A simple and robust statistical test for detecting the presence of recombination. Genetics 2006, 172, 2665–2681. [Google Scholar] [CrossRef]
  40. Huson, D.H.; Kloepper, T.; Bryant, D. SplitsTree 4.0-Computation of phylogenetic trees and networks. Bioinformatics 2008, 14, 68–73. [Google Scholar] [CrossRef] [PubMed]
  41. Zhang, J.; Kapli, P.; Pavlidis, P.; Stamatakis, A. A general species delimitation method with applications to phylogenetic placements. Bioinformatics 2013, 29, 2869–2876. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, F.; Bonthond, G.; Groenewald, J.Z.; Cai, L.; Crous, P.W. Sporocadaceae, a family of coelomycetous fungi with appendage-bearing conidia. Stud. Mycol. 2019, 92, 287–415. [Google Scholar] [CrossRef] [PubMed]
  43. Hyde, K.D.; Norphanphoun, C.; Maharachchikumbura, S.S.N.; Bhat, D.J.; Jones, E.B.G.; Bundhun, D.; Chen, Y.J.; Bao, D.F.; Boonmee, S.; Calabon, M.S.; et al. Refined families of Sordariomycetes. Mycosphere 2020, 11, 305–1059. [Google Scholar] [CrossRef]
  44. Crous, P.W.; Carris, L.M.; Giraldo, A.; Groenewald, J.Z.; Hawksworth, D.L.; Hemández-Restrepo, M.; Jaklitsch, W.M.; Lebrun, M.H.; Schumacher, R.K.; Stielow, J.B.; et al. The genera of fungi-fixing the application of the type species of generic names-G 2: Allantophomopsis, Latorua, Macrodiplodiopsis, Macrohilum, Milospium, Protostegia, Pyricularia, Robillarda, Rotula, Septoriella, Torula, and Wojnowicia. IMA Fungus 2015, 6, 163–198. [Google Scholar] [CrossRef]
  45. Senanayake, I.C.; Maharachchikumbura, S.S.N.; Hyde, K.D.; Bhat, J.D.; Jones, E.G.; McKenzie, E.H.C.; Dai, D.Q.; Daranagama, D.A.; Dayarathne, M.C.; Goonasekara, I.D.; et al. Towards unraveling relationships in Xylariomycetidae (Sordariomycetes). Fungal Divers. 2015, 73, 73–144. [Google Scholar] [CrossRef]
  46. Wijayawardene, N.N.; Hyde, K.D.; Al-Ani, L.K.T.; Tedersoo, L.; Haelewaters, D.; Rajeshkumar, K.C.; Zhao, R.L.; Aptroot, A.; Leontyev, D.V.; Saxena, R.K.; et al. Outline of Fungi and fungus-like taxa. Mycosphere 2020, 11, 1060–1456. [Google Scholar] [CrossRef]
  47. Wijayawardene, N.N.; Hyde, K.D.; Dai, D.Q.; Sánchez-García, M.; Goto, B.T.; Saxena, R.K.; Erdoğdu, M.; Selçuk, F.; Rajeshkumar, K.C.; Aptroot, A.; et al. Outline of Fungi and fungus-like taxa—2021. Mycosphere 2022, 13, 53–453. [Google Scholar] [CrossRef]
  48. Sydow, H.; Mitter, J.H. Fungi indici-II. Ann. mycol. 1935, 33, 46–71. [Google Scholar]
  49. Species Fungorum. 2023. Available online: http://www.speciesfungorum.org/Names/Names.asp (accessed on 15 March 2023).
  50. Tangthirasunun, N.; Silar, P.; Bhat, D.J.; Maharachchikumbura, S.S.N.; Wijayawardene, N.W.; Bahkali, A.H.; Hyde, K.D. Morphology and phylogeny of two appendaged genera of coelomycetes: Ciliochorella and Discosia. Sydow 2015, 67, 217–226. [Google Scholar]
  51. Hyde, K.D.; Hongsanan, S.; Jeewon, R.; Bhat, D.J.; McKenzie, E.H.C.; Jones, E.B.G.; Phookamsak, R.; Ariyawansa, H.A.; Boonmee, S.; Zhao, Q.; et al. Fungal diversity notes 367–490: Taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers. 2016, 80, 1–270. [Google Scholar] [CrossRef]
  52. Samaradiwakara, N.P.; de Farias, A.R.G.; Tennakoon, D.S.; Aluthmuhandiram, J.V.; Bhunjun, C.S.; Chethana, K.T.; Kumla, J.; Lumyong, S. Appendage-Bearing Sordariomycetes from Dipterocarpus alatus Leaf Litter in Thailand. J. Fungi 2023, 9, 625. [Google Scholar] [CrossRef]
  53. Castlebury, L.A.; Rossman, A.Y.; Jaklitsch, W.J.; Vasilyeva, L.N. A preliminary overview of the Diaporthales based on large subunit nuclear ribosomal DNA sequences. Mycologia 2002, 94, 1017–1031. [Google Scholar] [CrossRef]
  54. Rossman, A.Y.; Farr, D.F.; Castlebury, L.A. A review of the phylogeny and biology of the Diaporthales. Mycoscience 2007, 48, 135–144. [Google Scholar] [CrossRef]
  55. Alvarez, L.V.; Groenewald, J.Z.; Crous, P.W. Revising the Schizoparmaceae: Coniella and its synonyms Pilidiella and Schizoparme. Stud. Mycol. 2016, 85, 1–34. [Google Scholar] [CrossRef]
  56. Jayawardena, R.S.; Hyde, K.D.; Jeewon, R.; Ghobad-Nejhad, M.; Wanasinghe, D.N.; Liu, N.; Phillips, A.J.; Oliveira-Filho, J.R.C.; da Silva, G.A.; Gibertoni, T.B.; et al. One stop shop II: Taxonomic update with molecular phylogeny for important phytopathogenic genera: 26–50 (2019). Fungal Divers. 2019, 94, 41–129. [Google Scholar] [CrossRef]
  57. Höhnel, F.V. Dritte vorlaufige Mitteilung mycologischer Ergebnisse (Nr. 201–304). Ber. Dtsch. Bot. Ges. 1918, 36, 309–317. [Google Scholar]
  58. Chethana, K.W.T.; Zhou, Y.; Zhang, W.; Liu, M.; Xing, Q.K.; Li, X.H.; Yan, J.Y.; Hyde, K.D. Coniella vitis sp. nov. is the common pathogen of white rot in Chinese vineyards. Plant Dis. 2017, 101, 2123–2136. [Google Scholar] [CrossRef] [PubMed]
  59. Senanayake, I.C.; Jeewon, R.; Chomnunti, P.; Wanasinghe, D.N.; Norphanphoun, C.; Karunarathna, A.; Pem, D.; Perera, R.H.; Camporesi, E.; McKenzie, E.H.C.; et al. Taxonomic circumscription of Diaporthales based on multigene phylogeny and morphology. Fungal Divers. 2018, 93, 241–443. [Google Scholar] [CrossRef]
  60. Cheewangkoon, R.; Groenewald, J.Z.; Verkley, G.J.M.; Hyde, K.D.; Wingfield, M.J.; Gryzenhout, M.; Summerell, B.A.; Denman, S.; Toanun, C.; Crous, P.W. Re-evaluation of Cryptosporiopsis eucalypti and Cryptosporiopsis-like species occurring on Eucalyptus leaves. Fungal Divers. 2010, 44, 89–105. [Google Scholar] [CrossRef]
  61. Crous, P.W.; Summerell, B.A.; Shivas, R.G.; Burgess, T.I.; Decock, C.A.; Dreyer, L.L.; Granke, L.L.; Guest, D.I.; Hardy, G.S.; Hausbeck, M.K.; et al. Fungal Planet description sheets: 107–127. Pers. Mol. Phylogeny Evol. 2012, 28, 138–182. [Google Scholar] [CrossRef] [PubMed]
  62. Crous, P.W.; Luangsa-Ard, J.J.; Wingfield, M.J.; Carnegie, A.J.; Hernández-Restrepo, M.; Lombard, L.; Roux, J.; Barreto, R.W.; Baseia, I.G.; Cano-Lira, J.F.; et al. Fungal Planet description sheets: 785–867. Pers. Mol. Phylogeny Evol. 2018, 41, 238–417. [Google Scholar] [CrossRef]
  63. Suwannarach, N.; Kumla, J.; Lumyong, S. Pseudoplagiostoma dipterocarpi sp. nov., a new endophytic fungus from Thailand. Mycoscience 2016, 57, 118–122. [Google Scholar] [CrossRef]
  64. Bezerra, J.D.; Pádua, A.P.; Oliveira, T.G.; Paiva, L.M.; Guarnaccia, V.; Fan, X.; Souza-Motta, C.M. Pseudoplagiostoma myracrodruonis (Pseudoplagiostomataceae, Diaporthales): A new endophytic species from Brazil. Mycol. Prog. 2019, 18, 1329–1339. [Google Scholar] [CrossRef]
  65. Phookamsak, R.; Hyde, K.D.; Jeewon, R.; Bhat, D.J.; Jones, E.C.G.; Maharachchikumbura, S.S.N.; Raspé, O.; Karunarathna, S.C.; Wanasinghe, D.N.; Hongsanan, S.; et al. Fungal diversity notes 929–1035: Taxonomic and phylogenetic contributions on genera and species of fungi. Fungal Divers. 2019, 95, 1–273. [Google Scholar] [CrossRef]
  66. Mu, T.C.; Zhang, Z.X.; Liu, R.Y.; Li, Z.; Zhang, X.G.; Xia, J.W. Morphological and molecular identification of Pseudoplagiostoma castaneae sp. nov. (Pseudoplagiostomataceae, Diaporthales) in Shandong Province, China. Nova Hedwig. 2022, 114, 171–180. [Google Scholar] [CrossRef]
  67. Tang, X.; Jayawardena, R.S.; Stephenson, S.L.; Kang, J.C. A new species Pseudoplagiostoma dipterocarpicola (Pseudoplagiostomataceae, Diaporthales) found in northern Thailand on members of the Dipterocarpaceae. Phytotaxa 2022, 543, 233–243. [Google Scholar] [CrossRef]
  68. Zhang, Z.; Liu, X.; Tao, M.; Liu, X.; Xia, J.; Zhang, X.; Meng, Z. Taxonomy, Phylogeny, Divergence Time Estimation, and Biogeography of the Family Pseudoplagiostomataceae (Ascomycota, Diaporthales). J. Fungi 2023, 9, 82. [Google Scholar] [CrossRef] [PubMed]
  69. Mendgen, K.; Hahn, M.; Deising, H. Morphogenesis and mechanisms of penetration by plant pathogenic fungi. Annu. Rev. Phytopathol. 1996, 34, 367–386. [Google Scholar] [CrossRef] [PubMed]
  70. Ryder, L.S.; Talbot, N.J. Regulation of appressorium development in pathogenic fungi. Curr. Opin. Plant Biol. 2015, 26, 8–13. [Google Scholar] [CrossRef] [PubMed]
  71. Ma, X.; Nontachaiyapoom, S.; Jayawardena, R.S.; Gentekaki, E.; Zhou, S.; Qian, Y.; Wen, T.; Kang, J. Endophytic Colletotrichum species from Dendrobium spp. in China and northern Thailand. MycoKeys 2018, 43, 23–57. [Google Scholar] [CrossRef] [PubMed]
  72. Green, K.A.; Eaton, C.J.; Savoian, M.S.; Scott, B. A homologue of the fungal tetraspanin Pls1 is required for Epichloë festucae expressorium formation and establishment of a mutualistic interaction with Lolium perenne. Mol. Plant Pathol. 2019, 20, 961–975. [Google Scholar] [CrossRef] [PubMed]
  73. Marasinghe, D.S.; Hongsanan, S.; Zeng, X.Y.; Jones, E.G.B.; Boonmee, S.; Hyde, K.D. Taxonomy and ecology of epifoliar fungi. Mycosphere 2022, 13, 558–601. [Google Scholar] [CrossRef]
  74. Marasinghe, D.S.; Hongsanan, S.; Zeng, X.Y.; Jones, E.B.G.; Chomnunti, P.; Boonmee, S.; Hyde, K.D. Taxonomic monograph of epifoliar fungi. Fungal Divers. 2023, 121, 139–334. [Google Scholar] [CrossRef]
  75. Demoor, A.; Silar, P.; Brun, S. Appressorium: The breakthrough in Dikarya. J. Fungi 2009, 5, 72. [Google Scholar] [CrossRef]
  76. Rai, M.; Agarkar, G. Plant-fungal interactions: What triggers the fungi to switch among lifestyles? Crit. Rev. Microbiol. 2016, 42, 428–438. [Google Scholar] [CrossRef]
  77. Marin-Felix, Y.; Groenewald, J.Z.; Cai, L.; Chen, Q.; Marincowitz, S.; Barnes, I.; Bensch, K.; Braun, U.; Camporesi, E.; Damm, U.; et al. Genera of phytopathogenic fungi: GOPHY 1. Stud. Mycol. 2017, 86, 99–216. [Google Scholar] [CrossRef]
  78. Konta, S.; Hongsanan, S.; Tibpromma, S.; Thongbai, B.; Maharachchikumbura, S.S.N.; Bahkali, A.H.; Hyde, K.D.; Boonmee, S. An advance in the endophyte story: Oxydothidaceae fam. nov. with six new species of Oxydothis. Mycosphere 2016, 7, 1425–1446. [Google Scholar] [CrossRef]
  79. Konta, S.; Hyde, K.D.; Phookamsak, R.; Xu, J.C.; Maharachchikumbura, S.S.N.; Daranagama, D.A.; McKenzie, E.H.C.; Boonmee, S.; Tibpromma, S.; Eungwanichayapant, P.D.; et al. Polyphyletic genera in Xylariaceae (Xylariales): Neoxylaria gen. nov. and Stilbohypoxylon. Mycosphere 2020, 11, 2629–2651. [Google Scholar] [CrossRef]
  80. Konta, S.; Tibpromma, S.; Karunarathna, S.C.; Samarakoon, M.C.; Steven, L.S.; Mapook, A.; Boonmee, S.; Senwanna, C.; Balasuriya, A.; Eungwanichayapant, P.D.; et al. Morphology and multigene phylogeny reveal ten novel taxa in Ascomycota from terrestrial palm substrates (Arecaceae) in Thailand. Mycosphere 2023, 14, 107–152. [Google Scholar] [CrossRef]
  81. Baldrian, P.; Větrovský, T.; Lepinay, C.; Kohout, P. High-throughput sequencing view on the magnitude of global fungal diversity. Fungal Divers. 2021, 114, 539–547. [Google Scholar] [CrossRef]
  82. Hyde, K.D.; Norphanphoun, C.; Chen, J.; Dissanayake, A.J.; Doilom, M.; Hongsanan, S.; Jayawardena, R.S.; Jeewon, R.; Perera, R.H.; Thongbai, B.; et al. Thailand’s amazing diversity: Up to 96% of fungi in northern Thailand may be novel. Fungal Divers. 2018, 93, 215–239. [Google Scholar] [CrossRef]
  83. Hyde, K.D.; Jeewon, R.; Chen, Y.J.; Bhunjun, C.S.; Calabon, M.S.; Jiang, H.B.; Lin, C.G.; Norphanphoun, C.; Sysouphanthong, P.; Pem, D.; et al. The numbers of fungi: Is the descriptive curve flattening? Fungal Divers. 2020, 103, 219–271. [Google Scholar] [CrossRef]
Figure 1. Maximum likelihood phylogram based on the combined ITS, 28S, and β-tub matrices of Ciliochorella. Bootstrap support values (ML-IQ ≥ 80%), maximum parsimony (MP ≥ 80%), and Bayesian posterior probabilities (PP ≥ 0.80) are given above the branches or at the nodes as ML-IQ/MP/PP. Hyphen (-) indicates bootstrap support values below 80% for ML-IQ and MP, and posterior probabilities below 0.80. Allelochaeta acuta (CBS 144168 and CPC 19289) and Discosia ravennica (MFLU 18-0131) are the outgroup taxa. Ex-type and reference strains are in bold, and our isolate is in red.
Figure 1. Maximum likelihood phylogram based on the combined ITS, 28S, and β-tub matrices of Ciliochorella. Bootstrap support values (ML-IQ ≥ 80%), maximum parsimony (MP ≥ 80%), and Bayesian posterior probabilities (PP ≥ 0.80) are given above the branches or at the nodes as ML-IQ/MP/PP. Hyphen (-) indicates bootstrap support values below 80% for ML-IQ and MP, and posterior probabilities below 0.80. Allelochaeta acuta (CBS 144168 and CPC 19289) and Discosia ravennica (MFLU 18-0131) are the outgroup taxa. Ex-type and reference strains are in bold, and our isolate is in red.
Pathogens 12 01407 g001
Figure 2. Split graph derived from the PHI analysis, generated for Ciliochorella. Our isolate is in red.
Figure 2. Split graph derived from the PHI analysis, generated for Ciliochorella. Our isolate is in red.
Pathogens 12 01407 g002
Figure 3. Maximum likelihood phylogram based on the combined ITS, 28S, Rpb2, and tef-1α matrices of Coniella. Bootstrap support values (ML-IQ ≥ 80%) and maximum parsimony (MP ≥ 80%), and Bayesian posterior probabilities (PP ≥ 0.80) are given above the branches or at the nodes as ML-IQ/MP/PP. Hyphen (-) indicates bootstrap support values below 80% for ML-IQ and MP, and posterior probabilities below 0.80. Melanconiella hyperopta (CBS 132231 and CBS 131696) are selected as outgroups. Ex-type and reference strains are in bold, and our isolate is in red.
Figure 3. Maximum likelihood phylogram based on the combined ITS, 28S, Rpb2, and tef-1α matrices of Coniella. Bootstrap support values (ML-IQ ≥ 80%) and maximum parsimony (MP ≥ 80%), and Bayesian posterior probabilities (PP ≥ 0.80) are given above the branches or at the nodes as ML-IQ/MP/PP. Hyphen (-) indicates bootstrap support values below 80% for ML-IQ and MP, and posterior probabilities below 0.80. Melanconiella hyperopta (CBS 132231 and CBS 131696) are selected as outgroups. Ex-type and reference strains are in bold, and our isolate is in red.
Pathogens 12 01407 g003
Figure 4. Maximum likelihood phylogram based on the combined ITS, 28S, β-tub, Rpb2, and tef-1α matrices of Pseudoplagiostoma. Bootstrap support values (ML-IQ ≥ 30%) and maximum parsimony (MP ≥ 30%), and Bayesian posterior probabilities (PP ≥ 0.80) are given above the branches or at the nodes as ML-IQ/MP/PP. Hyphen (-) indicates bootstrap support values below 30% for ML-IQ and MP, and posterior probabilities below 0.80. Apoharknessia eucalypti (CBS 142518), A. eucalyptorum (CBS 142519), and Apoharknessia insueta (CBS 111377 and CBS 114575) are the outgroup taxa. Ex-type and reference strains are in bold, and the new taxon is in bold red.
Figure 4. Maximum likelihood phylogram based on the combined ITS, 28S, β-tub, Rpb2, and tef-1α matrices of Pseudoplagiostoma. Bootstrap support values (ML-IQ ≥ 30%) and maximum parsimony (MP ≥ 30%), and Bayesian posterior probabilities (PP ≥ 0.80) are given above the branches or at the nodes as ML-IQ/MP/PP. Hyphen (-) indicates bootstrap support values below 30% for ML-IQ and MP, and posterior probabilities below 0.80. Apoharknessia eucalypti (CBS 142518), A. eucalyptorum (CBS 142519), and Apoharknessia insueta (CBS 111377 and CBS 114575) are the outgroup taxa. Ex-type and reference strains are in bold, and the new taxon is in bold red.
Pathogens 12 01407 g004
Figure 5. Split graph derived from the PHI analysis, generated for Pseudoplagiostoma. The novel species is in bold red.
Figure 5. Split graph derived from the PHI analysis, generated for Pseudoplagiostoma. The novel species is in bold red.
Pathogens 12 01407 g005
Figure 6. Results generated from the PTP analysis of Pseudoplagiostoma. The analysis was based on the ML-IQ topologies of the concatenated ITS, 28S, β-tub, and tef-1α matrices. Groups of species are denoted by colored branches, with blue-colored branches indicating that they are different species, and red-colored branches representing different strains of the same species. Numbers near the nodes are posterior probabilities. The new taxon is given in bold red.
Figure 6. Results generated from the PTP analysis of Pseudoplagiostoma. The analysis was based on the ML-IQ topologies of the concatenated ITS, 28S, β-tub, and tef-1α matrices. Groups of species are denoted by colored branches, with blue-colored branches indicating that they are different species, and red-colored branches representing different strains of the same species. Numbers near the nodes are posterior probabilities. The new taxon is given in bold red.
Pathogens 12 01407 g006
Figure 7. Ciliochorella sp. (MFLUCC 23-0239) (a) Leaf specimen. (b) Close up of conidiomata on a leaf of Jasminum sp. (c) Section through conidioma. (d,e) Conidiomata wall. (fi) Conidiophores, conidiogenous cells and developing conidia. (jn) Immature and mature conidia with appendages. (o) Germinated conidium (p,q) top (left) and reverse (right) of colonies on MEA after 7 and 14 days of incubation, respectively. (r) Appressorium. Scale bars: (b) = 1 mm, (c) = 100 μm, (e,h,jo) = 10 μm, (d,r) = 20 μm, (f,g,i) = 5 μm.
Figure 7. Ciliochorella sp. (MFLUCC 23-0239) (a) Leaf specimen. (b) Close up of conidiomata on a leaf of Jasminum sp. (c) Section through conidioma. (d,e) Conidiomata wall. (fi) Conidiophores, conidiogenous cells and developing conidia. (jn) Immature and mature conidia with appendages. (o) Germinated conidium (p,q) top (left) and reverse (right) of colonies on MEA after 7 and 14 days of incubation, respectively. (r) Appressorium. Scale bars: (b) = 1 mm, (c) = 100 μm, (e,h,jo) = 10 μm, (d,r) = 20 μm, (f,g,i) = 5 μm.
Pathogens 12 01407 g007
Figure 8. Coniella malaysiana (MFLUCC 23-0240) (a) Herbarium specimen with leaf spots. (b) Close up of conidiomata on a leaf of Jasminum sp. (c) Section through a conidioma. (d) Conidioma wall. (eh) Conidiophores, conidiogenous cells, and developing conidia. (il) Immature and mature conidia. (m) Germinated conidia (n) Top (upper) and reverse (lower) of colony on MEA after 5 and 14 days of incubation. (o,p) Appressoria. Scale bars: (b) = 500 μm, (c) = 50 μm, (d,e,m,o,p) = 10 μm, (fl) = 5 μm.
Figure 8. Coniella malaysiana (MFLUCC 23-0240) (a) Herbarium specimen with leaf spots. (b) Close up of conidiomata on a leaf of Jasminum sp. (c) Section through a conidioma. (d) Conidioma wall. (eh) Conidiophores, conidiogenous cells, and developing conidia. (il) Immature and mature conidia. (m) Germinated conidia (n) Top (upper) and reverse (lower) of colony on MEA after 5 and 14 days of incubation. (o,p) Appressoria. Scale bars: (b) = 500 μm, (c) = 50 μm, (d,e,m,o,p) = 10 μm, (fl) = 5 μm.
Pathogens 12 01407 g008
Figure 9. Conidial (a,do) and ascospores (b,c) morphology of Pseudoplagiostoma spp. (a) P. eucalypti (asexual morph) (b,c) P. eucalypti (ascospores with apical and basal appendages) (d) P. oldii (e) P. dipterocarpi (f) P. variabile (g) P. corymbiicola (h) P. corymbiae (i) P. myracrodruonis (j) P. mangiferae (k) P. dipterocarpicola (l) P. castaneae (m) P. alsophilae (n) P. bambusae (o) P. machili. Scale bars = 15 µm. (Redrawn from Cheewangkoon et al. [60]; Crous et al. [61,62]; Suwannarach et al. [63]; Bezerra et al. [64]; Phookamsak et al. [65]; Mu et al. [66]; Tang et al. [67]; Zhang et al. [68]).
Figure 9. Conidial (a,do) and ascospores (b,c) morphology of Pseudoplagiostoma spp. (a) P. eucalypti (asexual morph) (b,c) P. eucalypti (ascospores with apical and basal appendages) (d) P. oldii (e) P. dipterocarpi (f) P. variabile (g) P. corymbiicola (h) P. corymbiae (i) P. myracrodruonis (j) P. mangiferae (k) P. dipterocarpicola (l) P. castaneae (m) P. alsophilae (n) P. bambusae (o) P. machili. Scale bars = 15 µm. (Redrawn from Cheewangkoon et al. [60]; Crous et al. [61,62]; Suwannarach et al. [63]; Bezerra et al. [64]; Phookamsak et al. [65]; Mu et al. [66]; Tang et al. [67]; Zhang et al. [68]).
Pathogens 12 01407 g009
Figure 10. Pseudoplagiostoma jasmini (MFLUCC 23-0044, ex-holotype) (a) Leaf of Jasminum grandiflorum with spots. (b) Appearance of conidiomata on leaves. (c) Close up of conidioma on substrate. (d) Section through a conidioma. (eh) Conidiophores, conidiogenous cells, and developing conidia. (it) Conidia with guttules, septa and protruding hilum. (uv) Top (left) and reverse (right) of colonies on MEA after 7 and 14 days of incubation, respectively. (w) Appressorium. Scale bars: (b) = 1 mm, (c) = 200 μm, (d) = 100 μm, (et) = 10 μm, (w) = 20 μm.
Figure 10. Pseudoplagiostoma jasmini (MFLUCC 23-0044, ex-holotype) (a) Leaf of Jasminum grandiflorum with spots. (b) Appearance of conidiomata on leaves. (c) Close up of conidioma on substrate. (d) Section through a conidioma. (eh) Conidiophores, conidiogenous cells, and developing conidia. (it) Conidia with guttules, septa and protruding hilum. (uv) Top (left) and reverse (right) of colonies on MEA after 7 and 14 days of incubation, respectively. (w) Appressorium. Scale bars: (b) = 1 mm, (c) = 200 μm, (d) = 100 μm, (et) = 10 μm, (w) = 20 μm.
Pathogens 12 01407 g010
Table 1. GenBank accession numbers of sequences used in the phylogenetic analyses. Ex-type and reference strains are denoted with an ‘*’. Our isolates are in blue.
Table 1. GenBank accession numbers of sequences used in the phylogenetic analyses. Ex-type and reference strains are denoted with an ‘*’. Our isolates are in blue.
SpeciesIsolate NumberITS28Sβ-tubRpb2tef-1α
Allelochaeta acuta CBS 144168 *MH822973MH823023MH823160N/AN/A
Allelochaeta acuta CPC 19289MH822975MH823025MH823162N/AN/A
Apoharknessia eucalypti CBS 142518 *MG934432MN162172MG934505N/AN/A
Apoharknessia eucalyptorumCBS 142519 *KY979752KY979807KY979919N/AN/A
Apoharknessia insueta CBS 111377 *JQ706083AY720814N/AN/AMN271820
Apoharknessia insueta CBS 114575MN172402MN172370N/AN/AMN271821
Ciliochorella castaneaeNBRC 104545N/AAB433277N/AN/AN/A
Ciliochorella castaneaeNBRC 104546N/AAB433278N/AN/AN/A
Ciliochorella dipterocarpiMFLUCC 22-0132 *OP912991OP912990OQ127637N/AN/A
Ciliochorella dipterocarpiMFLUCC 23-0023OQ657982OQ657981OQ657298N/AN/A
Ciliochorella sp.MFLUCC 23-0239OR610581OR610582N/AN/AN/A
Ciliochorella mangiferaeMFLUCC 12-0310KF827444KF827445KF827478N/AN/A
Ciliochorella phanericolaMFLUCC 14-0984 *KX789680KX789681KX789682N/AN/A
ConiellaafricanaCBS 114133 *AY339344AY339293N/AKX833421KX833600
Coniella australiensisIMI 261318N/AN/AN/AN/AN/A
Coniella castaneicolaLGZ2MW672530MW856810N/AN/AN/A
Coniella crousiiNFCCI 2213HQ264189N/AN/AN/AN/A
Coniella diospyriCBS 145071 *MK047439MK047489N/AMK047543MK047562
Coniella diplodiellaCBS 111858 * AY339323 KX833335 N/A KX833423 KX833603
Coniella diplodiellaCBS 111857AY339331AY339286N/AN/AAY339357
Coniella diplodiopsisCBS 590.84 *AY339334AY339288N/AN/AAY339359
Coniella diplodiopsisCBS 10923 AY339332 AY339287 N/A KX833440 KX833624
Coniella duckeraeCBS 142045 * KY924929 N/AN/AN/AN/A
Coniella erumpensCBS 52378 * KX833535 KX833361 N/A KX833446 KX833630
Coniella eucalyptigenaCBS 139893 * KR476725 N/AN/AN/AN/A
Coniella eucalyptorumCBS 112640 * AY339338 AY339290 N/A KX833452 KX833637
Coniella eucalyptorum CBS 114852 KX833556 KX833380 N/A KX833464 KX833652
Coniella ferreirenseCBS 224.80 *MH861257MH873026N/AN/AN/A
Coniella ficiMFLU 18-2578 *MW114356MW114417N/AN/AN/A
Coniella ficiMFLU 18-2579MW114357MW114418N/AN/AN/A
Coniella fragariaeCBS 17249 * AY339317 AY339282 N/A KX833472 KX833663
Coniella fragariae CBS 45468 KX833571 KX833393 N/A KX833477 KX833670
Coniella fusiformisCBS 141596 * KX833576 KX833397 N/A KX833481 KX833674
Coniella granati CBS 132860 KX833577 KX833400 N/A KX833484 KX833677
Coniella heterosporaFMR: 15231LT800501LT800500N/ALT800502LT800503
Coniella hibisciCBS 109757 * KX833589 N/AN/AN/A KX833689
Coniella javanicaCBS 45568 * KX833583 KX833403 N/A KX833489 KX833683
Coniella koreana CBS 14397 KX833584 AF408378 N/A KX833490 KX833684
Coniella lanneaeCBS 141597 * KX833585 KX833404 N/A KX833491 KX833685
Coniella limoniformisCBS 111021 * KX833586 KX833405 N/A KX833492 KX833686
Coniella lustricolaDAOMC 251731 *MF631778MF631799N/AMF651900MF651899
Coniella lustricola DAOMC 251734 MF631781 MF631802 N/AN/AN/A
Coniella macrosporaCBS 52473 * KX833587 AY339292 N/A KX833493 KX833687
Coniella malaysianaCBS 141598 * KX833588 KX833406 N/A KX833494 KX833688
Coniella malaysianaMFLUCC 23-0240OR608286OR608334N/AOR601568OR601569
Coniella musaiaensisAR3534N/AN/AN/AN/AN/A
Coniella nicotianaeCBS 87572 * KX833590 KX833407 N/A KX833495 KX833690
Coniella nigraCBS 16560 * AY339319 KX833408 N/A KX833496 KX833691
Coniella obovataCBS 111025 AY339313 KX833409 N/A KX833497 KX833692
Coniella paracastaneicolaCBS 141292 * KX833591 KX833410 N/A KX833498 KX833693
Coniella peruensisCBS 110394 * KJ710463 KJ710441 N/A KX833499 KX833695
Coniella prostataCOAD 2597MZ727004MZ727000N/AMZ772858MZ772860
Coniella pseudodiospyriCBS 145540 *MK876381MK876422N/AMK876479MK876493
Coniella pseudodiospyriCBS 145541MK876382MK876423N/AMK876480MK876494
Coniella pseudogranatiCBS 137980 *KJ869132N/AN/AN/AN/A
Coniella pseudokoreanaMFLU 13-0282 *MF190146N/AN/AN/AN/A
Coniella pseudostramineaCBS 112624 * KX833593 KX833412 N/A KX833500 KX833696
Coniella quercicolaCBS 90469 * KX833595 KX833414 N/A KX833502 KX833698
Coniella solicolaCBS 76671 * KX833597 KX833416 N/A KX833505 KX833701
Coniella stramineaCBS 14922 AY339348 AY339296 N/A KX833506 KX833704
Coniella tibouchinaeCBS 131595 *JQ281774KX833418N/AKX833507JQ281778
Coniella vitisJZB 3700001 *KX890008KX890083N/AN/AKX890058
Coniella vitisJZB 3700002KX889992KX890067N/AN/AKX890042
Coniella wangiensisCBS 132530 *JX069873JX069857N/AKX833509KX833705
Discosia ravennicaMFLU 18-0131 *MT376615MT376617MT393594N/AN/A
Melanconiella hyperoptaCBS 132231 *MH866004MH877448N/AKX833510KX833706
Melanconiella hyperoptaCBS 131696JQ926281N/AN/AN/AN/A
Pseudoplagiostoma alsophilaeSAUCC WZ0451 *OP810625OP810631OP828586OP828578OP828580
Pseudoplagiostoma alsophilaeSAUCC WZ0152OP810626OP810632OP828587OP828579OP828581
Pseudoplagiostoma bambusaeSAUCC 1206-4 *OP810629OP810635OP828590N/AOP828584
Pseudoplagiostoma bambusaeSAUCC 1206-6OP810630OP810636OP828591N/AOP828585
Pseudoplagiostoma castaneaeSAUCC my0162 *MZ156982MZ156985MZ220325MZ220323MZ220321
Pseudoplagiostoma castaneaeSAUCC my0523MZ156983MZ156986MZ220326MZ220324MZ220322
Pseudoplagiostoma corymbiaeCBS 132529 *JX069861JX069845N/AN/AN/A
Pseudoplagiostoma corymbiicolaCBS 145052 *MK047425MK047476MK047577N/AMK047558
Pseudoplagiostoma dipterocarpiCMUETT57 *KR994682KR994683N/AN/AN/A
Pseudoplagiostoma dipterocarpicolaMFLUCC 21-0142 *OM228844OM228842OM219638N/AOM219629
Pseudoplagiostoma dipterocarpicolaMFLUCC 21-0114OM228843OM228841OM219637N/AOM219628
Pseudoplagiostoma eucalyptiCBS 124807 *GU973512GU973606GU973575N/AGU973542
Pseudoplagiostoma eucalyptiCBS 116382GU973514GU973608GU973577N/AGU973544
Pseudoplagiostoma inthanonenseMFLUCC 23-0262 *OR606510OR633320OR611920OR611921OR650831
Pseudoplagiostoma jasmini sp. nov.MFLUCC 23-0044 * OQ786078OQ786079OQ850148N/AOQ850145
Pseudoplagiostoma machiliSAUCC BW0233 *OP810627OP810633OP828588N/AOP828582
Pseudoplagiostoma machiliSAUCC BW0221OP810628OP810634OP828589N/AOP828583
Pseudoplagiostoma mangiferaeKUMCC 18-0197 *MK084824MK084825MK084823N/AN/A
Pseudoplagiostoma mangiferae8-1.1MN818665MN876855N/AN/AN/A
Pseudoplagiostoma myracrodruonisURM 7799 *MG870421MK982151MN019566MK977723MK982557
Pseudoplagiostoma myracrodruonisURM 8123MK982150MK982152MN019567MK977724MK982558
Pseudoplagiostoma oldiiCBS 124808 *GU973534GU973609GU993862N/AGU973564
Pseudoplagiostoma variabileCBS 113067 *GU973536GU973611GU993863N/AGU973566
N/A: Not applicable.
Table 2. Partition model selected for each locus for the Bayesian analyses.
Table 2. Partition model selected for each locus for the Bayesian analyses.
Model Selected under Akaike Information Criterion (AIC)
Gene Region (s)CiliochorellaConiellaPseudoplagiostoma
ITSHKY+GHKY+GGTR+G
28SGTR+IGTR+I+GGTR+I+G
β-tubGTR+IN/AHKY+I+G
Rpb2N/AGTR+I+GGTR+G
tef-1αN/AHKY+I+GHKY+G
N/A: Not applicable.
Table 3. Total number of characters, ML-IQ, and MP analysis parameters.
Table 3. Total number of characters, ML-IQ, and MP analysis parameters.
CiliochorellaConiellaPseudoplagiostoma
Number of characters in the combined alignment203329752738
Partition of each locusITS: 1–516
28S: 517–1362
β-tub: 1363–2033
ITS: 1–580
28S: 581–1412
Rpb2: 1413–2175
tef-1α: 2176–2493
ITS: 1–557
28S: 558–1375
β-tub: 1376–1855
tef-1α: 1856–2162
Rpb2: 2163–2738
Number of strains used (excluding outgroups)7 (5 species)51 (42 species)23 (15 species)
ML-IQ analysis parameters
ML optimization likelihood value−4043.048−17,415.761−12,799.621
ML Tree length0.2452.5031.687
Distinct alignment patterns140916874
Maximum parsimonious analysis parameters
MP length: Tree #129534072136
Constant1747 20341789
Parsimony-informative257769840
Parsimony-uninformative29172109
Tree #1CI0.9970.4450.678
RI0.9970.6840.802
RC0.9930.3040.543
HI0.0030.5550.322
Table 4. Morphological comparison of Ciliochorella sp. (MFLUCC 23-0239) and C. phanericola.
Table 4. Morphological comparison of Ciliochorella sp. (MFLUCC 23-0239) and C. phanericola.
Species
Species CharactersCiliochorella sp. MFLUCC 23-0239
(This Study)
C. phanericola MFLUCC 14-0984
[51]
ConidiomataSize 1000–1100 μm diam., 370–380 μm high1000–1200 μm diam., 170–200 μm high
Shape and colourSemi-immersed, carbonaceous, sometimes solitary, uniloculate, blackSemi-immersed, circular areas, carbonaceous, sometimes solitary, black
ConidiaSize (μm)11–15 × 2.4–3.813–15 × 2.8–3.5
L/W4.04.1
ShapeAllantoid to sub-cylindrical, or sub-falcate to reniform, apex sometimes broadly obtuse, tapering towards a slightly curved base with an obconic basal cell, smoothAllantoid to sub-cylindrical, smooth
ColourHyaline to pale brownHyaline to pale brown
Septa1-euseptate2-septate
GuttulesPresentPresent
Appendages2 apical, 1 basal, tubular, filiform, flexuous; apical 12.5–18.5 μm long; basal 4–6.5 μm long2 apical, 1 basal, tubular; apical 15–23 μm long; basal 9–11.5 μm long
AppressoriaPresentNot observed
Reported morph(s)AsexualAsexual
Life style(s)SaprobicPathogen or saprobic on leaf
HostsJasminum sp.Phanera purpurea
Gene region(s)ITS, 28SITS, 28S, β-tub
L/W: length-to-width ratio.
Table 5. Morphological comparison between our strain and the ex-type of Coniella malaysiana.
Table 5. Morphological comparison between our strain and the ex-type of Coniella malaysiana.
Species
Species CharactersC. malaysiana MFLUCC 23-0240
(This Study)
C. malaysiana CBS 141598
[55]
ConidiomataSize 135–140 μm diam., 100–130 μm high550 μm diam.
Shape and colorSemi-immersed, sometimes erumpent, solitary, scattered or gregarious, uniloculate, globose to subglobose, blackImmersed or superficial, globose to depressed, initially hyaline, becoming olivaceous to black with age
ConidiaSize (μm)8.2–13.1 × 4–58.5–11 × 3.5–4.5
L/W2.62.5
ShapeSmooth, thick-walled, fusiform to truncate to sub-ellipsoidal, sometimes obovoid, tapering towards a slightly curved apex and base, wider in the middleThick-walled, fusoid to ellipsoid, inequilateral, apex acutely rounded, widest in the middle, tapering to a truncate base
ColorHyaline when immature, becoming pale to dark brown upon maturationHyaline to pale brown
SeptaAseptateAseptate
GuttulesPresentPresent
AppressoriaPresentNot observed
Reported morph(s)AsexualAsexual
Life styleAssociated with leaf spotsPlant pathogenic
HostsJasminum sp.Corymbia torelliana
Gene region(s)ITS, 28S, Rpb2, tef-1αITS, 28S, Rpb2, tef-1α
L/W: Length to width ratio.
Table 6. Morphological comparison of Pseudoplagiostoma jasmini with P. dipterocarpicola.
Table 6. Morphological comparison of Pseudoplagiostoma jasmini with P. dipterocarpicola.
Species
Species CharactersP. jasmini
MFLUCC 23-0044
(This Study)
P. dipterocarpicola
MFLUCC 21-0142
[67]
ConidiomataSize150–230 μm diam., 140–200 μm high113–288 μm diam., 63–153 μm high
Shape and colorPycnidial, semi-immersed, globose to subglobose, pale brown, surrounded with black marginPycnidial with pale yellow cylindrical strips of exuding conidia, subglobose, subcuticular to epidermal, unilocular, irregularly breaking through plant tissue at the center, medium to dark brown
ConidiaSize (μm)14–22 × 6.5–119–22 × 4–7.5
L/W2.02.7
ShapeCylindrical to truncate or ellipsoidal, elongated, reniform, pyriform or obovoid, apex broadly obtuse, tapering towards slightly curved baseEllipsoidal to elongated, apex broadly obtuse, straight, or slightly curved at base, often slightly narrow at middle, base tapering to flat protruding scar
ColorHyalineHyaline
Septa0–2-septateAseptate
GuttulesPresentPresent
Conidial wall (μm)Smooth, 0.6–1.3Smooth, 0.5–1.5
HilumMostly present and prominentPresent or absent
AppressoriaPresentNot observed
Reported morph(s)AsexualAsexual
Life styleAssociated with leaf spotsAssociated with twigs and fruits
HostsJasminum grandiflorumDipterocarpus sp.
Gene region(s)ITS, 28S, β-tub, tef-1αITS, 28S, β-tub, tef-1α
L/W: Length-to-width ratio.
Table 7. Genetic distance (%) between Pseudoplagiostoma species (grouped according to PTP results) in the concatenated ITS, 28S, β-tub, and tef-1α genetic markers.
Table 7. Genetic distance (%) between Pseudoplagiostoma species (grouped according to PTP results) in the concatenated ITS, 28S, β-tub, and tef-1α genetic markers.
Group 1 (%)Group 2 (%)Group 3 (%)Group 4 (%)
Group 1: P. machiliN/A5.6711.012.3
Group 2: P. alsophilae5.67N/A10.111.8
Group 3: P. dipterocarpicola11.010.1N/A9.11
Group 4: P. jasmini12.311.89.11N/A
N/A: not applicable.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gomdola, D.; McKenzie, E.H.C.; Hyde, K.D.; Bundhun, D.; Jayawardena, R.S. Appressoria-Producing Sordariomycetes Taxa Associated with Jasminum Species. Pathogens 2023, 12, 1407. https://doi.org/10.3390/pathogens12121407

AMA Style

Gomdola D, McKenzie EHC, Hyde KD, Bundhun D, Jayawardena RS. Appressoria-Producing Sordariomycetes Taxa Associated with Jasminum Species. Pathogens. 2023; 12(12):1407. https://doi.org/10.3390/pathogens12121407

Chicago/Turabian Style

Gomdola, Deecksha, Eric H. C. McKenzie, Kevin D. Hyde, Digvijayini Bundhun, and Ruvishika S. Jayawardena. 2023. "Appressoria-Producing Sordariomycetes Taxa Associated with Jasminum Species" Pathogens 12, no. 12: 1407. https://doi.org/10.3390/pathogens12121407

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