Research Article
Print
Research Article
Two new species of Ganoderma (Ganodermataceae, Basidiomycota) from Southwest China
expand article infoJun He, Xiao-Jun Li, Wan-Zhong Tan, Xiao-Qu Wu§|, Dan Wu, Zong-Long Luo, Qi Wu Zhou, E-Xian Li§, Shu-Hong Li§
‡ West Yunnan University, Lincang, China
§ Biotechnology and Germplasm Resources Institute, Yunnan Academy of Agricultural Sciences, Kunming, China
| Yunan University, Kunming, China
¶ Dali University, Dali, China
Open Access

Abstract

Ganoderma is a large and diverse genus containing fungi that cause white rot to infect a number of plant families. This study describes G. phyllanthicola and G. suae as new species from Southwest China, based on morphological and molecular evidence. Ganoderma phyllanthicola is characterized by dark brown to purplish black pileus surface with dense concentric furrows, pale yellow margin, irregular pileipellis cells, small pores (5–7 per mm) and ellipsoid to sub-globose basidiospores (8.5–10.0 × 6.0–7.5 µm). Ganoderma suae is characterized by reddish brown to oxblood red pileus surface and lead gray to greyish-white pore surface, heterogeneous context, wavy margin and almond-shaped to narrow ellipsoid basidiospores (8.0–10.5 × 5.0–7.0 μm). The phylogeny of Ganoderma is reconstructed with multi-gene sequences: the internal transcribed spacer region (ITS), the large subunit (nrLSU), translation elongation factor 1-α gene (TEF-1α) and the second subunit of RNA polymerase II (RPB2). The results show that G. suae and G. phyllanthicola formed two distinct line-ages within Ganoderma. Descriptions, illustrations and phylogenetic analyses results of the two new species are presented.

Key words

2 new taxa, Ganodermataceae, Morphology, Phylogeny, Taxonomy

Introduction

Ganodermataceae is one of the main families of polypores with fourteen accepted genera: Amauroderma Murrill, Amaurodermellus Costa-Rezende, Drechsler-Santos & Góes-Neto, Cristataspora Robledo & Costa-Rezende, Foraminispora Robledo, Costa-Rez. & Drechsler-Santos, Furtadoella B.K. Cui & Y.F. Sun, Ganoderma P. Karst., Haddowia Steyaert, Humphreya Steyaert, Magoderna Steyaert, Neoganoderma B.K. Cui & Y.F. Sun, Sanguinoderma Y.F. Sun, D.H. Costa & B.K. Cui, Sinoganoderma B.K. Cui, J.H. Xing & Y.F. Sun, Tomophagus Murrill and Trachydermella B.K. Cui & Y.F. Sun (Costa-Rezende et al. 2020; Sun et al. 2022), of which most species are classified in the genus Ganoderma.

The word Ganoderma is derived from the Greek words “Gano”, meaning “shiny”, and “derma”, meaning “skin” (Loyd et al. 2018). The genus Ganoderma (Polyporales, Basidiomycota) was described by Fries (1821) based on Polyporus lucidus (Curtis) Fr. and typified by Ganoderma lucidum (Curtis) P. Karst. from Europe (Fries 1821; Karsten 1881). Ganoderma is a globally distributed genus of wood-decaying fungi that encompass important species for forestry, medicine, food, and cultural traditions, in which morphological delimitation has been challenging due to its large plasticity and wide distribution across various regions (Luangharn et al. 2021). In the past few decades, DNA or amino acid sequence analyses have provided effective tools for taxonomists to combine data. These modern techniques have helped to clarify the distribution of different species complexes in the genus Ganoderma, and have revealed some instances of misidentification (Kinge et al. 2015; Fryssouli et al. 2020).

The genus is characterized by laccate or non-laccate basidiocarps, sessile to stipitate basidiomata, white to pale yellow margin, and red-brown colored truncate double-walled basidiospores, an apical germinal pore, thin and colourless external wall (exosporium), with a brown to dark brown interwall pillars (endosporium), and the ability to cause white rot in woody plants (Karsten 1881; Moncalvo and Ryvarden 1997). Furthermore, these species hold different characteristics, such as the shape and the color of the fruit body, host specificity, and geographical origin, which are used to identify individual members of the species. The species concept in the genus Ganoderma is thus not universally accepted nor well established due to the highly variable morphological features of the species (Wang et al. 2014; Náplavová et al. 2020).

Currently, based on credible morphological and phylogenetic evidence, 191 species of Ganoderma have been described worldwide (He et al. 2022; Sun et al. 2022; Vinjusha et al. 2022; Cabarroi-Hernández et al. 2023). Ganoderma is economically important, due to the fact that members of the genus are regarded as valuable medicinal mushrooms (Hapuarachchi et al. 2018a). Several Ganoderma species are known to be prolific sources of a high number of natural bioactive compounds such as polysaccharides, triterpenoids, sterols, and secondary metabolites (Richter et al. 2015). Approximately 45 species of Ganoderma are recorded in Chinese Fungi (Sun et al. 2022), of which Ganoderma lucidumlingzhi” and G. sinense which used to be listed in Chinese Pharmacopeia to prevent and treat many diseases and are listed in Chinese Pharmacopeia, and which are included in the homologous list of medicine and food (Li et al. 2018). Furthermore, Ganoderma was included in the American Herbal Pharmacopoeia and Therapeutic Compendium (Hapuarachchi et al. 2018b). They are commonly named as “Lingzhi” or “Rui–zhi” in China, “Youngzhi” in Korea, “Reishi” in Japan and “Ganoderma” in the USA (Liu et al. 2015). These natural bioactive compounds are used to treat and remedy many pathological diseases, including traditional medicine for treating neurasthenia, debility of prolonged illness, insomnia, arthritis, asthma, anorexia, dizziness, chronic hepatitis, hypercholesterolemia, mushroom poisoning, coronary heart disease, hypertension, prevention of acute mountain sickness, deficiency fatigue’, carcinoma, and bronchial cough in the elderly (Wang et al. 2020). In addition, Ganoderma products come in the form of various commercial products of Ganoderma such as powders, dietary supplements, coffee, tea, spore products, drinks, syrup, toothpaste, soap, lotion, and capsules, and have been commercialized as effective food and drug supplements for health benefits (Lai et al. 2004).

Southwest China contains some of the highest concentrations of fungal biodiversity in the world, and Yunnan Province, in particular, has a varied topography, environmental conditions, and a variety of habitats for a diverse range of fungi (He et al. 2022). Despite the advancement in taxonomic studies of Ganoderma species diversity, many novel species are still being discovered (He et al. 2021; He et al. 2022). During our investigations of macrofungi in Southwest China, a couple of specimens of Ganoderma were collected. In the current study, the phylogenetic analyses of Ganoderma were carried out based on the combined sequence dataset of ITS + nLSU + TEF1–α + RPB2 gene regions. Subsequent morphological and molecular studies uncovered two undescribed species. These species are illustrated and described below.

Materials and methods

Specimen collection

During the rainy season from June 2019 to September 2023, four Ganoderma specimens were collected in southwest China. They were photographed in the field, then macro-mophology was described on fresh basidiomata, on the same day of collection. Specimens were there after thoroughly dried at 45 °C (Hu et al. 2022), in a thermostatic drier, stored in sealed plastic bags, and deposited in the herbarium of Kunming Institute of Botany, Chinese Academy of Sciences Academia Sinica (KUN-HKAS).

Morphological studies

Colour codes were determined following Kornerup and Wanscher (1978). For microscopic characteristics, anatomical and cytological characteristics including basidia, basidiospores, hyphal system, and pileipellis were observed and photographed using a Nikon ECLIPSE Ni-U microscope (Nikon, Japan) at magnifications up to × 1000. Tarosoft(R) Image Frame Work (IFW) was used for the measurement of photomicrographs, and Adobe Photoshop CS5 software was used to process images for making photo plates (He et al. 2021).

The following abbreviations are used: IKI = Melzer’s reagent, IKI– = neither amyloid nor dextrinoid, KOH = 10% potassium hydroxide, CB = Cotton Blue, CB+ = cyanophilous. The notation [n/m/p] specifies that measurements were made on “n” basidiospores from “m” basidiomata and “p” collections. Basidiospore dimensions are given as (a) b–av– c (d). Where a and d refer to the lower and upper extremes of all measurements, respectively, b-c the range of 95% of the measured values, L = mean spore length (arithmetic average of all spores), W = mean spore width (arithmetic average of all spores), Q is the length/width ratio of basidiospores, Qm denotes the average of n measured basidiospores and SD is their standard deviation. Results are presented as Q = Qm ± SD.

DNA extraction, PCR amplification, and sequencing

Genomic DNA was extracted from dry specimens using the Ezup Column Fungi Genomic DNA Purification Kit following manufacturer instructions. Primers pairs for PCR were respectively ITS1F/ITS5 (White et al. 1990), LR5/LR0R (Liu et al. 1999), TEF1–983 / TEF1–1567R (Matheny et al. 2007), and RPB2–6f / fRPB2–7cR (Liu et al. 1999), respectively. Primer sequences are available in the WASABI database at the AFTOL website (aftol.org). The PCR mixture was prepared in a 30 μL final volume, with 15 μL 2× Taq Plus Master Mix II (Sangon Biotechnology Co., Kunming, China), 12 μL ddH2O, 0.5 μL 10 μM of forward and reverse primers, 2 μL DNA. The PCR thermal cycle program for ITS and nrLSU amplification was conducted using the following profiles: 94 °C for 5 min, 35 cycles of 94 °C for 30 s, 53 °C for 50 s, 72 °C for 1 min, and 72 °C for 10 min. The PCR cycling for TEF1-α was as follows: initial denaturation at 94 °C for 5 min, followed by 35 cycles at 94 °C for 30 s, 55 °C for 30 sec and 72 °C for 1 min, and 72 °C for 10 min. The PCR cycling for RPB2 was as follows: initial denaturation at 94 °C for 5 min, followed by 35 cycles at 94 °C for 30 s, 50 °C for 50 s and 72 °C for 1 min, and 72 °C for 10 min. PCR products were checked on 1% agarose gels stained with ethidium bromide under UV light. The PCR products were purified and sequenced by the Sangon Biotech Limited Company (Shanghai, China). Raw DNA sequences were assembled and edited in Sequencher 4.1.4, and the assembled DNA sequences were deposited in GenBank (Table 1).

Table 1.

Names, voucher numbers, origins, and their corresponding GenBank accession numbers of the taxa used in the phylogenetic analyses. The new species sequences generated sequences is show in bold, after the species name and the type specimens show “T” after the number.

Species Voucher/strain Origin GenBank accession numbers
ITS nLSU TEF1–α RPB2
Ganoderma acaciicola Cui16815T Australia MZ354895 MZ355005 MZ245384
G_ acaciicola Cui16813 Australia MZ354893 MZ355003 MZ245382
G. artocarpicola HL173T Yunnan, China ON994239 OP456495 OP508442 OP508428
G. artocarpicola HL188 Yunnan, China ON994240 OP380253 OP508441 OP508427
G. aridicola Dai12588T South Africa KU572491 KU572502
G. austroafricanum CBS138724T South Africa KM507324 KM507325 MK611970
G. austroafricanum CMW25884 South Africa MH571693 MH567296
G. boninense WD2085 Japan KJ143906 KJ143925 KJ143965
G. boninense WD2028 Japan KJ143905 KU220015 KJ143924 KJ143964
G. bubalinomarginatum Dai20075 T Guangxi, China MZ354926 MZ355010 MZ221637 MZ245388
G. bubalinomarginatum Dai20074 Guangxi, China MZ354927 MZ355040 MZ221638 MZ245389
G. carocalcareum DMC513 Cameroon EU089970
G. carocalcareum DMC322 T Cameroon EU089969
G. casuarinicola HKAS104639 Thailand MK817650 MK817654 MK871328 MK840868
G. casuarinicola Dai16336 T Guangdong, China MG279173 MG367565 MG367508
G. concinnum Robledo3235 Brazil MN077523 MN077557
G. concinnum Robledo3192 Brazil MN077522 MN077556
G. curtisii CBS100132 NC, USA JQ781849 KJ143927 KJ143967
G. curtisii CBS100131 NC, USA JQ781848 KJ143926 KJ143966
G. destructans CBS139793 T South Africa NR132919 NG058157
G. destructans Dai16431 South Africa MG279177 MG367569 MG367512
G. dunense CMW42150 South Africa MG020249 MG020228
G. dunense CMW42157 T South Africa MG020255 MG020227
G. ecuadorense URM89449 Ecuador MK119828 MK119908 MK121577 MK121535
G. ecuadorense URM89441 Ecuador MK119827 MK119907 MK121576 MK121534
G. enigmaticum Dai15971 Africa KU572487 KU572497 MG367514
G. enigmaticum Dai15970 Africa KU572486 KU572496 MG367513
G. heohnelianum Cui13982 Guangxi, China MG279178 MG367570 MG367515
G. heohnelianum Dai11995 Yunnan, China KU219988 KU220016 MG367550 MG367497
G. hochiminhense MFLU19_2225 Vietnam MN396662 MN396391 MN423177
G. hochiminhense MFLU19_2224 T Vietnam MN398324 MN396390 MN423176
G. lingzhi Dai20895 Liaoning, China MZ354904 MZ355006 MZ221668 MZ245413
G. lingzhi HL56 Yunnan, China ON994247 OP380262 OP508423
G. martinicense 246TX TX, USA MG654185 MG754737 MG754858
G. martinicense LIPSWMart0855 T Martinique, France KF963256
G. suae L4651 T Yunnan, China PP869243 PP869250 PP894782 PP894784
G. suae L4817 Yunnan, China PP869244 PP869251 PP894783
G. mexicanum MUCL55832 Martinique MK531815 MK531829 MK531839
G. mexicanum MUCL49453 Martinique MK531811 MK531825 MK531836
G. mirabile Cui18271 Malaysia MZ354958 MZ355067 MZ221672 MZ345729
G. mirabile Cui18283 Malaysia MZ354959 MZ355069 MZ221673 MZ345730
G. mizoramense UMNMZ5 India KY643751 KY747490
G. mizoramense UMNMZ4T India KY643750
G. multipileum Cui13597 Hainan, China MZ354899 MZ355043 MZ221675 MZ345732
G. multipileum L4989 Yunnan, China ON994249 OP380264 OP508447 OP508432
G. multiplicatum CC8 China KU569515 KU570915
G. multiplicatum Dai17395 Brazil MZ354903 MZ221678 MZ345734
G. multiplicatum SPC9 Brazil KU569553 KU570951
G. multiplicatum URM83346 Brazil JX310823 JX310837
G. myanmarense MFLU19_2167 T Myanmar MN396330 MN428672
G. myanmarense MFLU19_2169 Myanmar MN396329 MN398325
G. nasalanense GACP17060211 T Laos MK345441 MK346831
G. nasalanense GACP17060212 Laos MK345442 MK346832
G. orbiforme HL43 Yunnan, China ON994250 OP380265 OP508435
G. orbiforme TNM F0018838 China JX840350
G. parvulum MUCL52655 Guiana, French MK554770 MK554717 MK554755
G. parvulum MUCL47096 Cuba MK554783 MK554721 MK554742
G. philippii Cui14443 Hainan, China MG279188 MG367578 MG367524
G. philippii MFLU19/2222 Thailand MN401410 MN398326 MN423174
G. polychromum 330OR OR, USA MG654196 MG754742
G. polychromum MS343OR OR, USA MG654197 MG754743
G. ravenelii MS187FL FL, USA MG654211 MG754745 MG754865
G. ravenelii NC_8349 USA AY456341
G. resinaceum LGAM462 Greece MG706250 MG706196 MG837858 MG837821
G. resinaceum LGAM448 Greece MG706249 MG706195 MG837857 MG837820
G. resinaceum MUCL38956 Netherlands MK554772 MK554723 MK554747
G. resinaceum MUCL52253 France MK554786 MK554737 MK554764
G. rodriguezii M–11926 Cuba OQ079179
G. rodriguezii 269TX USA MG654352
G. ryvardenii HKAS58053 T South Africa HM138670
G. ryvardenii HKAS58054 South Africa HM138671
G. sessile 113FL FL, USA MG654307 MG754748 MG754867
G. sessile 111TX TX, USA MG654306 MG754747 MG754866
G. sichuanense Cui16343 China MZ354928 MZ355011 MZ221692 MZ345741
G. sichuanense Dai19651 Sri Lanka MZ354929 MZ355031 MZ221693 MZ345742
G. sinense Wei5327 Hainan, China KF494998 KF495008 KF494976 MG367529
G. sinense HL109 Yunnan, China ON994252 OP380267 OP508438 OP508425
G. steyaertanum MEL2382783 Australia KP012964
G. steyaertanum 6WN 20B Indonesia KJ654462
G. phyllanthicola L4948 T Yunnan, China PP869245 PP869252
G. phyllanthicola HL308 Yunnan, China PP869246 PP869253
G. thailandicum HKAS104640 T Thailand MK848681 MK849879 MK875829 MK875831
G. thailandicum HKAS104641 Thailand MK848682 MK849880 MK875830 MK875832
G. tropicum Dai16434 Hainan, China MG279194 MZ355026 MG367585 MG367532
G. tropicum HL186 Yunna, China ON994253 OP380268 OP508440
G. tuberculosum GVL40 Veracruz, Mexico MT232634
G. tuberculosum JV1607_62 Costa Rica MZ354944 MZ355087 MZ221710
G. weberianum CBS21936 Philippines MK603804 MK611974 MK611972
G. weberianum Dai19673 China MZ354930 MZ355032 MZ221712 MZ358829
G. wiiroense UMN21GHA T Ghana KT952363 KT952364
G. wiiroense UMN20GHA Ghana KT952361 KT952362
G. zonatum FL03 FL_USA KJ143922 KJ143942 KJ143980
G. zonatum FL02 FL_USA KJ143921 KJ143941 KJ143979
Amauroderma rugosum Cui9011 Guangdong, China KJ531664 KU572504 MG367506

Sequencing and sequence alignment

Sequences newly generated in this study and sequences obtained from GenBank (Table 1) were analyzed. The related sequences were determined by using a BLAST search to reveal the closest matches with taxa in Ganoderma and recent relevant publications (Sun et al. 2022). Sequences were aligned using MAFFT v.7 (http://mafft.cbrc.jp/alignment/server/) (Katoh and Standley 2013) and then checked visually and manually optimized using BioEdit v.7.0.9 (Hall 1999), to allow maximum alignment and minimize gaps. Ambiguous regions were excluded from the analyses and gaps were treated as missing data. The phylogeny website tool “ALTER” (Glez-Peña et al. 2010) was used to convert the alignment fasta file to Phylip format for RAxML analysis and AliView and PAUP 4.0 b 10 were used to convert the alignment fasta file to a Nexus file for Bayesian analysis (Swofford 2003).

Phylogenetic analyses

A maximum likelihood (ML) analysis was performed at the CIPRES web portal (Miller et al. 2010) using RAxML v.8.2.12 as part of the “RAxML-HPC2 on TG” tool (Miller et al. 2010). A general time-reversible model (GTR) was applied with a discrete gamma distribution and four rate classes. Fifty thorough ML tree searches were conducted out in RAxML v.8.2.11 under the same model. One thousand non-parametric bootstrap iterations were run with the GTR model and a discrete gamma distribution. The resulting replicates were plotted onto the best scoring tree obtained previously. Since no supported conflict (BS ≥ 60%) was detected among the topologies, the four single-gene alignments were concatenated using SequenceMatrix (Vaidya et al. 2011).

The Bayesian analyses were performed using PAUP v.4.0b10 and MrBayes v.3.2 (Ronquist et al. 2012), and the best-fit model of sequences evolution was estimated via MrModeltest 2.3 (Guindon and Gascuel 2003; Nylander 2004; Darriba et al. 2012). Markov Chain Monte Carlo (MCMC) sampling approach was used to calculate posterior probabilities (PP) (Rannala and Yang 1996). Bayesian analyses of six simultaneous Markov chains were run for one million generations and trees were sampled every 100th generation with a total of 10,000 trees. The first 2000 trees were discarded and the remaining trees were used for calculating posterior probabilities in the majority rule consensus tree.

Phylogenetic trees were visualized using FigTree v1.4.4 (http://tree.bio.ed.ac.uk/software/figtree/), and editing and typesetting was done using Adobe Illustrator CS5 (Adobe Systems Inc., USA). Sequences derived in this study were deposited in GenBank (http://www.ncbi.nlm.nih.gov). The final sequence alignments and the phylogenetic trees are available at TreeBase (http://www.treebase.org, accession number: 31439).

Results

Phylogenetic analyses

In this study, eleven sequences were newly generated from specimens of Ganoderma spp. and deposited in GenBank (Table 1), all collected from Yunnan Province, China. The dataset comprised combined ITS + nrLSU + TEF1-α + RPB2 sequences data from 94 specimens, representing 46 taxa in Ganodermataceae. The aligned dataset comprised 2633 characters including gaps (ITS: 1–576; nrLSU: 577–1423; TEF1-α: 1424–1959; RPB2: 1969–2633) of which Amauroderma rugosum Cui 9011 as the outgroup taxon (Fig. 1, Sun et al. 2020). The likelihood of the final tree was evaluated and optimized under GAMMA. The best RAxML tree with a final likelihood value of -13209.788540 is presented (Fig. 1). The matrix had 855 distinct alignment patterns, with 37.25% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.223857, C = 0.250309, G = 0.275931, T = 0.249903; substitution rates AC = 1.104623, AG = 5.821128, AT = 1.180576, CG = 1.273064, CT = 8.821984, GT = 1.000000, α = 0.169579, Tree-Length: 1.075203. Best model for the ITS + nLSU + TEF1-α + RPB2 dataset estimated and applied in the Bayesian analysis were HKY+I+G for ITS and RPB2 (Lset nst=2, rates=invgamma; Prset statefreqpr=dirichlet (1,1,1,1)), GTR+I+G for nrLSU and TEF1-α (Lset nst=6, rates=invgamma; Prset statefreqpr=dirichlet (1,1,1,1). ML analysis resulted in a similar and applied in the Bayesian in equal frequency of nucleotides. Bootstrap support values with a maximum likelihood (ML) equal to or greater than 60%, and Bayesian posterior probabilities (PP) equal to or greater than 0.90 are given above the nodes (Fig. 1).

Figure 1. 

Maximum likelihood (ML) tree based on combined ITS + nrLSU + TEF1-α + RPB2 sequence data. Bootstrap support values with a maximum likelihood (ML) equal to or greater than 60% and Bayesian posterior probabilities (PP) equal to or greater than 0.90 are given above the nodes, shown as “ML/PP”. New species are indicated in bold blue.

The phylogeny demonstrated that our four Ganoderma-like specimens were clustered into two different lineages with high support, represented two new species, G. phyllanthicola (100% BS and 1.00 BP; Fig. 1) and G. suae (100% BS and 1.00 BPP; Fig. 1). Ganoderma phyllanthicola sp. nov. clustered as a sister clade with G. castaneum BK Cui, JH Xing & YF Sun G. tropicum (Jungh.) Bres. and G. philippii Bres. & Henn. ex Sacc. with strong statistical support (90%ML/0.99PP, Fig. 1), but forming a distinct lineage. Ganoderma suae sp. nov. was sister to G. resinaceum Boud. with high statistical support (100%ML/1.00PP, Fig. 1).

Taxonomy

Ganoderma phyllanthicola J. He & S.H. Li, sp. nov.

MycoBank No: 853508
Fig. 2

Diagnosis

Differs from other species in the genus by its sessile and coriaceous basidiomata, dark brown to purplish black pileus surface with dense concentric furrows, pale yellow margin, irregular pileipellis cells, broadly ellipsoid to subglobose basidiospores and truncated apex, exospore walls smooth, endospore walls with dense spinules.

Etymology

The epithet ‘phyllanthicola’ refers to the host tree genus Phyllanthus.

Holotype

China. Yunnan Province., Honghe City, Mengzi County, on living tree of Phyllanthus emblica, alt. 1685 m, Jun He, 26 August. 2019, L4948(HKAS 123776).

Description

Basidiomata annual, sessile and broadly attached, coriaceous, hard corky to woody hard. Pileus single or dimidiate, sub-circular, flabelliform to shell-shaped, applanate, projecting up to 22 cm, 12 cm wide and 1.9 cm thick at base. Pileus surface dark brown(8F8), purplish black(8F3) to reddish brown(6F8) and covered by a thin hard crust, laccate, glabrous and shiny, with dense concentric furrows. Margin pale yellow(4A3) to generally concolorous, entire, subacute, slightly wavy. Context up to 0.8 cm thick, homogeneous, cinnamon brown(6D7) to chestnut brown(8E5), with black melanoid lines, hard corky. Tubes 0.5–1.1 cm long, concolorous with the base of the context, corky, unstratified. Pores 5–7 per mm, circular to subcircular, dissepiments slightly thick, entire; pores surface greyish white(2B1) when fresh, orange grey(5B2) to pale brown(6D6) when bruising and drying.

Figure 2. 

Ganoderma phyllanthicola (HKAS 123776, holotype) A, B basidiomata C pore surface D cut side of pileus E, F sections of pellis G skeletal hyphae from context H binding hyphae from context I generative hyphae from tubes J–N basidiospores. Scale bars: 20 μm (G); 10 μm (E, F, H, I); 5 μm (J–N).

Hyphal system trimitic. Generative hyphae 1.0–2.0 μm in diameter, colorless, thin-walled, with clamps connections; skeletal hyphae 2.0–5.0 μm in diameter, thick-walled with a wide to narrow lumen or sub-solid, arboriform with few branches, yellowish to golden yellow; binding hyphae 1.0–3.0 μm in diameter, thick-walled, branched and flexuous, pale yellow, scarce; all the hyphae IKI–, CB+; tissues darkening in KOH.

Pileipellis a crustohymeniderm, cells 15–33 × 5–9 μm, thick-walled to sub-solid, composed of irregular, narrowly clavate end cells, straight to flexuose, smooth or with a few small apical protuberances, yellowish to golden-yellow, sometimes with apical granulations, apex slightly amyloid.

Basidiospores ellipsoid to subglobose, apex truncated or subacute, yellowish to yellowish brown, IKI–, CB+, inamyloid; double-walled with distinctly thick walls, exospore wall smooth, endospore walls with interwall pillars; (40/2/2) (8.5) 9.0–9.6–10.0 (11.0) × 6.0–6.8–7.0 (7.5) µm, L = 9.65 µm, W = 6.75 µm, Q = (1.24) 1.38–1.52 (1.55), Qm = 1.43 ± 0.07 (including myxosporium). Basidia not observed.

Additional specimens examined

China, Sichuan Province, Panzhihua City, Miyi County, on a decaying tree of Phyllanthus sp., alt. 1035 m, Jun-He, 15 August 2023, HL308.

Notes

In the phylogenetic analyses, Ganoderma phyllanthicola is clustered as a sister taxon to G. castaneum with strong statistical support (100% ML and 1.00PP, Fig. 1). Morphologically, both species share similar characteristics of the sessile basidiomata and non-stratified tubes. However, G. castaneum differs from G. phyllanthicola in having buff and obtuse pileus margin, regular palisade pileipellis, heterogeneous context, smaller basidiospores (6.2–8.5 × 4.2–6.3 μm) with smooth endospore walls, Sun et al. 2022). Ganoderma tropicum and G. philippii have similar homogeneous context, but G. tropicum has a stipitate basidiomata and buff pileus margin, samller basidiospores (6.8–10.0 × 4.0–6.4 μm), and (Steyaert 1972). Ganoderma philippii has wavy like pileus margin and brown context with black melanoid lines, smaller and obovoid basidiospores (6.0–8.0 × 3.0–4.0 μm, Moncalvo and Ryvarden 1997).

Ganoderma aridicola described from South Africa is similar to G. phyllanthicola in the sessile basidiomata with dark brown pileus surface, homogeneous context, small pores and ellipsoid basidiospores. However, G. aridicola differs by the distinctly stratified tubes and lacks branched or protuberant apical cells (Xing et al. 2016). Besides, the phylogenetic analyses separated G. aridicola and G. phyllanthicola (Fig. 1). Ganoderma multiplicatum also has pale yellow margin and irregular pileipellis cells., but it differs from G. phyllanthicola by the photo brown to reddish brown pileus surface, short stipe (1.8–3 cm) and ellipsoid basidiospores (6.0–10.0 × 4.5–7.0 μm, Gottlieb and Wright 1999).

Ganoderma suae J. He & S.H. Li, sp. nov.

MycoBank No: 853506
Fig. 3

Diagnosis

Differs from other species in the genus by its large and substipitate basidiomata, reddish brown to oxblood red pileus surface with concentric furrows and radial rugose, whitish and wavy margin, almond-shaped basidiospores, heterogeneous context and non-stratified tubes.

Etymology

The epithet ‘suae’ refers to the Chinese mycologist Prof. Hong-Yan Su, for her great contribution to the mycology.

Holotype

China. Yunnan Province., Honghe City, lvchun County, on a dead stump of a broad-leaved tree, alt. 1392 m, Jun He, 24 June 2019, L4651(HKAS 123791).

Description

Basidiomata annual, sessile to substipitate, and occasionally imbricate, woody-corky, light in weight. Pileus round-flabelliform to reniform, slightly convex to applanate; surface glabrous, projecting up to 15 cm, 10 cm wide and 2 cm thick at base. Pileus surface reddish brown(6F8) to oxblood red(9E7), weakly to strongly laccate, and covered by a thin hard crust, concentrically zonate or azonate. Margin whitish to generally concolorous, entire, acute to obtuse, smooth to irregularly wavy. Context up to 0.8 cm thick, heterogeneous, the upper layer greyish white(2B1), the lower layer cinnamon brown (6D7) to chestnut brown(6F7), bearing distinct concentric growth zones, without black melanoid lines, hard corky and fibrous. Tubes 0.2–1.2 cm long, grayish brown (6B3), corky, unstratified. Pores 4–6 per mm, circular to angular, dissepiments slightly thick, entire; pore surface lead gray (2D2) to greyish white (2B1) when fresh, golden grey (4C2) to soot brown(5F5) when bruising or aging. Stipe up to 4.5 cm long and 3.0 cm diam, generally short and thick, cylindrical, horizontal or lateral, fibrous to spongy, reddish brown (6F8) to dark brown (8F8), concolorous to generally darker than pileus.

Figure 3. 

Ganoderma suae (HKAS 123791, holotype) A, B basidiomata C pore surface D cut side of pileus E sections of pellis F skeletal hyphae from context E generative hyphae from tubes H binding hyphae from context I–K basidia and basidioles L–O basidiospores. Scale bars: 30 μm (E–H); 10 μm (I–K, O); 5 μm (L–N); 20 mm (P, Q).

Hyphal system trimitic. Generative hyphae 2.0–3.0 μm in diameter, colorless, thin-walled, hyaline, unbranched, abundant, with clamp connections; skeletal hyphae 3.0–9.0 μm in diameter, thick-walled with a narrow lumen to subsolid, non-septate, moderately branched, orange yellow to golden-yellow, predominant; binding hyphae 1.0–2.0 μm in diameter, subthick-walled to solid, non-septate, frequently branched, interwoven, colourless to yellowish, scarce, notably thinner and paler than skeletal hyphae; all the hyphae IKI–, CB+; tissues darkening in KOH.

Culture characteristics. Initially, white to yellowish white, pale yellow when growing, become orange white, pale orange, light orange and some reddish yellow to dark brown around the plugged circle of active mycelium after incubation for 3 weeks.

Pileipellis a crustohymeniderm, cells 24–43 × 6–11 μm, thick-walled to sub-solid, apical cells narrowly clavate to clavate, slightly inflated, yellowish to golden-yellow, without granulations in the apex; negative or apex slightly amyloid.

Basidiospores almond-shaped to narrow ellipsoid, apex subacute, with apical germ pore, yellowish to yellowish-brown, IKI–, CB+, inamyloid; double-walled, exospore smooth, endospore with coarse echinulate, exosporium with inter-walled pillars 0.5–0.6 μm thick; (80/4/2) (8.0) 9.0–9.7–10.5 × (5.0) 5.5–6.1–6.5 (7.0) μm, L = 9.70 µm, W = 6.10 µm, Q = (1.38) 1.45–1.79 (1.97), Qm = 1.61 ± 0.13 (including myxosporium). Basidia barrel-shaped to widely clavate, colorless, with a clamp connection and four sterigmata, thin-walled, 9–18 × 9–12 µm; basidioles pear-shaped to fusiform, colourless, thin-walled, 8–14 × 6–11 µm.

Additional specimens examined

China, Yuannan Province, Lingcang City, Yun County, on a living Quercus sp. tree, alt. 1516 m, Jun He, 4 August 2019, L4817(HKAS 123777).

Notes

Phylogenetic analyses showed that Ganoderma suae clusters as a sister taxon to G. resinaceum with good statistical support (100% ML/1.00 PP, Fig. 1). Morphologically, G. resinaceum differs from G. suae by having smaller basidiomata, reddish brown to oxblood red pileus surface and wavy margin, homogeneous context, longer pileipellis (34–59 × 6.2–9.3 μm), and larger basidiospores (11.2–12.5 × 6.5–7.4 μm, Náplavová et al. 2020; Ryvarden 2000; Torres-Torres et al. 2012).

Ganoderma zonatum also has sessile basidiomata and a whitish pileus margin, but it differs from G. suae by having an apex widened to swollen of pileipellis cells (30–70 × 5–12 μm), and larger basidiospores (11.2–12.5 × 6.5–7.4 μm, Murrill 1902).

Discussion

Ganoderma has long been regarded as one of the most important genera of medicinal fungi worldwide with more than 45 species described in China. To date, 36 species of Ganoderma have been reported from Southwest China (Yunnan, Tibet, Guizhou, and Sichuan), including 16 species originally described from China, namely G. alpinum B.K. Cui, J.H. Xing & Y.F. Sun, G. artocarpicola J. He & S.H. Li, G. dianzhongense J. He, H.Y. Su & S.H. Li, G. ellipsoideum Hapuar., T.C. Wen & K.D. Hyde 2018, G. esculentum J. He & S.H. Li, G. leucocontextum T.H. Li, W.Q. Deng, S. H. Wu, D. M. Wang & H.P. Hu, G. mutabile Y. Cao & H.S. Yuan, G. obscuratum J. He & S. H. Li, G. ovisporum H.D. Yang, T.C. Wen. G. puerense B.K. Cui, J.H. Xing & Y.F. Sun, G. sanduense Hapuar., T.C. Wen & K.D. Hyde, G. sichuanense J.D. Zhao & X.Q. Zhang, G. subangustisporum B.K. Cui, J.H. Xing & Y.F. Sun, G. weixiense Karun. & J.C. Xu, G. yunnanense J. He & S. H. Li and G. yunlingense B.K. Cui, J.H. Xing & Y.F. Sun (Zhao et al. 1983; Cao et al. 2012; Li et al. 2015; Hapuarachchi et al. 2019; He et al. 2021; He et al. 2022; Sun et al. 2022; Yang et al. 2022). During the last five years, the diversity of Ganoderma in Southwest China was mainly reported from Yunnan Province and Guizhou Province (Hapuarachchi et al. 2019; Luangharn et al. 2021; He et al. 2022; Sun et al. 2022). These studies show that there is an unrecognized diversity of Ganoderma species in southwest China. More potential new species of Ganoderma may be discovered in the future.

In this study, two new species viz G. phyllanthicola and G. suae from Southwest China are introduced based on morphology and multigene phylogeny. Ganoderma phyllanthicola and G. suae satisfied the generic concept of the genus Ganoderma (Karsten 1881). They comprise subglobose to ellipsoid or ovoid basidiospores, truncated, double-walled with thick walls, exospore wall semi-reticulate, endospore wall smooth or with conspicuous spinules, homogeneous or heterogeneous context and laccate with variable ornamentation pileus surface. When compared with each other, G. phyllanthicola and G. suae occupied distinct and distant positions in the multilocus phylogenetic tree, and the morphology of their basidiomata also exhibits distinct macro- and microscopic characters that can further differentiate the two species. Thus, based on convergent results from morphology and molecular data analyses, G. phyllanthicola and G. suae are considered to be new species to science.

Ganoderma phyllanthicola was closely related to G. castaneum, G. philippii and G. tropicum in the phylogeny inferred from the concatenated sequence data set. Morphologically, they are easily distinguishable by some macro- and microscopic characters of their basidiomata. Contrary to G. phyllanthicola, G. castaneum has a broadly attached, flabelliform, chestnut brown pileus surface with wide concentric ridges, heterogeneous context, regular palisade pileipellis cells, and broadly ellipsoid basidiospores not obviously truncated with smooth endospore walls (Sun et al. 2022; Table 2). Ganoderma philippii and G. tropicum, contrary to the new species, is characterized by flabelliform to circular, non-coriaceous basidiomata and very much smaller basidiospores (Steyaert 1972; Moncalvo and Ryvarden 1997; Table 2). Moreover, Ganoderma enigmaticum can be easily distinguished from G. phyllanthicola by the stipitate basidiomata and regular pileipellis cells (Coetzee et al. 2015). Ganoderma orbiforme has biannual or perennial basidiospores and longer pileipellis cells than those of G. phyllanthicola (Ryvarden 2000; Table 2).

Table 2.

Morphological comparison of Ganoderma phyllanthicola sp. nov., and G. suae sp. nov., with their closest relatives in the combined phylogeny.

Species Shape Context Pileipellis cells Pores (per mm) Basidiospores (μm) Reference
Ganoderma aridicola Sessile, dimidiate 2.4–3 µm thick, homogeneous, fuscous cells clavate, 30–55 × 5–8 μm 6–8 9.7–11.2 × 7.0–7.8 Xing et al. 2016
G. castaneum sessile, flabelliform up to 1.6 cm thick, heterogeneous, the upper layer pale straw yellow, the lower layer dark brown, composed of regular palisade, clavate end cells 25–40 × 3–5 μm 4–6 6.2–8.5 × 4.2–6.3 Sun et al. 2022
G. phyllanthicola sessile, sub-circular, flabelliform to shell-shaped up to 0.8 cm thick, homogeneous, cinnamon brown to chestnut brown composed of irregular, narrowly clavate end cells, straight to flexuous or irregular, 15–33 × 5–9 μm 5–7 8.5–11.0 × 6.0–7.5 this study
G. enigmaticum stipitate, globular context soft, homogenous, dark brown amyloid elements 20–46 × 5.5–9 um 3–5 8.0–11.0 × 3.5–6.0 Coetzee et al. 2015
G. lucidum stipitate to sessile thinner context of white to slightly cream color context amyloid hyphal end cells up to 7–11 μm diam 4–5 7.7–11.5 × 5.2–8.4 Ryvarden and Gilbertson 1993
G. multiplicatum sessile, flabelliform, applanate or convex up to 2 cm thick, homogeneous, cinnamon colour, darker toward
the tubes
cells clavate, cylindrical or irregular, 38–65 × 5.6–10 µm 5–6 6.0–10.0 × 4.5–7.0 Gottlieb and Wright 1999
G. orbiforme sessile, flabelliform or spathulate context up to 0.4–1.0 cm thick, triplex composed of apically acanthus like branched cells, 50–100 X 6–12 μm 4–7 7.1–12.6 × 5.2–7.7 Ryvarden 2000; Wang et al. 2014
G. philippii sessile, flabelliform to circular up to 1.4 cm thick, homogeneous, brown 5–6 6.0–8.0 × 3.0–4.0 Steyaert 1972, Moncalvo and Ryvarden 1997
G. polychromum sessile to o substipitate, flabelliform pink buff to cinnamon buff concentric growth zones 4–5 10.3–18.3 × 7.0–11.9 Murrill 1908
G. resinaceum sessile to stipitate, round-flabelliform 0.4-1.3 cm thick, homogeneous context, wood-coloured to pale tawny brown., with resinous incrustations cells clavate, narrowly clavate, or almost cylindrical, 34–59 × 6.2–9.3 μm, 3–4 9.0–13.0 × 6.0–8.0 Steyaert 1972; Ryvarden 2000; Náplavová et al. 2020
G. sessile sessile, pileus sometimes imbricate, conchate to flabelliform context thin, soft corky or woody, radially fibrous, concentrically zonate, ochraceous cylindric, smooth elements, 60–75 × 7–10 µm 4–5 12.0–16.0 × 6.0–8.0 Murrill 1902; Gottlieb and Wright 1999
G. suae sessile to substipitate, variable, reniform up to 0.8 cm thick, heterogeneous, the upper layer greyish white, the lower layer cinnamon brown, without resinous incrustations cells clavate, 24–43 × 6–11 μm 4–6 8.0–10.5 × 5.0–7.0 this study
G. tropicum usually sessile, sometimes laterally stipitate, flabelliform to shell-shaped or circular up to 2.2 cm thick, homogeneous, dark brown cells clavate, sometimes branched or protuberant, inflated and flexuous, 19–32
× 4–9 μm
4–6 6.8–10.0 × 4.0–6.4 Ryvarden 1981
G. vivianimercedianum sessile to substipitate, flabelliform in pole view 1–1.5 cm thick, homogeneous, caramel above and dark brown toward the tubes, cells clavate, apex occasionally slightly widened, 36–65 × 7.2–12 µm 3–5 9.0–12.0 × 6.0–8.0 Torres-Torres et al. 2008
G. zonatum sessile, applanate to convex homogeneous, slightly zonate, dark brow cells cylindrical to clavate, 30–70 × 5–12 µm 4–5 12.0–14.0 × 6.0–9.0 Loyd et al. 2018

Ganoderma resinaceum is known to be a Northern Hemisphere species, mainly occurring in Europe (Patouillard 1889; Moncalvo et al. 1995; Ryvarden and Melo 2014). The European specimens are easily recognized in the field by thick, soft and pale context. The first signs of genetic diversity within G. resinaceum were observed by Moncalvo (2000), and Loyd et al. (2018) showed that G. resinaceum sensu American auctores encompassed at least two distinct species, viz. G. polychromum and G. sessile. Cabarroi-Hernández et al. (2019) studies confirmed that G. resinaceum sensu auctores from China, East Africa, Europa, and both North and South America represented a species complex. Study of phylogenetic inferences based on multilocus sequences by Hernand éz et al. (2019) also showed that G. resinaceum represents a species complex.

Our results based on polygenic phylogenetic analysis also confirm that Ganoderma resinaceum represents a species complex, encompassing several distinct species, namely G. platense, G. polychromum, G. sessile, and G. suae. Ganoderma suae emerges as a newly recognized species within the G. resinaceum sensu complex group (Fig. 1). Ganoderma suae is characterized by its annual basidiomata, reddish brown to oxblood red pileus surface, heterogeneous context without resinous incrustations (without black melanoid lines), wavy margin and almond-shaped basidiospores not obviously truncated, endospore walls with dense spinules (8.0–10.5 × 5.0–7.0 μm), can be easily distinguished from G. resinaceum (Ryvarden 2000). Náplavová et al. 2020 confirmed the presence of two distinct genotypes (genotype A and genotype B) in European G. resinaceum by comparing partial sequences of the TEF1-α region and the 25 s LSU rRNA gene. Their study also showed that basidiospore sizes range between 9.6–14.4 × 6.0–8.4 µ m in genotype A and 6–12.0 × 7.2–9.6 µ m in genotype B. Besides, specimens of both genotypes share the same pileus surface (glossy with resinous layer) and almost identical coloration. Only the context color was lighter brown beige to sand yellow in genotype A and darker brown beige to ochre brown in genotype B. Ganoderma resinaceum from Europe has a special laccate and glossy with resinous layer pileus surface, homogeneous context, and larger basidiospores cells than those of G. suae. (Ryvarden 2004; Náplavová et al. 2020). Thus, Ganoderma suae from China and G. resinaceum from Europe should be recognized as two different species. Table 2 presents a morphological comparison between the new species and its closest phylogenetic neighbors. Although we are of the opinion that G. suae well represent a species on its own, more material, ideally from various localities, and DNA sequences, is necessary to reveal the species diversity and kinship of G. resinaceum complex groups.

Recent studies have shown that the specimen G. resinaceum collected from China is inconsistent with the original description; therefore, it is clear that G. resinaceum is not distributed in China (Sun et al. 2022). It’s noteworthy that we have also collected a sample (HL199) from Yunnan Province, which differs from both G. resinaceum and G. suae in terms of its distinct macro-morphology and multi-gene sequences. Regrettably, the specimens were sterile and micromorphological data were missing. In the future, collecting additional specimens will be crucial for revealing the true distribution and diversity of G. resinaceum complex groups in China.

Acknowledgment

We would like to thank Qian-Qiu Luo, Xiao Han, Tong Lv and De-Chao Chen for their help on sample collection, DNA extraction, and PCR amplification.

Additional information

Conflict of interest

The authors have declared that no competing interests exist.

Ethical statement

No ethical statement was reported.

Funding

This research was funded by Yunnan science and technology talents and platform plan (Project ID: 202305AF150162); Central guidance for local scientific and technological de­velopment funds (Project ID: 202307AB110001); Yunnan Science Technology plan project of Science and Technology Department (Project ID: 202301BD070001-021); Lingcan City Technology Innovation Talents Training Objects (Project ID 202304AC100001-RC03); Lincan City to build the national Sustainable Development agenda innovation demonstration zone science and technology project (Project ID 202204AC100001-A01).

Author contributions

Conceptualization:JH, ZLL, SHL. Formal analysis: JH. Funding acquisition: EXL, SHL, XJL. Inves tigation: JH, XQW. Methodology: JH, WZT, DW. Resources: XJL, WZT, QWZ. Software: JH. Supervision: ZLL, SHL. Writing – original draft: JH. Writing – review and editing: SHL, ZLL.

Author ORCIDs

Jun He https://orcid.org/0000-0001-7027-7206

Wan-Zhong Tan https://orcid.org/0000-0002-9355-5798

Zong-Long Luo https://orcid.org/0000-0001-7307-4885

Data availability

All of the data that support the findings of this study are available in the main text or Supplementary Information.

References

  • Cabarroi-Hernández M, Villalobos-Arámbula AR, Torres-Torres MG, Decock C, Guzmán-Dávalos L (2019) The Ganoderma weberianum-resinaceum lineage: Multilocus phylogenetic analysis and morphology confirm G. mexicanum and G. parvulum in the Neotropics. MycoKeys 59: 95–131. https://doi.org/10.3897/mycokeys.59.33182
  • Cabarroi-Hernández M, Decock C, Welti S, Amalfi M, Villalobos-Arámbula A, Aliaga-Ramos D, Morera G, Macedo-Pérez E (2023) Ganoderma from Cuba: An approach to some species based on morphology and phylogenetic analyses. Biological Journal of the Linnean Society. Linnean Society of London 140(3): 323–357. https://doi.org/10.1093/biolinnean/blad055
  • Coetzee MP, Marincowitz S, Muthelo VG, Wingfield MJ (2015) Ganoderma species, including new taxa associated with root rot of the iconic Jacaranda mimosifolia in Pretoria, South Africa. IMA Fungus 6(1): 249–256. https://doi.org/10.5598/imafungus.2015.06.01.16
  • Costa-Rezende D, Robledo GL, Drechsler-Santos ER, Glen M, Gates G, de Madrignac Bonzi BR, Popoff OF, Crespo E, Góes-Neto A (2020, August) Taxonomy and phylogeny of polypores with ganodermatoid basidiospores (Ganodermataceae). Mycological Progress 19(8): 725–741. https://doi.org/10.1007/s11557-020-01589-1
  • Darriba D, Taboada GL, Doallo R, Posada DJ (2012) ModelTest 2: More models, new heuristics and parallel computing. Nature Methods 9(8): 772–772. https://doi.org/10.1038/nmeth.2109
  • Fryssouli V, Zervakis GI, Polemis E, Typas MA (2020) A global meta-analysis of ITS rDNA sequences from material belonging to the genus Ganoderma (Basidiomycota, Polyporales) including new data from selected taxa. MycoKeys 75: 71–143. https://doi.org/10.3897/mycokeys.75.59872
  • Glez-Peña D, Gómez-Blanco D, Reboiro-Jato M, Fdez-Riverola F, Posada D (2010) ALTER: Program-oriented conversion of DNA and protein alignments. Nucleic Acids Research 38: 14–18. https://doi.org/10.1093/nar/gkq321
  • Hall TA (1999) BioEdit: A user–friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 41: 95–98.
  • Hapuarachchi KK, Elkhateeb WA, Karunarathna SC, Cheng CK, Bandara AR, Kakumyan P, Hyde KD, Daba G, Wen TC (2018a) Current status of global Ganoderma cultivation, products, industry and market. Mycosphere 9(5): 1025–1052. https://doi.org/10.5943/mycosphere/9/5/6
  • Hapuarachchi KK, Karunarathna SC, McKenzie EHC, Wu XL, Kakumyan P, Hyde KD, Wen TC (2018b) High phenotypic plasticity of Ganoderma sinense (Ganodermataceae, Polyporales) in China. Asian Journal of Mycology 2(1): 1–47. https://doi.org/10.5943/ajom/2/1/1
  • Hapuarachchi KK, Karunarathna SC, Phengsintham P, Yang HD, Kakumyan P, Hyde KD, Wen TC (2019) Ganodermataceae (Polyporales): Diversity in Greater Mekong Subregion countries (China, Laos, Myanmar, Thailand and Vietnam). Mycosphere : Journal of Fungal Biology 10(1): 221–309. https://doi.org/10.5943/mycosphere/10/1/6
  • He J, Luo ZL, Tang SM, Li YJ, Li SH, Su HY (2021) Phylogenetic analyses and morphological characters reveal two new species of Ganoderma from Yunnan province. China. MycoKeys 84: 141–162. https://doi.org/10.3897/mycokeys.84.69449
  • He J, Han X, Luo ZL, Li EX, Tang SM, Luo HM, Niu KY, Su X, Li S-H (2022) Species diversity of Ganoderma (Ganodermataceae, Polyporales) with three new species and a key to Ganoderma in Yunnan Province, China. Frontiers in Microbiology 13: 1035434. https://doi.org/10.3389/fmicb.2022.1035434
  • Hu Y, Karunarathna SC, Li H, Galappaththi MC, Zhao CL, Kakumyan P, Mortimer PE (2022) The impact of drying temperature on basidiospore size. Diversity 14(4): 239. https://doi.org/10.3390/d14040239
  • Karsten PA (1881) Enumeratio boletinearum et polyporearum Fennicum, systemate novo dispositarium. Revue Mycologicum 3: 16–19.
  • Katoh K, Standley DM (2013) MAFFT multiple sequence alignment software Version 7: Improvements in performance and usability. Molecular Biology and Evolution 30(4): 772–780. https://doi.org/10.1093/molbev/mst010
  • Kinge T R, Mih AM, Neves MD, Sousa DR, Frota MZ (2015) Diversity and distribution of species of Ganoderma in south western Cameroon. Journal of Yeast and Fungal Research 6(2): 17–24. https://doi.org/10.5897/JYFR2014.0150
  • Kornerup A, Wanscher JH (1978) Methuen handbook of colour (3rd edn.) Methuen. London, England.
  • Lai T, Gao Y, Zhou SF (2004) Global marketing of medicinal Ling Zhi mushroom Ganoderma lucidum (W.Curt:Fr.) Lloyd (Aphyllophoromycetideae) products and safety concerns. International Journal of Medicinal Mushrooms 6(2): 189–194. https://doi.org/10.1615/IntJMedMushr.v6.i2.100
  • Li TH, Hu HP, Deng WQ, Wu SH, Wang DM, Tsering T (2015) Ganoderma leucocontextum, a new member of the G. lucidum complex from southwestern China. Mycoscience 56(1): 81–85. https://doi.org/10.1016/j.myc.2014.03.005
  • Li LF, Liu HB, Zhang QW, Li ZP, Wong TL, Fung HY, Zhang JX, Bai SP, Lu AP, Han QB (2018) Comprehensive comparison of polysaccharides from Ganoderma lucidum and G. sinense: Chemical, anti-tumor, immunomodulating and gut-microbiota modulatory properties. Scientific Reports 8(1): 6172. https://doi.org/10.1038/s41598-018-22885-7
  • Loyd AL, Barnes CW, Held BW, Schink MJ, Smith ME, Smith JA, Blanchette RA (2018) Elucidating ‘lucidum’: Distinguishing the diverse laccate Ganoderma species of the United States. PLoS ONE 13(7): e0199738. https://doi.org/10.1371/journal.pone.0199738
  • Luangharn T, Karunarathna SC, Dutta AK, Paloi S, Lian CK, Huyen LT, Pham HND, Hyde KD, Xu JC, Mortimer PE (2021) Ganoderma (Ganodermataceae, Basidiomycota) species from the Greater Mekong Subregion. Journal of Fungi 7: 1–83. https://doiorg/103390/jof7100819
  • Matheny PB, Wang Z, Binder M, Curtis JM, Lim YW, Nilsson RH, Hughes KW, Petersen RH, Hofstetter V, Ammirati JF, Schoch C, Langer GE, McLaughlin DJ, Wilson AW, Crane PE, Frøslev T, Ge ZW, Kerrigan RW, Slot JC, Vellinga EC, Liang ZL, Aime MC, Baroni TJ, Fischer M, Hosaka K, Matsuura K, Seidl MT, Vaura J (2007) Hibbett DS Contributions of rpb2 and tef1 to the phylogeny of mushrooms and allies (Basidiomycota, Fungi). Molecular Phylogenetics and Evolution 43(2): 430–451. https://doi.org/10.1016/j.ympev.2006.08.024
  • Miller MA, Pfeiffer W, Schwartz T (2010) Creating the CIPRES Science Gateway for inference of large phylogenetic trees In Proceedings of the 2010 Gateway Computing Environments Workshop (GCE) New Orleans LA USA 14 November, 8 pp. https://doi.org/10.1109/GCE.2010.5676129
  • Moncalvo JM, Wang HF, Hseu RS (1995) Phylogenetic relationships in Ganoderma inferred from the internal transcribed spacers and 25S ribosomal DNA sequences. Mycologia 87(2): 223–1238. https://doi.org/10.1080/00275514.1995.12026524
  • Moncalvo JM, Ryvarden LA (1997) A nomenclatural study of the Ganodermataceae Donk. Synopsis Fungorum 11: 1–114.
  • Murrill WA (1902) The Polyporaceae of North America part I The genus Ganoderma. Bulletin of the Torrey Botanical Club 29: 599–608 https://doi.org/10.2307/2478682
  • Murrill WA (1908) Polyporaceae, Part 2. North American Flora 9: 73–131.
  • Náplavová K, Beck T, Pristaš P, Gáperová S, Šebesta M, Piknová M, Gáper J (2020) Molecular Data Reveal Unrecognized Diversity in the European Ganoderma resinaceum. Forests 11(8): 850. https://doi.org/10.3390/f11080850
  • Nylander JA (2004) A MrModeltest 20; Program Distributed by the Author; Uppsala University, Uppsala Sweden.
  • Patouillard NT (1889) Le genre Ganoderma. Bull. Soc. Mycol. Fr. 5: 64–80.
  • Rannala B, Yang Z (1996) Probability distribution of molecular evolutionary trees: A new method of phylogenetic inference. Journal of Molecular Evolution 43(3): 304–311. https://doi.org/10.1007/BF02338839
  • Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S, Larget B, Liu L, Suchard MA, Huelsenbeck JP (2012) MrBayes 32: Efficient Bayesian phylogenetic inference and model choice across a large model space. Systematic Biology 61(3): 539–542. https://doi.org/10.1093/sysbio/sys029
  • Ryvarden L (1981) Type studies in the Polyporaceae 12 Species described by F W Junghuhn. Persoonia 11: 369–372.
  • Ryvarden L (2004) Neotropical polypores part 1. Introduction, ganodermataceae & hy-menochaetaceae. Synopsis Fungorum 19: 1–229.
  • Ryvarden L, Gilbertson RL (1993) European polypores Part 1 Abortiporus Lindtneria. Synopsis Fungorum 6: 1–387.
  • Ryvarden L, Melo I (2014) Poroid Fungi of Europe; Fungiflora. Oslo, Norway, 455.
  • Steyaert RL (1972) Species of Ganoderma and related genera mainly of the Bogor and Leiden Herbaria. Persoonia 7: 55–118.
  • Sun YF, Costa-Rezende DH, Xing JH, Zhou JL, Zhang B, Gibertoni TB, Gates G, Glen M, Dai Y-C, Cui B-K (2020) Multi-gene phylogeny and taxonomy of Amauroderma s lat (Ganodermataceae). Persoonia 44(1): 206–239. https://doi.org/10.3767/persoonia.2020.44.08
  • Sun YF, Xing JH, He XL, Wu DM, Song CG, Liu S, Vlasák J, Gates G, Gibertoni TB, Cui BK (2022) Species diversity systematic revision and molecular phylogeny of Ganodermataceae (Polyporales Basidiomycota) with an emphasis on Chinese collections. Studies in Mycology 101(1): 287–415. https://doi.org/10.3114/sim.2022.101.05
  • Torres-Torres MG, Guzmán-Dávalos L, Guggliota AM (2008) Ganoderma vivianimercedianum sp nov and the related species G. perzonatum. Mycotaxon 105: 447–454.
  • Torres-Torres MG, Guzmán-Dávalos L, Guggliota AM (2012) Ganoderma in Brazil: Known species and new records. Mycotaxon 121(1): 93–132. https://doi.org/10.5248/121.93
  • Vaidya G, Lohman DJ, Meier R (2011) SequenceMatrix: Concatenation software for the fast assembly of multi-gene datasets with character set and codon information. Cladistics 27(2): 171–180. https://doi.org/10.1111/j.1096-0031.2010.00329.x
  • Vinjusha N, Kumar TKA (2022) . The Polyporales of Kerala. SporePrint Books, Calicut, Kerala, India 90–108.
  • Wang L, Li J, Zhang J, Li Z, Liu H, Wang Y (2020) Traditional uses chemical components and pharmacological activities of the genus Ganoderma P Karst: A review. RSC Advances 10(69): 84–97. https://doi.org/10.1039/D0RA07219B
  • White TJ, Bruns T, Lee S, Taylor J (1990) Amplification and direct sequencing of fungal ri-bosomal RNA genes for phylogenetics In: Innis MA, Gelfand DH, Sninsky JJ, White TJ (Eds) PCR protocols: a guide to methods and applications. Academic, San Diego / New York, 315–322. https://doi.org/10.1016/B978-0-12-372180-8.50042-1
  • Xing JH, Song J, Decock C, Cui BK (2016) Morphological characters and phylogenetic analysis reveal a new species within the Ganoderma lucidum complex from South Africa. Phytotaxa 266: 115–124. https://doiorg/1011646/phytotaxa26625
  • Yang HD, Ding Y, Wen TC, Hapuarachchi KK, Wei DP (2022) Ganoderma ovisporum sp. nov. (Polyporales, Polyporaceae) from Southwest China. Biodiversity Data Journal 10: 1–31. https://doi.org/10.3897/BDJ.10.e80034
  • Ye L, Karunarathna SC, Mortimer PE, Li HL, Qiu MH, Peng XR, Luangharn T, Li YJ, Promputtha I, Hyde K, Xu JC (2019) Ganoderma weixiensis (Polyporaceae Basidiomycota) a new member of the G. lucidum complex from Yunnan Province China. Phytotaxa 423: 75–86. https://doi.org/10.11646/phytotaxa.423.2.3
  • Zhao JD, Xu LW, Zhang XQ (1983) Taxonomic studies on Ganodermataceae of China II. Acta Mycologica Sinica 2: 159–167.

Supplementary material

Supplementary material 1 

Phylogenetic tree

Author: Jun He

Data type: pdf

Explanation note: Maximum likelihood (ML) tree is based on combined ITS + nrLSU + TEF1-α + RPB2 sequence data. Bootstrap support values with a maximum likelihood (ML) greater than 60% and Bayesian posterior probabilities (PP) greater than 0.90 given above the nodes, shown as “ML/PP”.

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (66.95 kb)
login to comment