Biodiversity Data Journal : Taxonomy & Inventories
PDF
Taxonomy & Inventories
Ganoderma ovisporum sp. nov. (Polyporales, Polyporaceae) from Southwest China
expand article infoHong-De Yang‡,§,|, Yong Ding, Ting-Chi Wen¶,§,#, Kalani Kanchana Hapuarachchi¤,§,|,#, De-Ping Wei|,¤,§,#,«
‡ Key Laboratory of Forest Biotechnology in Yunnan, Southwest Forestry University, Kunming, China
§ The Engineering Research Center of Southwest Bio–Pharmaceutical Resources Ministry of Education, Guizhou University, Guiyang, China
| Center of Excellence in Fungal Research, Mae Fah Luang University, Chiang Rai, Thailand
¶ State Key Laboratory Breeding Base of Green Pesticide and Agricultural Bioengineering, Guiyang, China
# The Mushroom Research Centre, Guizhou University, Guiyang, China
¤ State Key Laboratory Breeding Base of Green Pesticide and Agricultural Bioengineering, Key Laboratory of Green Pesticide and Agricultural Bioengineering, Ministry of Education, Guizhou University, Guiyang, China
« Department of Entomology and Plant Pathology, Faculty of Agriculture, Chiang Mai University, Chiang Mai, Thailand
Open Access

Abstract

Background

Ganoderma is a white-rot fungus with a cosmopolitan distribution and includes several economically important species. This genus has been extensively researched due to its beneficial medicinal properties and chemical constituents with potential nutritional and therapeutic values. Traditionally, species of Ganoderma were identified solely based on morphology; however, recent molecular studies revealed that many morphology-based species are conspecific. Furthermore, some type species are in poor condition, which hinders us from re-examining their taxonomic characteristics and obtaining their molecular data. Therefore, new species and fresh collections with multigene sequences are needed to fill the loopholes and to understand the biological classification system of Ganoderma.

New information

In a survey of Ganoderma in Guizhou Province, southwest China, we found a new species growing on soil and, herein, it was identified by both morphology and phylogenetic evidence. Hence, we propose a new species, Ganoderma ovisporum sp. nov. This species is characterised by an annual, stipitate, laccate basidiome, with a red–brown to brownish-black pileus surface and pale white pores, duplex context, clavate pileipellis terminal cells, trimitic hyphal system, ellipsoid basidiospores with dark brown eusporium bearing coarse echinulae and an obtuse turgid appendix. Phylogenetic analyses confirmed that the novel species sisters to G. sandunense with high bootstrap support. Furthermore, the RPB2 sequence of G. sandunense is supplied for the first time. Notably, we re-examined the type specimen of G. sandunense and provide a more precise description of the duplex context, pileipellis terminal cells and basidia. All species collected are described and illustrated with coloured photographs. Moreover, we present an updated phylogeny for Ganoderma, based on nLSU, ITS, RPB2 and TEF1-α DNA sequence data and species relationships and classification are discussed.

Keywords

one new species, Ganoderma, morphology, phylogeny, taxonomy

Introduction

Ganodermataceae Donk is a large family of Polyporales and Ganoderma P. Karst is the most speciose genus in the family (Hapuarachchi et al. 2016a, Hapuarachchi et al. 2017, He et al. 2019). Before the molecular era, Polyporales with double-walled basidiospores with a pigmented endosporium ornamented with columns or ridges and a smooth hyaline exosporium were usually placed in Ganodermataceae (Moncalvo and Ryvarden 1997). This family is comprised of ten genera: Amauroderma Murrill, Amaurodermellus Costa-Rezende, Drechsler-Santos & Góes-Neto, Foraminispora Robledo, Costa-Rezende & Drechsler-Santos, Furtadomyces Leonardo-Silva, Cotrim & Xavier-Santos, Ganoderma P. Karst, Haddowia Steyaert, Humphreya Steyaert, Sanguinoderma Y.F. Sun, D.H. Costa & B.K. Cu,Tomophagus Murrill and Trachyderma Imazeki (Richter et al. 2014, Costa-Rezende et al. 2017, Costa-Rezende et al. 2020, Sun et al. 2020, Leonardo-Silva et al. 2022). However, Ganodermataceae has been treated as a synonym of Polyporaceae (Justo et al. 2017). There have been several discrepancies regarding the treatment of Justo et al. (2017); in particular, the studied collection of Ganodermatoid specimens was insufficient to establish a stable taxonomic and systematic placement in a phylogenetic context because some herbarium materials have been destroyed or cannot be found, lacking molecular and morphological data and the characterised double-walled basidiospores in Ganodermataceae are quite different from those in Polyporaceae (Cui et al. 2019, Costa-Rezende et al. 2020). In this study, we subsequently followed Justo et al. (2017) since the phylogenetic analyses are more convincing and objective than morphological results.

Ganoderma was introduced by Karsten (1881) and typified by G. lucidum (Curtis) P. Karst. (syn. Polyporus lucidus; bas. Boletus lucidus Curtis), a species with stipitate and laccate white-rot Polypore fungi (Karsten 1881, Pegler and Young 1973, Moncalvo and Ryvarden 1997, Keypour et al. 2020). The membership of Ganoderma has been subsequently extended, including species with sessile, non-laccate basidiocarps and pigmented, ellipsoid to ovoid, ornamented, double-walled basidiospores (Murrill 1902, Pegler and Young 1973, Steyaert 1980, Cao and Yuan 2013, Papp 2016). Moncalvo and Ryvarden (1997) accepted 148 Ganoderma species before the molecular era, of which 65% are recognised as only one or some species, but represented different morphology-based species (Ryvarden 2000, Smith and Sivasithamparam 2003, Torres-Torres and Guzmán-Dávalos 2012). Recently, 180 species of Ganoderma were accepted, whereas nearly 500 species are estimated worldwide, of which 60% are awaiting discovery (He et al. 2019, He et al. 2022).

Despite their economic importance, the taxonomy of Ganoderma remains uncertain due to a slew of confusion and misconceptions. During the past several decades, many species of Ganoderma have been delimited, based on the presence of stipe, laccate or non-laccate, the context of pileus and the microscopic characteristics of basidiospores (Chang and Chen 1984, Seo and Kitamot 1998, Wu et al. 2004, Torres-Torres and Guzmán-Dávalos 2012, Zhou et al. 2016, Tchotet-Tchoumi et al. 2019). In general, it is difficult and subjective to identify Ganoderma species solely based on morphological evidence, as their phenotypic traits are sensitive to extrinsic factors, such as illumination, ventilation and humidity (Szedlay et al. 1999, Demoulin 2010, Yang and Feng 2013, Hapuarachchi et al. 2019a). Therefore, morphology-based identification brought Ganoderma into a state of taxonomic chaos (Smith and Sivasithamparam 2003, Coetzee et al. 2015, López-Peña et al. 2019, Náplavová et al. 2020). Compared to morphology, molecular methods have turned out to be more effective in resolving intraspecific relationships with Ganoderma (Yamashita and Hirose 2016, Fryssouli et al. 2020, Gunnels et al. 2020, Jiang et al. 2021, Shen et al. 2021). Phylogenetic markers, such as IGS, nrSSU, ITS, nrLSU, mtSSU, β-TUB, RPB1, RPB2 and TEF1-α sequences, were independently or conjointly used to infer intraspecific relationships within Ganoderma (Cao et al. 2012, Zhou et al. 2015, Xing et al. 2018, Hapuarachchi et al. 2019a, Liu et al. 2019, Ye et al. 2019). In particular, the multilocus phylogeny incorporating sequences from ITS, nrLSU, TEF1-α and RPB2 was applied to give a phylogenetic framework for species delimitation in this genus (Xing et al. 2018, Ye et al. 2019, Tchotet-Tchoumi et al. 2019, Wu et al. 2020, He et al. 2021, Cao et al. 2021). Furthermore, some researchers steered using a combination of morphological, chemotaxonomic and molecular strategies to elevate a steady taxonomy for Ganoderma and resolve taxonomic ambiguities (Richter et al. 2014, Welti et al. 2015).

Ganoderma has a cosmopolitan distribution and most of the species are known from tropical and sub-tropical regions (He et al. 2019). This fungus grows as saprobes or parasites on deciduous and coniferous trees and some of them are considered as plant pathogens that cause basal stem butt rot and root rot (Pinruan et al. 2010, Ding et al. 2020, Mafia et al. 2020, Mohd et al. 2020). Species of Ganoderma play an important role in the nutrient mobilisation process of woody plants. They possess lignocellulose–decomposing enzymes with effective mechanisms for bioenergy production and bioremediation (Coetzee et al. 2015, Kües et al. 2015). In the natural environment, a basidiome has the ability to produce innumerable basidiospores that can be spread by air- or rain-driven and insect vectors (Tuno 1999, Kadowaki et al. 2011, Almaguer et al. 2014, Sadyś et al. 2014). The infection of a plant host by pathogenic Ganoderma species starts with the landing of the basidiospore on the wound trunk or root, followed by germination and colonisation (Rees et al. 2009, Rees et al. 2012, Hushiarian et al. 2013, Ayin et al. 2019). Basal stem rot caused by G. boninense is the main disease that leads to yield losses and death of oil palm, which account for 50% of substantial economic losses to Southeast Asia’s palm oil industry (Hushiarian et al. 2013, Lee and Chang 2016, Midot et al. 2019). Red roots caused by G. philippii are a serious disease of commercial Acacia mangium in Malaysia and India (Glen et al. 2014). Since different Ganoderma species produce different characteristics and pathogenicity, species identification is difficult, which in turn, leads to significant difficulty in disease control (Wong et al. 2012).

Ganoderma was first reported from China by Teng (1934), with four species including G. lucidum and one variety. More than 80 species have been introduced so far and several extensive studies have been carried out to investigate Ganoderma diversity in China, with new species being introduced (Zhao and Zhang 2000, Wu et al. 2004, Dai et al. 2009, Cao et al. 2012, Hapuarachchi et al. 2015, Hapuarachchi et al. 2018c, Wu et al. 2019, Liu et al. 2019). However, the majority of Ganoderma species reported from China have not been subjected to systematic studies (Wang 2012, Hapuarachchi et al. 2016b, Hapuarachchi et al. 2018a, Wang et al. 2019). The objective of the present study is to introduce a novel Ganoderma species, from Guizhou Province, southwest China, with descriptions, colour photographs, illustrations and a multigene phylogeny.

Materials and methods

Ganoderma samples were collected from Sandong Township, Sandu Shuizu Autonomous County, Guizhou Province, China, during the rainy season of July 2020. They were dried and preserved as outlined in Hapuarachchi et al. (2019b). The materials used in this study were deposited at Guizhou University (GACP) and the Herbarium of Kunming Institute of Botany Academia Sinica (HKAS).

Morphological study

Macro-morphological characteristics were described, based on dried material and the photographs provided here. Colour codes (e.g. 8E8) are from Kornerup and Wanscher (1978). Pileus was sectioned with a razor blade and mounted in 5% potassium hydroxide (KOH) solution. Pileipellis, hyphal systems of pileus, basidia and basidiospores were observed and captured using a compound microscope (Leica DM2500) equipped with a camera. Images were measured with Leica Application Suite X (LAS X). In the description section, the number, length, width and length/width ratio of the measured basidiospores are denoted with symbols n, L, W and Q, respectively. The Faces of Fungi number was registered by following Jayasiri et al. (2015).

DNA Extraction, PCR and Sequencing

Genomic DNA was extracted from dried specimens using an HP Fungal DNA Kit (OMEGA, USA) following the protocol of the manufacturer. PCR amplification was performed in a final volume of 50 µl reaction mixture that contained 25 µl 2x BenchTopTM Taq Master Mix (Biomigas), 19 µl distilled water, 2 µl (10 µM) of each primer and 2 µl template DNA. The large subunit ribosomal RNA (LSU), the internal transcribed spacer (ITS), the translation elongation factor (TEF1-α) and the RNA polymerase II second largest subunit (RPB2) were amplified with primer pairs LROR/LR5 (Vilgalys and Hester 1990), ITS5/ITS4 (White et al. 1990), EF1-983F/EF1-1567R (Rehner and Buckley 2017) and RPB2-5f/RPB2-7cR (Liu et al. 1999). PCR amplification reactions were performed with a T100 Thermal Cycler (T100™, Bio-Rad, USA). The procedures used for amplification of ITS were as follows: initial denaturation at 95°C for 3 min, followed by 35 cycles of denaturation at 95°C for 30 s, annealing at 58°C for 30 s, elongation at 72°C for 1 min and a final extension at 72°C for 5 min. The cycling conditions of LSU, TEF1-α and RPB2 consisted of initial denaturation at 95°C for 3 min, followed by 35 cycles of denaturation at 95°C for 30 s, annealing at 56°C for 30 s, elongation at 72°C for 1.3 min and a final extension at 72°C for 10 min. PCR products were verified by 1% agarose gel electrophoresis and sent to Sangon Biotech (Shanghai, China) for purification and sequencing.

Sequence Alignment and Phylogenetic Analysis

The raw sequences generated in this study were assembled with ChromasPro (2.1.8). Megablast analysis was conducted using the assembled ITS and RPB2 sequences as the query to check the closely-related taxa. The taxa used in our phylogenetic analysis were selected, based on megablast results and related publications (Table 1). Alignments were performed using MAFFT v. 7 (http://mafft.cbrc.jp/alignment/server/index.html, Katoh and Standley 2013). The resulting alignments were improved manually when necessary, using BioEdit v. 7.0.5.2 (Hall 1999). The introns in TEF and RPB2 were removed, based on the published CDS sequence in GenBank. The aligned ITS1, 5.8S, ITS2, LSU, TEF1-α and RPB2 sequences were concatenated with SequenceMatrix v.1.7.8 (Vaidya et al. 2011). Maximum Likelihood (ML) analysis was performed using RAxMLHPC2 (Stamatakis 2014) on the CIPRES Science Gateway v. 3.3 (Miller et al. 2010). The phylogenetic tree was inferred from four gene-partition analyses, using the GTRCAT model with 25 categories, with settings that the number of bootstrap replicates to 1,000. PartitionFinder v.2 (Lanfear et al. 2017) was used to estimate the best-fit model of nucleotide evolution, with the dataset subdivided into 10 data partitions (TEF 1st codon positions, TEF 2nd codon positions and TEF 3rd codon positions; RPB2 1st codon positions, RPB2 2nd codon positions and RPB2 3rd codon positions; ITS1; 5.8S; ITS2; LSU) and the following settings: branch lengths = unlinked, models = all, model_selection = AICc and search = greedy. Bayesian Inference (BI) analysis was performed in the CIPRES Science Gateway using MrBayes on XSEDE v. 3.2.7a. The GTR+F+I+G4 (TEF 1st codon positions, TEF 2nd codon positions, RPB2 1st codon positions, RPB2 2nd codon positions, LSU and 5.8S), GTR+F+G4 (TEF 3rd codon positions), GTR+F+G4 (RPB2 2nd codon positions), SYM+G4 (ITS1 and ITS2) were selected as the best model. Two runs of four chains were run until the average standard deviation of split frequencies dropped below 0.01, which occurred after 2,360,000 generations. Tree was sampled every 1000th generation and the chain temperature was decreased to 0.05 to improve convergence. The convergence of the runs was checked using TRACER v.1.6 (Rambaut et al. 2013). The first 25% of the resulting samples were discarded as burn-in and posterior probabilities were calculated from the remaining sampled trees (Larget and Simon 1999). In both ML and BY analyses, Foraminispora concentrica (Cui 12644) and Foraminispora yinggelingensis (Cui 13618) were selected as the outgroup (Sun et al. 2020). ML bootstrap values and BY posterior probabilities greater than or equal to 70% and 0.95, respectively, were considered significant support. The phylogenetic tree was visualised with FigTree version 1.4.0 available at http://tree.bio.ed.ac.uk/software/figtree/ (Rambaut 2012).

Table 1.

The species, specimens and GenBank accession numbers of sequences used in this study

Species Voucher Geographic origin GenBank accession numbers References
ITS LSU EF-1 RPB2
G. adspersum SFC20141001-16 Korea KY364251 KY393284 KY393270 Jargalmaa et al. (2017)
G. adspersum SFC20160115-20 Korea KY364254 KY393286 KY393272 Jargalmaa et al. (2017)
G. angustisporum Cui 14578 China MG279171 MG367564 Xing et al. (2018)
G. angustisporum Cui 13817 (T) China MG279170 MG367563 MG367507 Xing et al. (2018))
G. applanatum SFC20150930-02 Korea KY364258 KY393288 KY393274 Jargalmaa et al. (2017)
G. applanatum Wei 5787a China KF495001 KF494978 GenBank
G. aridicola Dai 12588 (T) Africa KU572491 KU572502 Xing et al. (2016)
G. australe DHCR417 (HUEFS) Australia MF436676 MF436673 MF436678 Costa-Rezende et al. (2017)
G. australe ZRL20151500 China LT716076 KY418900 KY419088 Zhao et al. (2017)
G. boninense WD 2085 Japan KJ143906 KJ143925 KJ143965 Zhou et al. (2015)
G. boninense WD 2028 Japan KJ143905 KU220015 KJ143924 Zhou et al. (2015)
G. carnosum MUCL 49464 France MG706220 MG706168 MG837838 MG837793 GenBank
G. carnosum GC011ND Slovakia MK415266 MK995647 Náplavová et al. (2020)
G. carocalcareum DMC 322 (T) Cameroon EU089969 Douanla-Meli and Langer (2009)
G. casuarinicola Dai 16339 China MG279176 MG367568 MG367511 Xing et al. (2018)
G. casuarinicola Dai 16336 (T) China MG279173 MG367565 MG367508 Xing et al. (2018)
G. chalceum URM80457 Brazil JX310812 JX310826 GenBank
G. concinnum Robledo 3235 MN077523 MN077557 Costa-Rezende et al. (2020)
G. concinnum Robledo 3192 MN077522 MN077556 Costa-Rezende et al. (2020)
G. curtisii CBS 100131 USA JQ781848 KJ143926 KJ143966 Zhou et al. (2015)
G. curtisii CBS 100132 USA JQ520164 KJ143927 KJ143967 Zhou et al. (2015)
G. destructans CBS 139793 (T) South Africa NR_132919 NG_058157 Coetzee et al. (2015)
G. destructans Dai 16431 South Africa MG279177 MG367569 MG367512 Xing et al. (2018)
G. dianzhongense L4331(T) China MW750237 MZ467043 He et al. (2021)
G. dianzhongense L4737 China MW750238 MW839000 He et al. (2021)
G. ecuadorense URM 89449 Brazil MK119828 MK119908 MK121577 MK121535 Sun et al. (2020)
G. ecuadorense URM 89441 Brazil MK119827 MK119907 MK121576 MK121534 Sun et al. (2020)
G. eickeri CMW50325 Africa MH571689 MH567290 Tchotet-Tchoumi et al. (2019)
G. eickeri CMW 49692 (T) Africa NR_165524 Tchotet-Tchoumi et al. (2019)
G. ellipsoideum MFLU 19-2221 China MN398339 MN428664 MN423157 GenBank
G. ellipsoideum GACP 14080966 (T) China NR_160617 Hapuarachchi et al. (2018c)
G. enigmaticum Dai 15971 South Africa KU572487 KU572497 MG367514 Xing et al. (2016)
G. enigmaticum Dai 15970 South Africa KU572486 KU572496 MG367513 Xing et al. (2016)
G. esculentum L4935 (T) China MW750242 MW839004 He et al. (2021)
G. esculentum L4946 China MW750243 He et al. (2021)
G. flexipes VT17102301 Viet Nam MK345430 MK346830 Hapuarachchi et al. (2019b)
G. flexipes Wei5491 China JQ781850 KJ143968 Cao et al. (2012)
G. gibbosum SFC20150918-08 Korea KY364271 KY393291 KY393278 Jargalmaa et al. (2017)
G. gibbosum SFC20150918-03 Korea KY364270 KY393290 KY393277 Jargalmaa et al. (2017)
G. hoehnelianum Dai 11995 China KU219988 KU220016 MG367550 MG367497 Xing et al. (2018)
G. hoehnelianum Cui 13982 China MG279178 MG367570 MG367515 Xing et al. (2018)
G. knysnamense CMW 47755 (T) South Africa NR_165523 MH567261 Tchotet-Tchoumi et al. (2019)
G. knysnamense CMW49688 Africa MH571683 MH567266 Tchotet-Tchoumi et al. (2019)
G. leucocontextum Dai 15601 China KU572485 KU572495 MG367516 Xing et al. (2016)
G. leucocontextum GDGM 40200 (T) China KM396272 Li et al. (2015)
G. lingzhi Dai12574 (IFP) China KJ143908 JX029977 JX029981 Zhou et al. (2015)
G. lingzhi Cui9166 China KJ143907 JX029974 JX029978 Cao et al. (2012)
G. lobatum JV 1008/31 USA KF605671 MG367553 MG367499 Xing et al. (2018)
G. lobatum JV 1008/32 USA KF605670 MG367554 MG367500 Xing et al. (2018)
G. lucidum BR 4195 France KJ143909 KJ143969 Zhou et al. (2015)
G. lucidum K 175217 Italy KJ143911 KJ143929 KJ143971 Zhou et al. (2015)
G. lucidum Cui 14405 China MG279182 MG367574 MG367520 Xing et al. (2018)
G. lucidum CCBAS 707 Europe MG706231 MG706177 MG837846 MG837805 GenBank
G. martinicense UMNTN1 USA MG654178 MG754738 MG754860 Loyd et al. (2018)
G. martinicense He 2240 USA MG279163 MG367557 MG367503 Xing et al. (2018)
G. mbrekobenum UMN7-4 GHA Ghana KX000898 KX000899 Crous et al. (2016)
G. mbrekobenum UMN7-3 GHA Ghana KX000896 KX000897 Crous et al. (2016)
G. meredithiae UMNFL50 USA MG654103 MG754735 MG754862 Loyd et al. (2018)
G. meredithiae CBS 271.88 (T) USA NR_164435 NG_067432 Vu et al. (2019)
G. meredithiae UMNFL64 USA MG654188 MG754734 MG754861 Loyd et al. (2018)
G. mexicanum MUCL: 55832 Martinique MK531815 MK531829 MK531839 Cabarroi-Hernández et al. (2019)
G. mexicanum MUCL: 49453 Martinique MK531811 MK531825 MK531836 Cabarroi-Hernández et al. (2019)
G. mizoramense UMN-MZ5 India KY643751 Crous et al. (2017)
G. mizoramense UMN-MZ4 (T) India KY643750 Crous et al. (2017)
G. multipileum Cui 14373 China MG279184 MG367575 MG367521 Xing et al. (2018)
G. multipileum Dai 9447 China KJ143914 KJ143932 KJ143973 Zhou et al. (2015)
G. multiplicatum Dai 12320 China KU572490 KU572500 Xing et al. (2016)
G. multiplicatum Dai 13710 China KU572489 KU572499 Xing et al. (2016)
G. mutabile Yuan2289 China JN383977 Cao and Yuan (2013)
G. mutabile CLZhao 982 China MG231527 Cao and Yuan (2013)
G. nasalaense LPDR17060212 Laos MK345442 MK346832 Hapuarachchi et al. (2019b)
G. nasalaense GACP 17060211 (T) Laos NR_164048 NG_066439 Hapuarachchi et al. (2019b)
G. orbiforme Cui 13880 China MG279187 MG367577 MG367523 Xing et al. (2018)
G. orbiforme Cui 13918 China MG279186 MG367576 MG367522 Xing et al. (2018)
G. oregonense JV 0108/93 USA KF605620 MG367558 MG367504 Xing et al. (2018)
G. oregonense CBS 265.88 USA JQ781875 KJ143933 KJ143974 Zhou et al. (2015)
G. ovisporum HKAS123193 (T) China MZ519547 MZ519545 MZ547661 This study
G. ovisporum GACP20071602 China MZ519548 MZ519546 MZ547662 This study
G. perzonatum URM 89437 Brazil MK119830 MK121579 Sun et al. (2020)
G. perzonatum SP445990 Brazil KJ792750 GenBank
G. pfeifferi LGAM 336-ACAM DD2118 Greece MG706232 MG706178 MG837847 MG837806 GenBank
G. pfeifferi Dai 12683 Greece MG279165 MG367560 Xing et al. (2018)
G. philippii Cui 14444 China MG279189 MG367579 MG367525 Xing et al. (2018)
G. philippii MFLU 19-2223 Thailand MN401411 MN398327 MN423175 GenBank
G. podocarpense QCAM6422 Ecuador MF796661 MF796660 GenBank
G. polychromum 330OR USA MG654196 MG754742 Loyd et al. (2018)
G. polychromum UMNOR3 USA MG654204 MG754744 Loyd et al. (2018)
G. ravenelii MS187FL USA MG654211 MG754745 MG754865 Loyd et al. (2018)
G. ravenelii 150FL USA MG654207 Loyd et al. (2018)
G. resinaceum MUCL: 38956 Netherlands MK554772 MK554723 MK554747 Cabarroi-Hernández et al. (2019)
G. resinaceum MUCL: 52253 France MK554786 MK554737 MK554764 Cabarroi-Hernández et al. (2019)
G. ryvardenii GanoTK41 Cameroon JN105699 Kinge et al. (2012)
G. ryvardenii GanoTK43 Cameroon JN105695 Kinge et al. (2012)
G. sandunense GACP18012502 China MK345451 MZ547664 Hapuarachchi et al. (2019b)
G. sandunense GACP18012501 (T) China NR_164049 MZ547663 Hapuarachchi et al. (2019b)
G. sessile 228DC USA MG654319 MG754750 MG754869 Loyd et al. (2018)
G. sessile JV 1209/27 USA KF605630 KJ143937 KJ143976 Zhou et al. (2015)
G. shandongense Dai 15791 China MG279192 MG367582 MG367528 Xing et al. (2018)
G. shandongense Dai 15787 China MG279191 MG367581 MG367527 Xing et al. (2018)
G. shanxiense HSA 539 China MK764269 MK789681 Liu et al. (2019)
G. shanxiense BJTC FM423 (T) China MK764268 MK783937 MK783940 Liu et al. (2019)
G. sichuanense CGMCC 5.2175 (T) China NR_152892 KC662404 Yao et al. (2013)
G. sinense Cui 13835 China MG279193 MG367583 MG367530 Xing et al. (2018)
G. sinense Wei 5327 China KF494998 KF495008 KF494976 MG367529 Xing et al. (2018)
G. steyaertanum 6-WN-16(M)-A Indonesia KJ654461 Glen et al. (2014)
G. steyaertanum V-64-3 Indonesia KJ654433 Glen et al. (2014)
G. stipitatum MUCL: 52655 French Guiana MK554770 MK554717 MK554755 Cabarroi-Hernández et al. (2019)
G. stipitatum MUCL: 43863 Cuba MK554769 MK554739 MK554745 Cabarroi-Hernández et al. (2019)
G. subamboinense UMNFL100 USA MG654373 MG754762 Loyd et al. (2018)
G. subamboinense SPC1 Brazil KU569546 KU570945 Bolaños et al. (2016))
G. tenue GTEN24-1 China DQ424977 GenBank
G. tenue GTEN24-2 China DQ424978 GenBank
G. thailandicum HKAS 104641a Thailand MK848682 MK849880 MK875830 MK875832 Luangharn et al. (2019a)
G. thailandicum HKAS 104640a (T) Thailand MK848681 MK849879 MK875829 MK875831 Luangharn et al. (2019a)
G. tropicum Dai 16434 China MG279194 MG367585 MG367532 Xing et al. (2018)
G. tropicum KUMCC 18–0046a Thailand MH823539 MH883621 Luangharn et al. (2019b)
G. tsugae Cui 14112 China MG279196 MG367587 MG367534 Xing et al. (2018)
G. tsugae Dai 12760 USA KJ143920 KJ143940 KJ143978 Zhou et al. (2015)
G. tuberculosum UMNFL117 USA MG654359 MG754771 Loyd et al. (2018)
G. tuberculosum 233FL USA MG654367 MG754873 Loyd et al. (2018)
G. weberianum CBS 219.36 Philippines MH855780 MH867289 MK611974 MK611972 Cabarroi-Hernández et al. (2019)
G. weberianum CBS 128581 Taiwan MH864975 MH876427 MK636693 MK611971 Cabarroi-Hernández et al. (2019)
G. weixiensis HKAS 100649 (T) China NR_166271 NG_067863 MK302442 Ye et al. (2019)
G. weixiensis HKAS100650 China MK302445 MK302447 MK302443 Ye et al. (2019)
G. wiiroense UMN-21-GHA Ghana KT952363 KT952364 Crous et al. (2015))
G. wiiroense MIN 938704 (T) Ghana NR_158480 NG_064392 Crous et al. (2015)
G. williamsianum Dai 16809 Thailand MG279183 MG367588 MG367535 Xing et al. (2018)
G. williamsianum Wei 5032 China KU219994 KU220024 Song et al. (2016)
G. zonatum FL-03 USA KJ143922 KJ143942 KJ143980 Zhou et al. (2015)
G. zonatum FL-02 USA KJ143921 KJ143941 KJ143979 Zhou et al. (2015)
Foraminispora concentrica Cui 12644 (T) China NR_158325 NG_064396 MK121561 MK121499 Sun et al. (2020)
F. yinggelingensis Cui 13618 (T) China NR_174805 MK119900 MK121570 MK121536 Sun et al. (2020)

Taxon treatments

Ganoderma ovisporum H.D. Yang, T.C. Wen, sp. nov.

Material   Download as CSV 
Holotype:
  1. scientificName:
    Ganoderma ovisporum
    ; kingdom:
    Fungi
    ; phylum:
    Basidiomycota
    ; class:
    Agaricomycetes
    ; order:
    Polyporales
    ; family:
    Polyporaceae
    ; genus:
    Ganoderma
    ; country:
    China
    ; countryCode:
    CN
    ; stateProvince:
    Guizhou
    ; county:
    Sandu Shuizu Autonomous County
    ; locality:
    Sandong Township
    ; verbatimElevation:
    612 m
    ; verbatimLatitude:
    25°70′ N
    ; verbatimLongitude:
    107°96′ E
    ; year:
    2020
    ; month:
    July
    ; day:
    16
    ; habitat:
    Terrestrial
    ; fieldNotes:
    Rotten wood, in dry dipterocarp forest and in upper mixed deciduous forest and growing up from soil
    ; recordedBy:
    Hongde Yang
    ; identifiedBy:
    Hongde Yang
    ; type:
    HKAS123193
    ; collectionID:
    SD2020071601
    ; occurrenceID:
    HKAS123193

Description

Basidiome annual, stipitate, corky, strongly laccate, becoming lighter when dry. Pileus 3 × 5 cm, up to 0.9 cm thick at the base, applanate, subreniform, upper surface red-brown (8E8) when fresh, becoming brownish-black (6C8) when dry, with slightly concentrically sulcate, radially rugose, irregularly tuberculate bumps and ridges overlying the context. Margin is slightly obtuse, yellow-brown (5D8) or concolorous with the pileus. Pore surface pale white (4A2). Pores nearly round to round, 3–4 per mm, dissepiments thin to slightly thick. Context up to 0.3 cm thick, corky, homogeneous at the periphery, becoming three-layered towards the centre, upper layer creamy-white (6E4), middle layer pale brown (6E4), lower layer brown (6D1), without concentric growth zone, black melanoid band absent. There is a line of independent or confluent, laterally arranged tubes inserted between the upper and middle layers of the context. Tubes up to 0.6 cm long, brownish (6E7). Stipe slightly darker than pileus, lateral, subcylindrical, 4-7 cm long, up to 1 cm in diam. Basidia not observed. Basidiospores (12.5–)13.0–13.5–15.0(–15.5) × (9.0–)9.5–10.0–10.5(–11.5) μm (Qm = 1.3, Q = 1.0–1.7,n = 30, with myxosporium), ellipsoid to broadly ellipsoid, ovoid, brown, double-walled, with a dark brown eusporium bearing coarse echinulae and an obtuse turgid appendix, overlaid by a hyaline, smooth myxosporium. Pileipellis hymeniodermiformic, yellowish-brown, terminal cells clavate, entire, brown (5D6), thick-walled, hollow, 18–29 × 6–11 μm. Hyphal system trimitic, generative hyphae 3.5–6 μm in diam., hyaline, colourless, thin-walled with clamp connections; skeletal hyphae 3–6 μm in diam., thick-walled to nearly solid, sometimes branched; binding hyphae 1.5–3 μm in diam., thick-walled, nearly solid, colourless (Fig. 1).

Figure 1.  

Ganoderma ovisporum (HKAS123193, holotype). ab Basidiome; c Pileus; d Pore surface; e Pores; f Sections of pileus; gi Pileipellis terminal cell; j–m Basidiospores; n Skeletal hyphae; o Generative hyphae; p Binding hyphae. Scale bars: g = 50 µm; h–i = 30 µm; j–m = 10 µm; n = 100 µm; o = 10 µm; p = 100 µm.

Etymology

Referring to the ovoid basidiospores.

Notes

Ganoderma ovisporum clusters with G. sandunense in the multigene phylogenetic tree (Fig. 3), the former is similar to the latter by having 98% and 97% homology in ITS and RPB2 sequence data, respectively. These two species are similar in having wide ovoid basidiospores and inhabiting deciduous coniferous mixed forests. However, G. ovisporum differs from G. sandunense in having inconspicuously concentric rings near the pileus margin, lateral stipe and shorter pileipellis terminal cells (18–29 × 6–11 μm), while conspicuously concentric zones and vertically-arranged ridges or grooves, central stipe and longer pileipellis terminal cells (50–95 × 8–13.5 μm) have been observed in the latter. By considering both phylogenetic evidence and morphological observations, we conclude our collection is a new species in Ganoderma.

Ganoderma sandunense Hapuar., T.C. Wen & K.D. Hyde

Material   Download as CSV 
Holotype:
  1. scientificName:
    Ganoderma sandunense
    ; kingdom:
    Fungi
    ; phylum:
    Basidiomycota
    ; class:
    Agaricomycetes
    ; order:
    Polyporales
    ; family:
    Polyporaceae
    ; genus:
    Ganoderma
    ; country:
    China
    ; countryCode:
    CN
    ; stateProvince:
    Guizhou
    ; county:
    Sandu Shuizu Autonomous County
    ; verbatimElevation:
    590 m
    ; verbatimLatitude:
    24°54′N
    ; verbatimLongitude:
    107°53′E
    ; year:
    2018
    ; month:
    January
    ; day:
    25
    ; habitat:
    Terrestrial
    ; fieldNotes:
    Rotten wood, growing up from the soil
    ; recordedBy:
    Ting-Chi Wen
    ; identifiedBy:
    Kalani Hapuarachchi
    ; type:
    GACP18012501
    ; collectionID:
    GACP18012501
    ; occurrenceID:
    GACP18012501

Description

Basidiome annual, stipitate, corky, strongly laccate. Pileus hemispherical, projecting 8 cm, up to 4 cm wide and 1.5 cm thick. Pileal surface reddish-black (8E8) to brownish-black (6C8), with distinctly concentrically sulcate, vertically-arranged ridges or grooves. Margin obtuse, concolorous with the pileus. Pore surface whitish-yellow (4A2) to light brown (6D4). Pores nearly circular, 3–5 per mm, dissepiments thin. Context up to 0.5 cm thick, inconspicuous triplex, fawn (5C5) to creamy-white (5A1) to dark brown (5E6), without concentric growth zone, black melanoid band absent. There is a line of independent or confluent, laterally-arranged tubes inserted between the upper and middle layer of the context. Tubes up to 1.2 cm long, dark brown (7F8). Stipe slightly darker than pileus, central, subcylindrical, up to 8 cm, 0.5 cm in diam. Basidia broadly ellipsoid, 21–25.5 × 13.5–17.5 μm, with four sterigmata. Basidiospores (12.3–)13.2–13.7–14.2(–15.7) × (9.0–)10–10.3–10.6(–12.5) μm (Qm = 1.3, Q = 1.0–1.7, n = 30, with myxosporium), ellipsoid to broadly ellipsoid, brown (7E5). Pileipellis cells clavate like, entire, brownish-orange (5C5), 50–95 × 8–13.5 μm. Hyphal system trimitic, generative hyphae 4-6 μm in diam., hyaline, colourless, thin-walled with clamp connections; skeletal hyphae 3.5–6 μm in diam., thick-walled to nearly solid, sometimes branched; binding hyphae 1-2 μm in diam., thick-walled, nearly solid, colourless (Fig. 2).

Figure 2.  

Ganoderma sandunense (GACP18012501, holotype). a Basidiome; b Pore surface; c Sections of pileus; de Pileipellis terminal cell; f Basidia; gj Basidiospores; k Skeletal hyphae and binding hyphae; l Generative hyphae; m Binding hyphae. Scale bars: d–e= 50 μm; f–h = 20 μm; i–j = 10 μm; k = 100 μm; l–m = 50 μm.

Notes

Ganoderma sandunense was introduced by Hapuarachchi et al. (2019b) with ITS sequence. In addition, the description of its basidia is absent in their publication. In this study, the holotype of G. sandunense was loaned from Herbarium (GACP) and re-examined. We have refined this species with a more detailed illustration. Furthermore, we provided RPB2 sequence data of this species, which is an important phylogenetic marker used for intraspecific delimitation within Ganoderma.

Identification keys

Keys to 22 species of laccate Ganoderma species in China

1 Distributed in China with gymnosperms as substrates G. tsugae
Distributed in China with angiosperms as substrates 2
2 Basidiome sessile 3
Basidiome stipitate to substipitate 5
3 Pileipellis terminal cells regular, clavate, occasionally with blunt outgrowth and protuberance, context present melanoid bands, basidiospores 8–12 × 3.8–5.2 µm G. angustisporum
Pileipellis terminal cells are irregular, mainly composed of clavate cells or branched cells with blunt outgrowths in the lateral part or protuberances in the apical 4
4 Melanoid bands absent in the context, concentric growth zones present in the context, basidiospores 9.2–12 × 6.8–8.4 μm G. mutabile
Melanoid bands present in the context, concentric growth zones absent in the context, basidiospores 8–13.5 × 4.2–6.3 μm G. boninense
5 Distributed in tropical regions 6
Distributed mainly in temperate regions 8
6 Basidiome notably with a long, lateral stipe, pileus smaller, basidiospores with coarsely echinulate, 8.5–11 × 5–7 μm G. flexipes
Basidiome stipitate to substipitate, pileus dimidiate, mostly large 7
7 Pileus single or occasionally composed of many small pilei, concentric growth zones present in the context, basidiospores with fine and long echinulate, 8–11.3 × 5–12.8 μm G. multipileum
Pileus is mostly single, concentric growth zones absent in the context, basidiospores with coarse and short echinulae, 8.5–12.5 × 5.5–7.5 μm G. orbiforme
8 Context nearly homogeneous to homogeneous 9
Context duplex to triplex 12
9 Pileus context white, pore surface white to cream, basidiospores 9.5–12.5 × 7–9 μm G. leucocontextum
Pileus context brownish to brown or darker 10
10 Pileipellis terminal cells are mostly irregular, context present melanoid bands and concentric growth zones, basidiospores 10.8–13.1 × 8.3–11 μm G. tropicum
Pileipellis terminal cells regular, cylindrical to clavate, context absent melanoid bands 11
11 Inhabiting deciduous forests, basidiospores ellipsoid, normally with an orderly arranged echinulae, basidiospores 10.7–12.8 × 7.0–9.0 μm G. sinense
Inhabiting bamboo forests 12
12 Pileipellis terminal cells 35–65 × 8–16 μm, basidiospores 11–12.5 × 6.5–7.5 μm G. bambusicola
Pileipellis terminal cells 20–55 × 10–15 μm, basidiospores 8.0–12.5 × 5.0–8.0 μm G. esculentum
13 Chlamydospores present in the context, basidiospores 7.8–10.4 × 5.2–6.4 μm G. weberianum
Chlamydospores absent in context 14
14 Basidiospores < 8 μm in width and < 12 μm in length 15
Basidiospores > 8 μm in width and > 9 μm in length 17
15 Basidiome corky, context soft, pores 2–4 per mm, pileipellis terminal cells regular, clavate, 20‒35 × 10‒12 μm, basidiospores 5.7‒8.3 × 2.6‒4.6 μm G. weixiensis
Basidiome corky to woody, context firm, pores 4–6 per mm, pileipellis terminal cells occasional with outgrowths 16
16 Growing on living trees of Casuarina equisetifolia, pileipellis terminal cells 40–70 × 5–13 µm, basidiospores 8.3–11.5 × 4.5–7 µm G. casuarinicola
Growing on deciduous trees, pileipellis terminal cells 20–40 × 7–15 μm, basidiospores 7–9.3 × 4.6–6.8 μm G. lingzhi
17 Basidiospores ellipsoid, with sinuous ridge-like echinulae, 12.3–13.8 × 8.5–9.8 μm G. lucidum
Basidiospores broadly ellipsoid, with coarse echinulae and an obtuse turgid appendix 18
18 Context brown to dark brown 19
Context greyish-white to fawn brown 20
19 Pores 4–5 per mm, pileipellis terminal cells 25–30 × 7.5–8.5 μm, basidiospores 11.0–13.0 × 8.0–9.5 μm G. shanxiense
Pores 5–8 per mm, pileipellis terminal cells 20–45 × 5.5–7.5 μm, basidiospores 9.0–12.5 × 6.5–9.0 μm G. dianzhongense
20 Distributed in Shandong Province, pileipellis terminal cells 17–25 × 4.5–7.5 μm, basidiospores 9‒13 × 6‒9 μm G. shandongense
Distributed in Guizhou Province 21
21 Basidiome with a central stipe, pileipellis terminal cells 50–95 × 8–13.5 μm, basidiospores 12.3–15.7 × 9.1–12.0 μm G. sandunense
Basidiome with a lateral stipe, pileipellis terminal cells 18–29 × 6–11 μm, basidiospores 12.5–15.5 × 9.0–11.5 μm G. ovisporum

Analysis

Phylogenetic analyses

Eight sequences of ITS, LSU and RPB2 were successfully amplified, but we failed to obtain the TEF1-α sequence from the two specimens HKAS123193 and GACP20071602. The newly-generated sequences and sequences from GenBank represented 132 specimens from 66 species, of which 21 were the type. The combined alignment of sequences comprised 3028 characters of 606, 1020, 809, 593 belonging to TEF1-α, RPB2, ITS and LSU, respectively. The final ML optimisation log-likelihood was -17354.28. The Bayesian Inference stopped at 2915000 generations when the average standard deviation of split frequencies reached 0.009904. The tree topologies derived from ML and BY were identical. Therefore, only the ML tree is shown (Fig. 3). The new species G. ovisporum and G. sandunense formed an individual clade in the phylogenetic tree (Fig. 3).

Figure 3.  

Phylogram for Ganoderma generated from Maximum Likelihood analysis of ITS, LSU, TEF1-α and RPB2 sequence data. Bootstrap support values for Maximum Likelihood and maximum parsimony greater than 70% and posterior probabilities of Bayesian Inference ≥ 0.95 are given above branches. Type specimens are marked with letter (T) and new species in this study are indicated in red.

Discussion

In this study, both phylogeny and morphology support G. ovisporum as a new species. Morphologically, it resembles other dark-coloured, laccate, stipitate Ganoderma species. However, it can be distinguished by having larger (12.5–15.5 × 9.0–11.5 μm), wide ovoid, dark brown-pigmented basidiospores. It is mostly similar to G. sandunense in having brownish-black pileus and similarly-sized basidiospores, as well distribution in Guizhou Province (Hapuarachchi et al. 2019b). The former species is distinct from the latter by having a lateral stipe and shorter pileipellis terminal cells (18–29 × 6–11 μm). Phylogenetically, G. ovisporum and G. sandunense are closely related, forming a distinct clade with basal position with strong support.

Ganoderma was extensively researched by the Chinese because it applied to medicine and food, together with the symbolic happiness and immortality culture, those being recognised as long as 2,000 years ago (Hapuarachchi et al. 2016b, Hapuarachchi et al. 2018b, Li et al. 2018, Cui et al. 2019, Du et al. 2021). Chinese taxonomists emphasised the morphological characteristics, such as stipe, pileus, pores, context, pileipellis terminal cells and basidiospores as keys to identity (Zhao 1989, He and Yu 1989, Zhang 1997). Keeping this method, Zhao and Zhang (2000) recorded 76 Ganoderma species from China, providing detailed illustrations. Wu and Dai (2005) identified 77 Ganoderma species with full description and colour photographs. Studies have been implemented to revise the taxonomy of Ganoderma in China by using molecular and morphology methods in the recent decade. The results indicated at least 23 species names are synonyms and confirmed that 24 species are distributed in China, 16 of which possess laccate basidiomes (Wang 2012, Chao 2013, Xing 2019). Since then, six species with laccate basidiomes have been described from China: G. bambusicola, G. dianzhongense, G. esculentum, G. sandunense, G. shanxiense and G. weixinense (Hapuarachchi et al. 2019b, Liu et al. 2019, Ye et al. 2019, Wu et al. 2020, He et al. 2021). Ganoderma taxonomy has undergone tremendous changes since both phenotypic features and phylogeny were used to delineate species (Gottlieb et al. 2000, Hapuarachchi et al. 2018c, Hapuarachchi et al. 2018a, Lin and Yang 2019, Tchotet-Tchoumi et al. 2019, He et al. 2022). Based on the aforementioned characteristics, we have provided a dichotomous key to 22 laccate species, including our new species from China.

Ganoderma could originate from Southeast Asia and later dispersal to the Northern Hemispheres, the Southern Hemispheres and the neotropics before 30 Mya years, during which species radiation and diversification events happened (Moncalvo and Buchanan 2008). Overviewing Ganoderma species worldwide, Imazeki (1939) concluded using subgenera Euganoderma and Elfvingia to accommodate species with laccate and non-laccate characters, respectively. In this study, a phylogenetic analysis was carried out using combined LSU, ITS, TEF1-α and RPB2 sequences from 66 species that included species previously placed in the above two subgenera. The topology of our phylogenetic tree is consistent with the morphology that the laccate species and non-laccate species tend to form groups. It is worth mentioning that the new species G. ovisporum group with the laccate species of G. carnosum, G. dianzhongense, G. leucocontextum, G. lucidum, G. oregonense, G. sandunense, G. shandongense, G. shanxiense, G. tsugae and G. weixiensis had strong support in both ML and Bayesian analyses. Those species were found in only or few ecological niches, except the widely cultivated G. leucocontextum, G. lucidum and G. tsugae (Gottlieb et al. 2000, Moncalvo and Buchanan 2008, Hapuarachchi et al. 2018b, Lin and Yang 2019). Therefore, many Ganoderma species are geographically restricted (He et al. 2022). However, the phylogenetic tree in the case of the laccate species G. pfeifferi and G. mutabile grouped with the non-laccate species G. adspersum, G. australe, G. eickeri, G. ellipsoideum, G. gibbosum, G. knysnamense, G. lobatum, G. podocarpense and G. williamsianum, indicating Euganoderma and Elfvingia are polyphyletic (Gottlieb et al. 2000). However, in fact, they are similar in having a substipitate to sessile basidiome and living as saprobes or parasites (Hapuarachchi et al. 2018c, Tchotet-Tchoumi et al. 2019). Consequently, biogeographic patterns and convergent evolution could explain the population structure and evolution of Ganoderma. Thus, a phylogeography study would help better understand the evolution of Ganoderma.

Acknowledgements

This work was financed by the Science and Technology Foundation of Guizhou Province (KY [2018]039 and No. [2019]2451-3) and by the Open Fund Project of Key Laboratory of Forest Biotechnology in Yunnan, Southwest Forestry University, China (51700201). The authors are very grateful to Professor Xing-Liang Wu for his valuable comments and suggestions.

References

login to comment