Previous Article in Journal
Exploring the Biocontrol Potential of Phanerochaete chrysosporium against Wheat Crown Rot
Previous Article in Special Issue
Comprehensive Characterization of Tuber maculatum, New in Uruguay: Morphological, Molecular, and Aromatic Analyses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Three New Truffle Species (Tuber, Tuberaceae, Pezizales, and Ascomycota) from Yunnan, China, and Multigen Phylogenetic Arrangement within the Melanosporum Group

1
College of Resources and Environment, Yunnan Agricultural University, Kunming 650100, China
2
Amway (China) Botanical R&D Center, Wuxi 214115, China
3
The Germplasm Bank of Wild Species, Yunnan Key Laboratory for Fungal Diversity and Green Development, Kunming Institute of Botany, Chinese Academy of Sciences, 132 Lanhei Road, Kunming 650201, China
4
Herbarium, Key Laboratory for Plant Diversity and Biogeography of East Asia, Kunming Institute of Botany, Chinese Academy of Sciences, Kunming 650201, China
5
Colegio de Postgraduados, Campus Montecillo, Microbiología, Edafología, Texcoco 56230, Mexico
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
J. Fungi 2024, 10(9), 640; https://doi.org/10.3390/jof10090640 (registering DOI)
Submission received: 30 July 2024 / Revised: 5 September 2024 / Accepted: 5 September 2024 / Published: 7 September 2024
(This article belongs to the Special Issue New Perspectives on Tuber Fungi)

Abstract

:
Based on a multi-locus phylogeny of a combined dataset of ITS, LSU, tef1-α, and rpb2 and comprehensive morphological analyses, we describe three new species from the Melanosporum group of genus Tuber and synonymize T. pseudobrumale and T. melanoexcavatum. Phylogenetically, the three newly described species, T. yunnanense, T. melanoumbilicatum and T. microexcavatum, differ significantly in genetic distance from any previously known species. Morphologically, T. yunnanense is distinctly different from its closest phylogenetically related species, T. longispinosum, due to its long shuttle-shape spores (average the ratio of spore length to spore width for all spores (Qm) = 1.74). Tuber melanoumbilicatum differs from the other species in having a cavity and long shuttle-shaped spores (Qm = 1.65). Although T. microexcavatum sampled ascomata have relatively low maturity, they can be distinguished from its closely related species T. pseudobrumale by the ascomata size, surface warts, and spore number per asci; additionally, phylogenetic analysis supports it as a new species. In addition, molecular analysis from 22 newly collected specimens and Genebank data indicate that T. pseudobrumale and T. melanoexcavatum are clustered into a single well-supported clade (Bootstrap (BS) = 100, posterior probabilities (PP) = 1.0); and morphological characteristics do not differ. Therefore, based on the above evidence and publication dates, we conclude that T. melanoexcavatum is a synonym of T. pseudobrumale. By taking into account current knowledge and combining the molecular, multigene phylogenetic clade arrangement and morphological data, we propose that the Melanosporum group should be divided into four subgroups. Diagnostic morphological features and an identification key of all known species in the Melanosporum group are also included. Finally, we also provide some additions to the knowledge of the characterization of T. pseudobrumale, T. variabilisporum, and T. pseudohimalayense included in subgroup 1 of the Melanosporum group.

1. Introduction

Tuber F.H. Wigg. (Tuberaceae, Pezizales) is renowned for its distinctive aroma, flavor, and nutritional benefits, including high fiber, protein, and low fat [1,2,3]. Some species of this genus are among the most highly prized edible wild mushrooms in international markets including mainly T. magnatum Pico and T. melanosporum Vittad. [4]. The genus Tuber is species-rich and geographically widespread, with over 180 species in the world mainly distributed in Europe, North America, and Southeast Asia [5]. In China, 82 phylogenetic species have been recorded, 68 of which are only found in this country, showing a high rate of endemism [6]. Tuber establish typical ectomycorrhizal symbioses with woody plant species belonging mainly to Pinaceae, Fagaceae, Corylaceae, and Betulaceae [7,8,9]. This relationship plays a role of paramount importance in the functioning of forest systems because it enhances plant nutrient absorption and resilience and also has a crucial role in plant–fungus–animal coevolution [10,11,12,13].
The study of the genus Tuber originated in Europe and the United States, where the Melanosporum group is considered one of the most important genealogical branches of the genus because of its food value, species diversity, and wide geographical distribution [5,14,15,16,17,18]. In terms of the timeline of species descriptions, the earliest described species of black truffle is T. melanosporum, which was described by the French mycologist Jean Louis Étienne Alexandre Brongniart in 1813 [19]. Since then, T. melanosporum has been one of the most studied species within the Tuber genus, attracting significant attention from researchers and food enthusiasts alike due to its culinary and economic value and its wide geographical distribution [2,20,21]. In 1831, the Italian botanist Carlo Vittadini provided a systematic description of species within the Tuber genus in his work “Monographia Tuberacearum”, which included T. brumale [22]. As one of the most widely distributed black truffles globally, T. brumale plays a crucial role in taxonomic studies of Tuber through morphological and molecular analyses [20,23]. In the late 19th century, the earliest Asian black truffle species was discovered in India. In 1892, Duthie collected a species resembling the T. melanosporum under Quercus incana Roxb. in the Mussoorie region of northern India. These specimens were sent to the Royal Botanic Gardens, Kew, and subsequently published as a new species, T. indicum Cooke and Massee [24]. In the 1980s, Zhang studied Chinese specimens resembling the type specimen of T. indicum. Based on the ornamentation of the spores, individuals primarily characterized by spiny ornamentation were identified as T. indicum, while those with regular or irregular reticulate ornamentation on the spores were described as a new species, named T. himalayense Zhang and Minter [25]. In 1988, Tao collected black truffles in Sichuan, China, and the following year, published T. sinense K. Tao and B. Liu [26]. Concurrently, Hu discovered a black truffle species under Q. glauca Thunb. in Nantou County, Taiwan, and named it T. formosanum H. T. Hu [27]. However, the publication did not specify a type specimen, unfortunately rendering it an invalid species name. It was not until 2013 that the species was formally published as T. formosanum by H. T. Hu and Y. Wang [28]. The categorization and naming of Asian black truffles, particularly those originating from China, have historically been subject to confusion and debate. In 1999, Roux et al. analyzed Asian black truffles from China using ITS fragments and found that these samples were divided into two phylogenetic distinct branches, representing T. indicum and T. himalayense [29]. Later scholars, including Zhang et al. and Feng et al. in China and Kinoshita et al. in Japan, held the same view on the interpretation of these two branches [30,31,32]. Meanwhile, T. sinense and T. formosanum are considered to be homonyms of T. indicum and T. himalayense, respectively. There are also scholars who have not agreed with the interpretation of these two branches and have replaced them as T. indicum-A and T. indicum-B [15,16,33]. In this classification’s proposal, T. sinense and T. himalayense are treated as synonyms of T. indicum [34]. One of the reasons for the confusion in the classification of black truffle species is that their morphological characteristics are highly variable, and a single morphological method cannot identify species with a high degree of similarity [30,32,34,35]. By 2018, phylogenetic analyses of more than 200 rDNA-ITS sequences from different parts of Asia by Fan et al. showed that the black truffle species cluster in Asia contained at least four species, with three branches containing the type sequences of T. sinense, T. formosanum, and T. yigongense L. Fan and W. P. Xiong, from China, respectively, and one branch containing T. longispinosum A. Kinosh. from Japan [32,35]. According to recent studies, the Asian black truffle species group is reported to consist of six species, of which four, namely T. formosanum, T. sinense, T. yigongense, and T. longispinosum, are supported by molecular evidence, while no plausible DNA data were available for T. indicum and T. himalayense samples of Indian origin [35]. In addition to the closely related species mentioned above, there are also species from one branch of the Melanosporum group that are vaguely delimited. Moreno et al. described T. pseudohimalayense G. Moreno, Manjón, Diez, and Garcia-Montero [36] and Wang et al. [37] described T. pseudoexcavatum Y. Wang and G. Moreno, but it has been concluded that T. pseudohimalayense and T. pseudoexcavatum are a single species [38,39]. While Asian black truffle species were discovered, Guevara et al. identified T. regimontanum Guevara, Bonito and Rodríguez in Mexico [40]. In recent years, three black truffle species, T. pseudobrumale Y. Wang and S. H. Li, T. melanoexcavatum Y. Wang and S. H. Li, and T. variabilisporum L. Fan and T. Li were described from southwestern China [6,41,42]. Therefore, currently, the Melanosporum group contains 11 species supported by molecular data.
In our recent investigations, we collected a number of black truffle species and conducted a combined detailed analysis of their morphology and polygenic sequences and compared them with those of previously known species. Our aim was to define the species diversity and phylogenetic relationships within the Melanosporum group. The results revealed the discovery of three new species, namely T. yunnanense S. P. Wan, R. Wang and F.Q. Yu, sp. nov. and T. melanoumbilicatum S. P. Wan, R. Wang, and F.Q. Yu, sp. nov., which were morphologically mature, and T. microexcavatum S. P. Wan, R. Wang, and F.Q. Yu, sp. nov., which lacks mature spores and is described solely by its macro- and micromorphological characterization of ascomata, peridium, and gleba as well as the immature spores and asci, along with the sequences of four genes sequenced from its DNA. Interestingly, our findings also indicated that the morphological phylogenetic analyses indicate that T. melanoexcavatum is a synonym of T. pseudobrumale.

2. Materials and Methods

2.1. Morphological Study

Fresh samples were collected from Yunnan, China. The macromorphological description was based on fresh ascomata and microscopic and macroscopic characteristics were described following the methods of Kumar et al. [43]. Hand-cut sections were mounted in 5% (w/v) KOH and examined under a light microscope (Leica DM2500, Leica Microsystems, Wetzlar, Germany). To assess the range of spore size, 376 ascospores were measured from T. yunnanense specimens and 423 from T. melanoumbilicatum specimens. Measurements of ascospores are given as (a–) b–c(–d), where b–c includes a minimum of 90% of the measured values. Extreme values (a and d) are given in parentheses. The abbreviation “Q” represents the ratio range of spore length to spore width calculated for each spore and “Qm ± ssd (sample standard deviation)” for the average Q of all spores ± ssd. For scanning electron microscopy (SEM), spores were scraped from the dried gleba onto double-sided tape and this was mounted directly on an SEM stub, coated with gold-palladium, and examined and photographed using a JSM-5600LV SEM (JEOL, Tokyo, Japan). The specimens are deposited at the Herbarium of Yunnan Agricultural University (YNAU).
In the present study, we conducted a comparative analysis of the morphological characters of all species of the Melanosporum group. Additionally, we studied the morphological features of 22 collected specimens resembling T. pseudobrumale in detail (Table 1). The ascomata surface ornamentation, peridium, asci, and ascospores of the 22 specimens were studied in detail, and at least 98 spores were measured for each specimen. A detailed morphological description of the harvested specimens was provided to further elucidate the morphology of T. pseudobrumale. Due to their recorded natural morphological variation, we also include photographs of the ascomata of two additional species: T. variabilisporum and T. pseudohimalayense.

2.2. Molecular Methods

The total DNA was extracted from pieces of dried ascomata with a modified CTAB procedure [44]. Standard and touchdown polymerase chain reaction (PCR) protocols along with fungal-specific primer sets were used to amplify and sequence four regions: the internal transcribed spacer of ribosomal RNA gene (ITS), the 28S large subunit of ribosomal RNA gene (LSU), elongation factor (tef1-α), and the second largest subunit of RNA polymerase II (rpb2) [5].
The PCR was performed using the following procedure: 25 μL of PCR reaction solution contained 1 μL DNA (50 ng/μL), 1 μL (5 μm) of each primer pair, 2.5 μL 10 × buffer (Mg2+plus), 1 μL dNTP (1 mM), 0.5 μL BSA (0.1%), 0.5 μL MgCl2 (2.0 mM), and 1 U of Taq DNA polymerase (Takara Tag, Takara Biotechnology, Dalian, China). The thermal cycling conditions were run as follows: an initial denaturation at 94 °C for 5 min, followed by 35 cycles of 94 °C for 1 min, annealing at 52 °C for 1 min, and 72 °C for 1 min. The final reaction was followed by extension at 72 °C for 10 min. The PCR products were verified on 1% agarose electrophoresis gels stained with ethidium bromide. The PCR products were subsequently purified and sequenced by Tsingke Biotech Corporation (Beijing, China). The obtained alignment and respective phylogenetic tree were deposited in the TreeBASE with the submission ID 31641.

2.3. Sequence Analysis

Sequences (ITS, LSU, tef1-α, and rpb2) from the studied specimens were compiled along with sequences from reference taxa curated in GenBank (http://www.ncbi.nlm.nih.gov/ (accessed on 12 October 2023)). A total of 109 taxa including holotypes were analyzed (Table A1). A dataset (ITS, LSU, tef1-α, and rpb2) was used to clarify the phylogenetic position of the new species. Two sequences derived from Choiromyces sichuanense S. P. Wan, R. Wang, and F.Q. Yu were selected and used as an outgroup.
Datasets were aligned using Multiple Alignment using Fast Fourier Transform (MAFFT) v.7.0 [45] and then manually edited with BioEdit v.7.0.9 as needed [46]. Phylogenetic relationships among the taxa were inferred using Maximum likelihood (ML) and Bayesian inference (BI) methods. ML bootstrap (BS) replicates (1000) were computed in Randomized Axelerated Maximum Likelihood (RAxML) with a rapid bootstrap analysis and a search for the best-scoring ML tree. Bayesian analysis was conducted using the selected model, with four chains sampled every 100 generations, over a total of 3,000,000 generations. The average standard deviations of split frequencies were less than 0.01 at the end of the run and effective sampling size (ESS) values exceeded 200. A majority rule consensus tree was built after discarding trees from 25% of the initial trees as burn-in. Posterior probabilities (PP) were calculated using the sumt command implemented in MrBayes.
Table 1. Morphological characteristics of all species within the Melanosporum group.
Table 1. Morphological characteristics of all species within the Melanosporum group.
CavityFungal TaxaVoucher
Specimens
Ascomata Surface Asci Spore NumberAscosporesSource
ColorWartsQ Interval of All Spores,
Average Value of All Spores and Main Shape
Size (μm)Ornamentation
YesTuber brumaleblackpenta- or
hexagonal flat warts
3–5nd
Qm = 1.6 ± 0.1,
mainly long ellipsoid
(25.3)–28.1–(33.7) × (15.7)–17.4–(19.1) μm3.4 ± 0.38 μm,
spino-reticulate
Vittadini (1831) [22]
Wang et al. (2006) [34]
Tuber pseudohimalayenseAH 18331blackPyramidal1–7(8)nd
Qm = 1.30 ± 0.09,
mainly broadly ellipsoidal,
1-spored: 34–35 × 25–30 μm
2-spored: 30–34 × 22–30 μm
3–7-spored: —
(including the ornamentation)
4–6(–8) μm, spino-reticulateMoreno et al. (1997) [36]
Wang et al. (2006) [34]
Chen & Liu (2011) [33]
Tuber melanoumbilicatumYNAU017
holotype
blacksolid,
irregular polygonal, pyramidal warts
(1–)2–7Q = 1.2–2.0,
Qm = 1.65 ± 0.1,
mainly long ellipsoid
1-spored: (42.5–)44.6–48.8(–54.5) × (25.7–)26.6–29.1(–30.2) μm
2-spored: (26.3–)30.9–41.4(–42.2) × (18.2–)18.5–23.9(–24.6) μm
3-spored: (19.2–)20.9–35.4(-36.9) × (13.7–)15.4–22.2(–26.2) μm
4-spored: (21.5–)24.3–31.7(–32.2) – (14.2–)14.9–20.2(–21.4) μm
5-spored: (21.7–)22.2–29.9(–33.1) × (13.1–)14.2–18.1(–20.2) μm
6-spored: (19.2–)21.7–28.0(–29.4) × (11.7)13.1–17.5(–18.9) μm
7-spored: (16.8–)18.9–26.8(–27.4) × (8.8–)12.5–16.3(–16.8) μm
0.9–8.2 μm,
spino-reticulate
This study
Tuber pseudobrumaleYAAS L3181
holotype
black low
pyramidal warts
3–7nd
Q = 1.27,
mainly ellipsoid
3-spored: 26–30 × 15.5–17.5 μm
4-spored: (22)23–25.5(27) × 14–17 μm
5-spored: (21)22–25(25.4) × 13.5–15(16) μm
6-spored: 21–23.5 × 13–14.5 μm
7-spored: 21–22.5 × 12–14 μm
4–5 μm spino-reticulateLi et al. (2014) [41]
Tuber melanoexcavatumYAAS L3605
holotype
blackpyramidal
warts
5–8nd
Q = 1.19,
mainly ellipsoid,
5-spored: 22–24.7 × 15.4–16.9 μm
6-spored: 21.4–24 × 14.3–16.0 μm
7-spored: 20–22 × 13.5–15.5 μm
8-spored: 18.7–21.2 × 12.6–15.1 μm
3–4 μm spino-reticulateWang et al. (2020) [42]
Tuber pseudobrumale(22 samples shown on Table A1)blacksolid, low, concave,
irregular, polygonal warts
1–6(–8)Q = 1.1–1.9,
Qm = 1.45 ± 0.2,
malily ellipsoid
1-spored: (19.1–)22.4–42.2(–52.3) × (13.8–)16.2–27.8(–36.7) μm
2-spored: (17.5–)21.8–36.3(–42.5) × (12.6–)16.1–23.9(–31.9) μm
3-spored: (13.8–)20.1–26.1(–40.6) × (12.4–)14.7–21.8(–29.0) μm
4-spored: (13.3–)17.9–28.3(–33.9) × (9.6–)13.5–19.3(–26.1) μm
5-spored: (13.0–)17.3–26.5(–32.0) × (9.6–)12.6–18.4(–22.8) μm
6-spored: (11.5–)16.7–25.4(–33.5) × (9.1–)12.3–18.2(–23.3) μm
7-spored: (9.6–)12.8–23.7(–28.4) × (5.6–)9.6–16.1(–18.1) μm
8-spored: (15.9–)18.2–21.9(–22.7) × (10.5–)12.4–15.1(–15.7) μm
0.4–9.6 μm,
spino-reticulate
This study
Tuber variabilisporumBJTC FAN362
holotype
dark brown
to black brown
verrucose1–5(–6)Q = 1.06–1.44,
nd
broadly ellipsoid and ellipsoid
1-spored: 30–37.5 × 21.75–27.5 μm
2-spored: 27.5–32.5 ×20–23.5 μm
3-spored: 22.5–32.5 × 18.75–21.25 μm
4-spored: 17.5–25.5 ×16.5–20 μm
5-spored: 16.5–22.5 × 13.25–17.5 μm
3–5 μm,
spino-reticulate
Fan et al. (2022) [6]
Tuber microexcavatumYNAU 1263
holotype
yellowish brownloose-textured, cracked irregular warts1–6nd
nd
ellipsoid
ndndThis study
Tuber sp. 5K229blackpyramidal warts5–8Q = 1.0–2.0,
nd
ellipsoid
15–20 × 10–15 μmnd
spino-reticulate
Kinoshita et al. (2011) [17]
NoTuber longispinosumK447
holotype
brown to dark greyishlow
polygonal warts
1–5(–6)Q = 1.0–2.1,
nd
ellipsoid to subglobose
1-spored: 31–41 × 22–30 μm
2-spored: 21–38 × 16–29 μm
3-spored: 19–34 × 15–26 μm
4-spored: 15–33 × 13–22 μm
5-spored: 16–31 × 12–20 μm
6-spored: 15–26 × 13–18 μm
3–7 (–12) μm, spinyKinoshita et al.
(2018) [32]
Tuber yunnanenseYNAU019
holotype
dark brown to blacksolid, irregular polygonal, clustered pyramidal ridged warts 1–5(–6)Q = 1.1–2.2,
Qm = 1.74 ± 0.1,
malily long shuttle-shaped
1-spored: (31.8–)32.3–52.8(–54.6) × (19.1–)20.2–33.1(–35.4) μm
2-spored: (25.4–)30.9–39.8(–43.6) × (15.7–)17.1–22.7(–25.1) μm
3-spored: (26.7–)28.1–35.9(–43.3) × (14.7–)15.4–21.4(–29.4) μm
4-spored: (25.1–)26.1–31.6(–32.7) × (14.3–)14.7–18.3(–20.6) μm
5-spored: (20.0–)20.6–28.2(–28.6) × (11.6–)12.8–17.1(–18.3) μm
6-spored: (22.5–)23.1–28.8(–29.4) × (11.7–)14.0–18.6(–19.2) μm
0.7–11.1 μm, spinyThis study
Tuber regimontanumITCV 909
holotype
dark brown to blackpyramidal verrucae1–4nd
nd
broadly fusiform to ellipsoid.
1-spore: 40–55 (–62) × 30–31 μm
2-spore: 37–42 × 25–26 μm
3-spore: 33–37 × 23–26 μm
4-spore: 28–35 × 18–22 μm
2–5 × 1–2 µm, spino-reticulateGuevara et al. (2008) [40]
Tuber yigongenseBJTC FAN731
holotype
dark brown
to blackish
pentagonal and pyramidal warts1–5nd
nd
malily ellipsoid
1-spored: 35–45 × 25–30 μm
2-spored: 30–37.5 × 20–25 μm
3–5-spored: 20–32.5 × 17.5–22.5 μm
2.5–4 µm, densely spino-reticulate Fan et al. (2018) [35]
Tuber sinense=
T. indicum?
MHSU
1633
brown, reddish brown or deeply brownverrucose1–4nd
nd
malily ellipsoid
1-spore: 32–36.5 × 43–49.5 μm
2-spores: 26.5–30 × 39–45.5 μm
3-4-spores: 22.5–25 × 30–35 μm (including spines)
3–6(–7) µm,
spiny
Tao et al. (1989) [26]
Wang et al. (2006) [34]
Chen (2007) [47]
Fan et al. (2022) [6]
Tuber formosanumHKAS
62628
dark reddish brown to dark grayish brownlow
pyramidal warts
1–4(–5) Q = (1.17–)1.27–1.62(–1.70)
nd
malily ellipsoid
1-spored: (27–)29–45(–48) × 20–32(–35) μm
2-spored: (26–)27–36(–39) × (18–)19–24(–28) μm
3-spored: 24–34 × (16–)18–23(–25) μm
4-spored:(25–)26–32(–33) × (17–)18–22 μm
2–5(–6) µm, spino-reticulateQiao et al. (2013) [28]
Tuber melanosporumblackish verrucose1–5Q = 1.4–2.1
nd
malily ellipsoid
28–32 × 16–21 μm2–4 μm.spiny,
spino-reticulate
Vittadini (1831) [22]
Wang et al. (2006) [34]
nd = not determined.

3. Results

3.1. Phylogenetic Analysis

Phylogenetic relationships were assessed using concatenated sequence data from four loci (ITS, LSU, tef1-α, and rpb2), totaling 3462 characters and including 109 representative sequences from various Tuber species. The Bayesian analysis yielded similar trees to the parsimony analysis; therefore, only the tree inferred from the ML analysis is shown in Figure 1.
Based on the sequences (ITS, LSU, tef1-α, and rpb2), all samples formed seven well-supported groups, representing the Melanosporum group, Rufum group, Excavatum group, Aestivum group, Puberulum group, Macrosporum group, and the outgroup (C. sichuanensis). All analyzed species in the Melanosporum group formed a monophyletic group with bootstrap support (BS = 100, PP = 1.0) and, based on the topology of the multigene phylogenetic analyses, we propose dividing them into four subgroups (termed then as subgroups 1, 2, 3, and 4).
The phylogenetic tree based on ITS, LSU, tef1-α, and rpb2 datasets confirmed the presence of 14 phylogenetic species in the Melanosporum group (Figure 1). Each of the three new species (T. microexcavatum, T. yunnanense, and T. melanoumbilicatum) formed separate phylogenetic branches corresponding to subgroup 1, subgroup 2, and subgroup 4, respectively. Tuber microexcavatum was identified as a new species in subgroup 1 due to its 79.6% ITS similarity with the closely related T. pseudobrumale. ITS rDNA sequence analysis also showed that T. pseudobrumale and T. melanoexcavatum have similarities ranging from 98.0% to 99.7%, suggesting that they are the same species. Therefore, subgroup 1 comprised five species, with ITS similarities ranging from 70.5% to 94.5%. Another new species, T. yunnanense, formed a separate branch in subgroup 2 with strong support (BS = 100, PP = 1.0). Therefore, subgroup 2 included seven species, five of which are Asian species including T. sinense, T. formosanum, T. yigongense, and T. yunnanense from China and T. longispinosum from Japan. The other two species are the well-known European species T. melanosporum and the Mexican species T. regimontanum. The ITS similarities among the species in subgroup 2 ranged from 87.0% to 93.0%. Subgroup 3 included only the European species T. brumale, while subgroup 4 only included the new Chinese species T. melanoumbilicatum, ITS rDNA sequence analysis showed that the similarity between subgroups 3 and 4 was 70.7%.

3.2. Taxonomy

3.2.1. Tuber yunnanense S. P. Wan, R. Wang and F.Q. Yu, sp. nov. Figure 2

MycoBank: MB849430
Etymology: Refers to the location of the type collection.
Typification: CHINA. Yunnan Province, Gongshan County, 29 October 2020, collected from Pinus sp., wsp973-1 (holotype YNAU019), dried specimens. GenBank: ITS = OK625306; LSU = OR661811; tef1-α = OR813081; rpb2 = OR832407.
Diagnosis: Ascomata are black with solid irregular polygonal pyramidal warts on the surface. Pseudoparenchymatous peridium. Each ascus contains 1–5(–6) spores, and some asci have short stalks. Ascospores are predominantly long shuttle-ellipsoids that are golden yellow, measuring 20.0–54.6 × 17.1–35.4 μm, with sharp spines.
Description: Ascomata are dark brown to black, irregularly spherical, and range from 1.9–5.5 cm in diameter. The surface features grooves and is covered with black solid irregular polygonal pyramidal ridged warts, 0.4–1.3 mm high, with occasional cracks and depressions at the apex. The peridium is composed of two layers, pseudoparenchymatous; the outer layer is 103.5–193.6 μm thick, composed of irregular cells, 3.7–15.6 × 2.8–10.9 μm, yellowish-brown, or hyaline; the inner layer is 45.5–259.9 μm thick, composed of intricately interwoven hyaline and thin-walled hyphae that are 0.5–2.4 μm in diameter. The gleba is solid, brown to black when mature, and marbled with white veins. It is composed of hyaline and interwoven thin-walled hyphae, 0.8–1.5 μm broad at the septa; with cylindrical to inflated cells, 10.6–23.2 × 7.2–16.3 μm. Asci are irregularly shaped, 47.2–128.9 × 30.9–98.2 μm (n = 142), with size and shape varying depending on the number of ascospores, with 1–5(–6) spores per ascus. Most asci are sessile, with a few having short stalks, 4.8–8.2 × 5.2–7.7 μm (n = 4). Ascospores are golden yellow, ellipsoid, or subglobose. Spikes of different lengths are attached to the outer layer of each spore. Spore sizes are as follows: in 1-spored asci: (31.8–)32.3–52.8(–54.6) × (19.1–)20.2–33.1(–35.4) μm, Q = 1.4–2.0, Qm = 1.7 ± 0.15, spines = 1.2–11.1 μm (n = 40); in 2-spored asci: (25.4–)30.9–39.8(–43.6) × (15.7–)17.1–22.7(–25.1) μm, Q = 1.4–2.1, Qm = 1.8 ± 0.13, spines = 0.7–9.5 μm (n = 60); in 3-spored asci (26.7–)28.1–35.9(–43.3) × (14.7–)15.4–21.4(–29.4) μm, Q = 1.1–2.2, Qm = 1.8 ± 0.16, spines = 0.9–7.9 μm (n = 84); in 4-spored: (25.1–)26.1–31.6(–32.7) × (14.3–)14.7–18.3(–20.6) μm, Q = 1.5–2.1, Qm = 1.8 ± 0.12, spines = 0.7–7.1 μm (n = 120); in 5-spored asci: (20.0–)20.6–28.2(–28.6) × (11.6–)12.8–17.1(–18.3) μm, Q = 1.1–2.0, Qm = 1.7 ± 0.16, spines = 0.9–6.8 μm (n = 60); in 6-spored asci: (22.5–)23.1–28.8(–29.4) × (11.7–)14.0–18.6(–19.2) μm, Q = 1.5–1.9, Qm = 1.6 ± 0.12, spines = 0.8–4.9 μm (n = 12), spiny, 0.7–11.1 μm.
Additional material examined: CHINA, Yunnan Province, Gongshan County, 29 October 2020, collected from Pinus sp., wsp973-2 (YNAU020). GenBank: ITS = OK625307; LSU = OR661812; tef1-α = OR813082; rpb2 = OR832408; ibid., wsp974-3 (YNAU0107). GenBank: ITS = OR665397; LSU = OR661813; tef1-α = OR813083; rpb2 = OR832409; CHINA, Sichuan Province, 12 September 2021, collected from Pinus sp., wsp1365 (YNAU0491). GenBank: ITS = OR250186; LSU = OR661814; tef1-α = OR813084; rpb2 = OR832410.
Figure 2. Tuber yunnanense (YNAU019, holotype). (A) Ascoma and gleba appearance; (B) Warts on surface ascoma; (C) Peridium hyphal arrangements; (D,E) Asci and ascospores under bright field microscopy; (F) Ascospore under scanning electronic microscopy. The scale bars are individually indicated for each image.
Figure 2. Tuber yunnanense (YNAU019, holotype). (A) Ascoma and gleba appearance; (B) Warts on surface ascoma; (C) Peridium hyphal arrangements; (D,E) Asci and ascospores under bright field microscopy; (F) Ascospore under scanning electronic microscopy. The scale bars are individually indicated for each image.
Jof 10 00640 g002

3.2.2. Tuber melanoumbilicatum S. P. Wan, R. Wang, and F.Q. Yu, sp. nov. Figure 3

MycoBank: MB849431
Etymology: The Latin name melanoumbilicatum derives from ‘melano’ (black) referring to black ascomata and ‘umbilicatum’, indicating the navel-like structure of the ascomata.
Typification: CHINA. Yunnan Province, Baoshan City, 5 December 2020, collected from Pinus sp., wsp1006 (holotype YNAU017), dried specimens. GenBank: ITS = OK625304; LSU = OR661815; tef1-α = OR832379; rpb2 = OR832411.
Diagnosis: Ascomata are black, with a distinct cavity, and are covered with black solid pyramidal warts. Pseudoparenchymatous peridium. Each ascus contains (1–)2–7 spores, sessile. Ascospores are predominantly long ellipsoid and black gold in color, measure 16.8–54.5 × 8.8–30.2 μm, and are spinoreticulate.
Description: Ascomata are black, with a distinct cavity, 2.7–3.5 cm in diameter. The surface is covered with black solid irregular polygonal sharp cones with verrucae convex 0.1–0.4 mm. The peridium is composed of two layers, pseudoparenchymatous; the outer layer, 142.3–255.4 μm thick, consists of subglobose to subangular cells, 6.6–29.5 × 3.5–11.5 μm, in yellow, pale brown, or hyaline; the inner layer, 55.1–133.4 μm thick, consists of intricately interwoven hyaline and thin-walled hyphae, 0.4–2.3 μm in diameter. The gleba is solid, white when young, and black when mature, marbled with white veins. It is composed of hyaline and interwoven thin-walled hyphae, 1.3–4.1 μm wide at the septa; with cylindrical to inflated cells, 3.8–10.8 × 3.5–6.7 μm. Asci are irregularly spherical and 61.5–81.9 × 47.4–72.9 μm (n = 106). Each ascus contains (1–)2–7 spores, sessile. Ascospores are black gold in color, mainly long ellipsoid, occasionally wide ellipsoidal, and long shuttle-shaped, with spines of varying lengths. Spore sizes are as follows: in 1-spored asci: (42.5–)44.6–48.8(–54.5) × (25.7–)26.6–29.1(–30.2) μm, Q = 1.5–1.9, Qm = 1.7 ± 0.15, spines = 2.2–8.0 μm (n = 5); in 2-spored asci: (26.3–)30.9–41.4(–42.2) × (18.2–)18.5–23.9(–24.6) μm, Q = 1.4–2.0, Qm = 1.7 ± 0.12, spines = 1.5–7.9 μm (n = 50); in 3-spored asci: (19.2–)20.9–35.4(–36.9) × (13.7–)15.4–22.2(–26.2) μm, Q = 1.2–2.0, Qm = 1.6 ± 0.16, spines = 1.2–8.2 μm (n = 48); in 4-spored asci: (21.5–)24.3–31.7(–32.2) × (14.2–)14.9–20.2(–21.4) μm, Q = 1.4–2.1, Qm = 1.7 ± 0.15, spines = 1.1–6.9 μm (n = 60); in 5-spored asci: (21.7–)22.2–29.9(–33.1) × (13.1–)14.2–18.1(–20.2) μm, Q = 1.2–2.0, Qm = 1.6 ± 0.15, spines = 0.9–6.5 μm (n = 100); in 6-spored asci: (19.2–)21.7–28.0(–29.4) × (11.7)13.1–17.5(–18.9) μm, Q = 1.3–2.0, Qm = 1.7 ± 0.13, spines = 0.9–6.3 μm (n = 90); and in 7-spored asci: (16.8–)18.9–26.8(–27.4) × (8.8–)12.5–16.3(–16.8) μm, Q = 1.2–2.0, Qm = 1.6 ± 0.15, spines = 0.9–6.0 μm (n = 70), spino-reticulate, reticulum with 6–10 meshes along the spore length and 6–8 across.
Additional material examined: CHINA, Yunnan Province, Baoshan City, 5 December 2020, collected from Pinus sp., wsp1006-1 (YNAU018). GenBank: ITS = OK625305; LSU = OR661816; tef1-α = OR832380; rpb2 = OR832412.
Figure 3. Tuber melanoumbilicatum (YNAU017, holotype). (A) Ascomatata and gleba in cross-section; (B) Warts on surface ascoma; (C) Close-up to gleba; (D) Peridium hyphal arrangement; (E) Asci and ascospores under bright field microscopy; (F) Ascospore under scanning electron microscopy. The scale bars are individually indicated for each image.
Figure 3. Tuber melanoumbilicatum (YNAU017, holotype). (A) Ascomatata and gleba in cross-section; (B) Warts on surface ascoma; (C) Close-up to gleba; (D) Peridium hyphal arrangement; (E) Asci and ascospores under bright field microscopy; (F) Ascospore under scanning electron microscopy. The scale bars are individually indicated for each image.
Jof 10 00640 g003

3.2.3. Tuber microexcavatum S. P. Wan, R. Wang, and F.Q. Yu, sp. nov. Figure 4

MycoBank: MB854066
Etymology: The species name microexcavatum is derived from Latin ‘micro’ referring to the small ascomata and ‘excavatum’ referring to ascomata having a navel-like cavity.
Typification: CHINA. Yunnan Province, Luquan County, 11 August 2022, collected from Platycarya strobilacea Maxim., occasionally on Pinus armandii Franch., wsp2087 (holotype YNAU1263), dried specimens. GenBank: ITS = OR250184; LSU = OR661838; tef1-α = OR832381; rpb2 = OR832413.
Diagnosis: Ascomata brown, with a distinct cavity, covered by yellow to brown loose-textured with cracked irregular warts. Pseudoparenchymatous peridium. Each ascus contains 1–6 spores, sessile.
Description: Ascomata has a distinct cavity, with a diameter of 0.7–0.9 cm. The surface is covered by yellow to brown loose-textured cracked irregular warts. The peridium is composed of two layers, pseudoparenchymatous; the outer layer that is 89.3–188.8 μm thick, composed of subglobose to subangular cells, 8.2–23.2 × 6.2–16.0 μm, yellowish brown, pale brown or hyaline; the inner layer is 137.6–224.7 μm thick, composed of intricately interwoven hyaline and thin-walled hyphae, 0.5–2.2 μm in diameter. The gleba is solid, marbled with white veins. It is composed of hyaline, interwoven, thin-walled hyphae, 1.1–3.4 μm broad at the septa, with cylindrical interwoven to inflated cells, 27.8–43.0 × 16.4–25.0 μm. Asci are irregularly spherical, 1–6 spored, sessile.
Additional material examined: CHINA, Yunnan Province, Luquan County, 11 August 2022, collected from P. strobilacea, occasionally on P. armandii, wsp2087-1 (YNAU1264). GenBank: ITS = OR250185; LSU = OR661839; tef1-α = OR832382; rpb2 = OR832414.
Figure 4. Tuber microexcavatum (YNAU1263, holotype). (A) Ascoma and gleba in cross-section; (B) Warts on surface ascoma; (C) Close-up to gleba; (D,E) Peridium hyphal arrangement; (F) Asci and ascospores under bright field microscopy. The scale bars are individually indicated for each image.
Figure 4. Tuber microexcavatum (YNAU1263, holotype). (A) Ascoma and gleba in cross-section; (B) Warts on surface ascoma; (C) Close-up to gleba; (D,E) Peridium hyphal arrangement; (F) Asci and ascospores under bright field microscopy. The scale bars are individually indicated for each image.
Jof 10 00640 g004

3.2.4. Tuber pseudobrumale Y. Wang and Shu H. Li, Mycol. Prog. 2014, 13, 1157–1163 (Figure 5)

Description: Ascomata have a distinct cavity, measuring 1.3–2.0 cm in diameter. The surface is covered with sharp brownish-yellow cones. The peridium is composed of two layers, pseudoparenchymatous; the outer layer is 49.3–112.7 μm thick, composed of subglobose to subangular cells, 4.6–28.6 × 3.1–9.6 μm, that are yellow, pale brown or hyaline; the inner layer 24.9–71.6 μm thick, composed of intricately interwoven hyaline and thin-walled hyphae, 0.4–1.9 μm in diameter. The gleba is solid, yellowish brown when mature, and marbled with white veins. It is composed of hyaline interwoven thin-walled hyphae, 0.6–1.6 μm at the septa, with cells are cylindrical, interwoven, or inflated, 3.1–28.4 × 2.2–13.6 μm. Asci are irregularly spherical, 49.5–63.9 × 32.0–59.9 μm, with each containing 1–7(–8) spores (n = 115), sessile. Ascospores are brownish-yellow, mainly long ellipsoids, though occasionally they are subglobose or long shuttle-shaped and have spines of varying lengths. Spore sizes are as follows: in 1-spored asci: (29.5–)30.2–38.1(–38.8) × (17.3–)19.6–25.2(–26.1) μm, Q = 1.1–1.9, Qm = 1.6 ± 0.16, spines = 1.0–6.1 μm (n = 40); in 2-spored asci: (24.4–)25.9–31.9(–32.7) × (15.1–)15.7–20.4(–20.9) μm, Q = 1.4–2.0, Qm = 1.6 ± 0.13, spines = 1.2–5.9 μm (n = 40); in 3-spored asci: (20.0–)20.6–27.6(–29.9) × (12.4–)14.2–17.9(–19.0) μm, Q = 1.3–1.9, Qm = 1.6 ± 0.14, spines = 1.2–5.5 μm (n = 60); in 4-spored asci: (16.6–)19.1–26.1(–27.9) × (10.1–)13.5–17.0(–17.3) μm, Q = 1.1–2.0, Qm = 1.5 ± 0.14, spines = 1.2–5.5 μm (n = 60); in 5-spored asci: (16.3–)19.3–24.3(–26.4) × (11.6–)12.0–15.9(–16.3) μm, Q = 1.2–1.9, Qm = 1.5 ± 0.14, spines = 1.2–5.0 μm (n = 60); in 6-spored asci: (16.0–)17.3–22.8(–25.0) × (11.3–)11.9–15.1(–15.7) μm, Q = 1.2–1.7, Qm = 1.5 ± 0.12, spines = 1.2–5.0 μm (n = 72); in 7-spored asci: (15.1–)16.9–21.8(–23.3) × (9.6–)10.4–13.9(–16.0) μm, Q = 1.2–1.8, Qm = 1.5 ± 0.12, spines = 0.7–4.4 μm (n = 56); and in 8-spored asci: (15.9–)18.2–21.9(–22.7) × (10.5–)12.4–15.1(–15.7) μm, Q = 1.3–1.8, Qm = 1.5 ± 0.19, spines = 1.8–5.0 μm (n = 1), spino-reticulate, reticulum with 6–9 meshes along the spore length and 6–8 across.
Specimens examined: CHINA, Yunnan Province, 26 November 2020, collected from Pinus sp., wsp991 (YNAU0126). GenBank: ITS = OR665398; LSU = OR825717; tef1-α = OR832383; rpb2 = OR832415; ibid., wsp992 (YNAU0127). GenBank: ITS = OR665399; LSU = OR661817; tef1-α = OR832384; rpb2 = OR832416; ibid., wsp993 (YNAU0128). GenBank: ITS = OR665400; LSU = OR661818; tef1-α = OR832385; rpb2 = OR832417; ibid., Weixi County, 5 December 2020, collected from Pinus sp., wsp1005 (YNAU0145). GenBank: ITS = OR665401; LSU = OR661819; tef1-α = OR832386; rpb2 = OR832418; ibid., wsp1009 (YNAU0148). GenBank: ITS = OR665402; LSU = OR661820; tef1-α = OR832387; rpb2 = OR832419; ibid., wsp1010 (YNAU0149). GenBank: ITS = OR665403; LSU = OR661821; tef1-α = OR832388; rpb2 = OR832420; ibid., wsp1011 (YNAU0150). GenBank: ITS = OR665404; LSU = OR661822; tef1-α = OR832389; rpb2 = OR832421; ibid., Baoshan City, 5 December 2020, collected from Pinus sp., wsp1012 (YNAU0151). GenBank: ITS = OR665405; LSU = OR661823; tef1-α = OR832390; rpb2 = OR832422; ibid., wsp1013 (YNAU0152). GenBank: ITS = OR665406; LSU = OR661824; tef1-α = OR832391; rpb2 = OR832423; ibid., Baoshan City, 20 December 2020, collected from Pinus sp., wsp1036 (YNAU0179). GenBank: ITS = OR665407; LSU = OR661825; tef1-α = OR832392; rpb2 = OR832424; ibid., wsp1037 (YNAU0180). GenBank: ITS = OR665408; LSU = OR661826; tef1-α = OR832393; rpb2 = OR832425; ibid., wsp1038 (YNAU0181). GenBank: ITS = OR665409; LSU = OR661827; tef1-α = OR832394; rpb2 = OR832426; ibid., Xiangyun County, 20 December 2020, collected from Pinus sp., wsp1052 (YNAU0221). GenBank: ITS = OR665410; LSU = OR661828; tef1-α = OR832395; rpb2 = OR832427; ibid., wsp1059 (YNAU0228). GenBank: ITS = OR665411; LSU = OR661829; tef1-α = OR832396; rpb2 = OR832428; ibid., wsp1060 (YNAU0229). GenBank: ITS = OR665412; LSU = OR661830; tef1-α = OR832397; rpb2 = OR832429; ibid., wsp1061 (YNAU0230). GenBank: ITS = OR665413; LSU = OR661831; tef1-α = OR832398; rpb2 = OR832430; ibid., wsp1064 (YNAU0233). GenBank: ITS = OR665414; LSU = OR661832; tef1-α = OR832399; rpb2 = OR832431; ibid., wsp1071 (YNAU0240). GenBank: ITS = OR665415; LSU = OR661833; tef1-α = OR832400; rpb2 = OR832432; ibid., Weixi County, 2 January 2021, collected from Pinus sp., wsp1107 (YNAU0275). GenBank: ITS = OR665416; LSU = OR661834; tef1-α = OR832401; rpb2 = OR832433; ibid., wsp1108 (YNAU0276). GenBank: ITS = OR665417; LSU = OR661835; tef1-α = OR832402; rpb2 = OR832434; ibid., wsp1109 (YNAU0277). GenBank: ITS = OR665418; LSU = OR661836; tef1-α = OR832403; rpb2 = OR832435; ibid., Weixi County, 31 October 2021, collected from Pinus sp., wsp1737 (YNAU0885). GenBank: ITS = OR665419; LSU = OR661837; tef1-α = OR832404; rpb2 = OR832436.
Figure 5. Tuber pseudobrumale (YNAU0221). (A) Ascomata and gleba in cross-section; (B) Warts on surface ascoma; (C) Close-up to gleba; (D) Peridium hyphal arrangement; (E) Asci and ascospores under bright field microscopy; (F,G) Ascospores under scanning electron microscopy. The scale bars are individually indicated for each image.
Figure 5. Tuber pseudobrumale (YNAU0221). (A) Ascomata and gleba in cross-section; (B) Warts on surface ascoma; (C) Close-up to gleba; (D) Peridium hyphal arrangement; (E) Asci and ascospores under bright field microscopy; (F,G) Ascospores under scanning electron microscopy. The scale bars are individually indicated for each image.
Jof 10 00640 g005

3.2.5. Tuber variabilisporum L. Fan and T. Li, Persoonia-Molecular Phylogeny and Evolution of Fungi. 2022, 48(1): 175–202 (Figure 6)

Notes: In the original description, Fan et al. [6] pointed out that “T. variabilisporum is characterized by its … ascomata without basal cavity…”; however, morphological analyses from four sampled specimens identified molecularly as T. variabilisporum (Figure 6A–D), show noticeable depressions on three ascomata (Figure 6A,B,D) and the fourth specimen exhibited less pronounced depression (Figure 6C). Thus, it can be inferred that the surface of the T. variabilisporum is generally depressed but the possibility that some specimens show only a slight depression can occur. Molecular analyses confirm that our specimens and the holotype clustered into a single species branch with high support (BS = 100, PP = 1.0).
Specimens examined: CHINA, Yunnan Province, Dali City, 5 December 2020, collected from Pinus sp., wsp1007-1 (YNAU0146). GenBank: ITS = PP784759; LSU = PP784585; tef1-α = PP796847; rpb2 = PP796857; ibid., Kunming City, 11 September 2021, collected from Pinus sp., wsp1346-1 (YNAU0468). GenBank: ITS = PP784760; LSU = PP784586; tef1-α = PP796848; rpb2 = PP796858; ibid., Lijiang City, 24 November 2021, collected from Q. glauca, wsp1778 (YNAU0925). GenBank: ITS = PP784761; LSU = PP784587; tef1-α = PP796849; rpb2 = PP796859; ibid., Chuxiong City, 12 December 2022, collected from Pinus sp., wsp2471 (YNAU1670). GenBank: ITS = PP784762; LSU = PP784588; tef1-α = PP796850; rpb2 = PP796860.

3.2.6. Tuber pseudohimalayense Moreno G, Díez M and García-Moreno, Mycotaxon. 1997, 63: 217–224 (Figure 6)

Notes: Moreno et al. [36] initially described the presence of uncertain cavities on the surface of the ascomata of T. pseudohimalayense. Wang et al. [48] later described cavities on the ascomata of T. pseudoexcavatum. Later on, Chen et al. [39] confirmed that T. pseudoexcavatum is a synonym of T. pseudohimalayense. In the present study, we observed clear cavities on the ascomata of all six analyzed specimens of T. pseudohimalayense (Figure 6E–J), further confirming the presence of surface depressions. Moreover, specimens of T. pseudohimalayense collected in this study clustered with the type specimen into a single species with high support (BS = 100, PP = 1.0). Additionally, they were grouped with Chinese T. pseudobrumale, T. microexcavatum, T. variabilisporum, and Japanese Tuber sp. 5, all of which exhibit concave surfaces (Figure 1).
Specimens examined: CHINA, Yunnan Province, Kunming City, 8 September 2022, collected from Pinus sp., wsp2167 (YNAU1344). GenBank: ITS = PP784768; LSU = PP784594; tef1-α = PP796856; rpb2 = PP796866; ibid., 22 November 2022, collected from P. armandii, wsp2411 (YNAU1608). GenBank: ITS = PP784763; LSU = PP784589; tef1-α = PP796851; rpb2 = PP796861; ibid., Chuxiong City, 12 December 2022, collected from Pinus sp., wsp2465 (YNAU1663). GenBank: ITS = PP784764; LSU = PP784590; tef1-α = PP796852; rpb2 = PP796862; ibid., wsp2466 (YNAU1664). GenBank: ITS = PP784765; LSU = PP784591; tef1-α = PP796853; rpb2 = PP796863; ibid., wsp2467 (YNAU1665). GenBank: ITS = PP784766; LSU = PP784592; tef1-α = PP796854; rpb2 = PP796864; ibid., wsp2467-1 (YNAU1666). GenBank: ITS = PP784767; LSU = PP784593; tef1-α = PP796855; rpb2 = PP796865.
Figure 6. Ascomata of T. variabilisporum (AD) and T. pseudohimalayense (EJ), showing cavities in their surfaces, with the exception of (E), which shows a less pronounced depression. (A) YNAU0146; (B) YNAU0468; (C) YNAU0925; (D) YNAU1670; (E) YNAU1664; (F) YNAU1608; (G) YNAU1663; (H) YNAU1665; (I) YNAU1666; (J) YNAU1344. (Scale bars = 1 cm).
Figure 6. Ascomata of T. variabilisporum (AD) and T. pseudohimalayense (EJ), showing cavities in their surfaces, with the exception of (E), which shows a less pronounced depression. (A) YNAU0146; (B) YNAU0468; (C) YNAU0925; (D) YNAU1670; (E) YNAU1664; (F) YNAU1608; (G) YNAU1663; (H) YNAU1665; (I) YNAU1666; (J) YNAU1344. (Scale bars = 1 cm).
Jof 10 00640 g006

4. Discussion

The description and study of black truffle species began in the early 19th century. Over time, research has gradually deepened, encompassing various aspects such as morphology, geographical distribution, genetics, and ecology [6,21,22,49,50,51]. Brongniart first introduced the Melanosporum group, characterized by its black or brown appearance and irregular surface texture; some species have distinct basal cavities, releasing unique strong aromas, with two peridium layers. Asci containing multiple spores are usually large and elliptical, with spines of different lengths [5,6,15,16,19,22,50,51]. However, relying solely on morphology and single-gene methods for species identification can lead to subjective biases and ambiguous or even incorrect delineations, especially when the number of samples is small and the sequence qualities are low [52,53]. This is particularly common in truffle species. For example, Asian black truffles such as T. indicum, T. himalayense, T. sinense and T. formosanum are generally poorly defined due to morphological similarities and variable sequence quality [33,54]. Although some studies speculate that T. indicum is T. sinense, this conclusion is not yet supported by molecular evidence from type specimens [6]. Taking into account this scenario, there is an urgent need to carry out an integrative approach combining morphological and molecular information in order to have deeper insights into the Melanosporum group. Therefore, we collected a series of samples from truffles belonging to this latter group and conducted initial classification based on morphological features. Subsequently, we conducted molecular analysis and sequenced four different regions for phylogenetic analysis. As a consequence, in this work, we describe three new species and redefined some other species, validating their taxonomic status. The results indicated that in some cases, the molecular and morphological characterization coincide, further supporting initial classifications. For example, species with depressions belonged to the so-called subgroups 1, 3, and 4, while species without depressions were all clustered in another subgroup called 2. However, in other cases, inconsistencies between morphological and molecular data were observed, suggesting that further insights are needed related to species diversity assessment (e.g., T. pseudobrumale and T. melanoexcavatum). Our study contributes to the knowledge of the taxonomic status of some species by adding the examination of more specimens and their morphological and molecular characterization.
Currently, there are at least 14 species in the Melanosporum group. Our phylogenetic tree based on a concatenated multilocus dataset revealed that the species belonging to this group are nested in four phylogenetic clades or subgroups. Subgroup 1 includes six species, T. pseudobrumale, T. melanoexcavatum, T. microexcavatum sp. nov., T. pseudohimalayense, T. variabilisporum, and Tuber sp. 5. Tuber microexcavatum sp. nov. was molecularly identified as a new species on the basis of less than 80% ITS similarity to any known species. Although T. microexcavatum has lower maturity spores, it can also be distinguished from the closely related species T. pseudobrumale by ascospore size, surface warts, and spore number.
We also re-evaluated and delimited the phylogenetic relationship between T. pseudobrumale and T. melanoexcavatum in subgroup 1. Molecularly, the types of T. pseudobrumale and T. melanoexcavatum and 22 examined specimens in this work, are highly clustered into a single branch, and the ITS similarities of both species sequences were greater than 97.9%.
Based on molecular data analysis, T. pseudobrumale and T. melanoexcavatum are closely related phylogenetically. ITS rDNA sequence analysis showed that these species have similarities ranging from 98.0% to 99.7%. Morphologically, these two species share common features in the original descriptions, such as the ellipsoidal shape of the spores. The only difference is the number of spores within the asci. Tuber pseudobrumale has ellipsoidal spores (Q = 1.27) and 3–7 spored asci according to Li et al. [42], while T. melanoexcavatum has 5–8 spored asci, and its spores are also ellipsoidal (Q = 1.19) [6]. It is noteworthy that T. pseudobrumale is assigned only one Q value (1.27) in the aforementioned article, whereas T. melanoexcavatum has also only one Q value (1.19). This later Q value should correspond to subglobose spores, however, in their original description, Wang et al. (2020) reported spore sizes as follows “… ascospores in 5-spored asci 22–24.7 × 15.4–16.9 μm, 6-spored asci 21.4–24 × 14.3–16.0 μm, 7-spored asci 20–22 × 13.5–15.5 μm, 8-spored asci 18.7–21.2 × 12.6–15.1 μm …”, which indicates that the spores were ellipsoid rather than subglobose. In order to elucidate whether T. pseudobrumale and T. melanoexcavatum may be different or the same species, we incorporated our own collected specimens with a detailed morphological characterization and analyzed previously published information. As a result, we were able to understand the morphological characteristics of T. pseudobrumale more comprehensively (Figure 5), including spore shape and Q value, which could adequately cover the aforementioned features. Based on these morphological, molecular, and phylogenetic analyses, it is possible to conclude that T. melanoexcavatum is a synonym of T. pseudobrumale.
Another two new species, T. yunnanense and T. melanoumbilicatum, were identified within subgroups 2 and 4, respectively. Phylogenetically, T. yunnanense belongs to subgroup 2, a clade that includes T. yigongense, T. sinense, T. formosanum, T. longispinosum, T. melanosporum, and T. regimontanum; they have pyramidal warts and no cavities in their ascomata. The ITS similarities between T. yunnanense and species of subgroup 2 were 87.0–93.0%. Morphologically, T. yunnanense is similar to T. longispinosum in the number of spores per asci having 1–5(–6) spiny spores, with a few asci possessing short stalks. However, they differ in spore color and size. The color of T. yunnanense is golden, whereas that of T. longispinosum is brown to dark brown. T. yunnanense spores are (20–)22–44(–55) × (12–)14–29(–35) μm, Q = 1.1–2.2, and those of T. longispinosum are (15–)21–35(–41) × (12–)15–26(–30) μm, Q = 1.0–2.1 [32]. The spores of T. yunnanense are mainly long shuttle-shaped (Qm = 1.74 ± 0.1), occasionally subglobose, whereas those of T. longispinosum are mainly ellipsoid, occasionally globose, and long shuttle-shaped. Subgroups 3 and 4 include T. brumale and T. melanoumbilicatum, respectively, both of which had concave ascomata. Molecular analysis reveals that the ITS similarity between these two species is 70%. In morphology, T. melanoumbilicatum resembles T. pseudobrumale in having a distinct basal cavity on the surface of ascomata and asci with multiple spores [41]. However, they differ significantly in spore shape. The spores of T. melanoumbilicatum are long oval with Qm = 1.65 ± 0.1, whereas those of T. pseudobrumale are ellipsoidal with Qm =1.43 ± 0.2.
The presence of 14 phylogenetic species in the Melanosporum group was confirmed by the phylogenetic tree based on the ITS, LSU, tef1-α, and rpb2 dataset (Figure 1). The Melanosporum group currently comprises 16 species, with 14 of them well supported by reliable molecular data, 13 of which are well-defined, and the status of the other three needs to be clarified by studying more collections, conducting more detailed morphological and molecular analyses. Among these, the species in subgroups 1, 3 and 4 all have distinct concavities. In the genus Tuber, the Melanosporum and Excavatum group have distinct cavities on the surface of the ascomata. This special surface structure may reflect the specialized adaptations of the Melanosporum and Excavatum groups in relation to their life histories and ecological environments [55]. Species of the Melanosporum and Excavatum groups, which typically grow in soil, have adapted to different habitat conditions. This adaptation likely involves the regulation of gas exchange, water uptake, and nutrient acquisition through their cavity structures [16]. Such morphological features may have gradually evolved, providing these fungi with effective mechanisms for survival and reproduction in environments with different conditions. Further studies on these cavity structures could reveal the biological characteristics and adaptive mechanisms of species of the Melanosporum and Excavatum groups. This research thus provides new insights and perspectives for the study of fungal taxonomy and ecology of the Melanosporum group of truffle species.
Key to the taxa of Melanosporum group
1. Ascomata has a cavity2
1. Ascomata no a cavity3
2. Ascomata yellowish brownT. microexcavatum
2. Ascomata black4
3. Spore shape, Q > 1.75
3. Spore shape, Q < 1.77
4. Spore shape, broadly ellipsoidal or ellipsoidal6
4. Spore shape, long ellipsoid8
5. Spore shape, mainly long shuttle-shapedT. yunnanense
5. Spore shape, mainly ellipsoid9
6. Spore shape, broadly ellipsoidalT. pseudohimalayense
6. Spore shape, mainly ellipsoid10
7. Height of spore ornamentation, 3–6(–7) µm, spinyT. sinense
7. Height of spore ornamentation, 2–5(–6) µm, spino-reticulateT. formosanum
8. Height of spore ornamentation, 3.4 ± 0.38 µm, spino-reticulateT. brumale
8. Height of spore ornamentation, 0.9–8.2 µm, spino-reticulateT. melanoumbilicatum
9. Height of spore ornamentation, spinyT. longispinosum
9. Height of spore ornamentation, spino-reticulate11
10. Height of spore ornamentation, 0.4–9.6 µm, spino-reticulateT. pseudobrumale
10. Height of spore ornamentation, spino-reticulate12
11. Height of spore ornamentation, 2.5–4 µm, densely spino-reticulateT. yigongense
11. Height of spore ornamentation, spino-reticulate13
12. Spore length and width, 15–20 × 10–15 μmTuber sp. 5
12. Spore length and width, 16.5–37.5 × 13.25–27.5 μmT. variabilisporum
13. Spore length and width, 28–55 × 18–31 μmT. regimontanum
13. Spore length and width, 28–32 × 16–21 μmT. melanosporum

Author Contributions

Writing—original draft preparation, R.W. (Rui Wang) and S.W.; writing—review and editing, F.Y., S.W. and J.P.-M.; methodology, Y.L., R.W. (Ruixue Wang), S.Y. and J.Y. (Jing Yuan); software, X.X., X.S. and J.Y. (Juanbing Yu); validation, F.Y. and S.W.; formal analysis, R.W. (Rui Wang), G.D., F.Y. and S.W.; data curation, R.W. (Rui Wang) and G.D.; visualization, R.W. (Rui Wang), Y.L. and R.W. (Ruixue Wang); supervision, F.Y.; project administration, S.W. and G.D.; funding acquisition, S.W. and G.D. All authors have read and agreed to the published version of the manuscript.

Funding

The work was financially supported by the National Natural Science Foundation of China (No. 32060008), the Basic Research Program of Yunnan (202201AT070268), the program “Evaluation of Tuber germplasm and development of mycorrhiza synthesis technology for new species (Am20230400BC)”, and the Yunnan Technology Innovation Program (202205AD160036).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

A publicly available dataset was analyzed in this study. The resulting alignments were deposited in TreeBASE (http://www.treebase.org; accession number 31641 (accessed on 22 August 2024))). All newly generated sequences were deposited in GenBank (https://www.ncbi.nlm.nih.gov/genbank/ (accessed on 12 October 2023), mentioned in the text in Table 1 and Table A1 and in Figure 1). All new taxa were deposited in MycoBank (https://www.mycobank.org/ (accessed on 21 May 2024)).

Acknowledgments

We are thankful to Zhijia Gu of the Key Laboratory for Plant Diversity and Biogeography of East Asia, Chinese Academy of Sciences, for the work of scanning electron microscopy (SEM). We are thankful to Chengjin Yu, Lei Cao, and Duancong Li for their kind help in this study. Jesús Pérez-Moreno acknowledges the support from the High-end Foreign Expert Project of Yunnan and Colegio de Postgraduados, Mexico.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

Appendix A

Table A1. Species, lineage voucher specimen, country, and corresponding GenBank accessions of the Tuber specimens used for the phylogenetic analysis.
Table A1. Species, lineage voucher specimen, country, and corresponding GenBank accessions of the Tuber specimens used for the phylogenetic analysis.
Taxon NameLineageVoucher SpecimenCountryITSLSUtef1rpb2
Choiromyces sichuanensisoutgroupYNAU003, holotypeSichuan, ChinaMW380902OK576632PP092037OR832439
Choiromyces sichuanensisoutgroupYNAU004Sichuan, ChinaOK585070OK576633PP092038OR832440
Tuber aestivumaestivumGB202ItalyJQ925679JX022565JQ954487
Tuber aestivumaestivumJT30500SwedenHM485340JQ954488
Tuber borchiipuberulumBJTC FAN217New ZealandKT067681KT067706KT067717OM584229
Tuber borchiipuberulumGB1/GB32ItalyFJ809852FJ809799JX022571JQ954492
Tuber brumalemelanosporumGB52ItalyHM485345JQ925683JQ954494
Tuber brumalemelanosporumGB53ItalyFJ748900JQ925684JQ954495
Tuber californicumpuberulumJT28058USAHM485346JQ925685JX022574JQ954496
Tuber canaliculatummacrosporumJT28215USAJQ925643JX022575JQ954497
Tuber canaliculatummacrosporumOSC59072USAHM485347JX022576JQ954498
Tuber crassitunicatumrufumBJTC FAN465, holotypeYunnan, ChinaMH115295OM366205OM649610OM584268
Tuber depressumexcavatumBJTC FAN340Sichuan, ChinaOM256744OM366187OM649592OM584250
Tuber depressumexcavatumBJTC FAN534Sichuan, ChinaOM256764OM366211OM649616
Tuber dryophilumpuberulumGB35ItalyJQ925644JQ925687JX022577
Tuber dryophilumpuberulumGB37ItalyHM485354JQ925688JX022578JQ954501
Tuber formosanummelanosporumBJTC FAN107Yunnan, ChinaMF621549OM366159OM649564OM584210
Tuber formosanummelanosporumBJTC FAN356Sichuan, ChinaMF627986OM366189OM649594OM584252
Tuber glabrummacrosporumBJTC FAN228, holotypeYunnan, ChinaKF002731OM366177OM649581OM584234
Tuber glabrummacrosporumBJTC FAN232, paratypeYunnan, ChinaKF002727OM366179OM649583OM584236
Tuber huidongenserufumBJTC FAN104Yunnan, ChinaJF921163OM366158OM649563OM584209
Tuber huidongenserufumBJTC FAN101Yunnan, ChinaOM311172OM366156OM649562OM584208
Tuber jinshajiangensepuberulumBJTC FAN406Yunnan, ChinaKX575841OM366199OM649604OM584262
Tuber jinshajiangensepuberulumBJTC FAN407Yunnan, ChinaKX575842OM366200OM649605OM584263
Tuber liaotongenserufumBJTC FAN550Beijing, ChinaMH115302OM366213OM649618OM584272
Tuber lijiangensepuberulumBJTC FAN307Yunnan, ChinaKP276188KP276203KP276206OM584244
Tuber lishanenserufumBJTC FAN718, holotypeShanxi, ChinaMH115303MH115304OM649622OM584276
Tuber lishanenserufumBJTCFAN683Shanxi, ChinaMH115305MH115306OM649621OM584275
Tuber longispinosummelanosporumK225 JapanAB553414AB553518AB553538AB553558
Tuber longispinosummelanosporumK447, holotypeJapanAB553423
Tuber macrosporummacrosporumJT13362ItalyHM485373FJ809838JX022590
Tuber magnatumaestivumGB12ItalyJQ925645JQ925700JX022591JQ954512
Tuber magnatumaestivumGB13ItalyJQ925646JQ925701JX022592JQ954513
Tuber melanoexcavatummelanosporumYAAS L3605, holotypeYunnan, ChinaKY081684
Tuber melanoexcavatummelanosporumYAAS L3636Yunnan, ChinaKY081685
Tuber melanosporummelanosporumGB200ItalyFJ748904JQ925703JX022594JQ954515
Tuber melanoumbilicatummelanosporumYNAU017, holotypeYunnan, ChinaOK625304OR661815OR832379OR832411
Tuber melanoumbilicatummelanosporumYNAU018Yunnan, ChinaOK625305OR661816OR832380OR832412
Tuber microexcavatummelanosporumYNAU1263Yunnan, ChinaOR250184OR661838OR832381OR832413
Tuber microexcavatummelanosporumYNAU1264Yunnan, ChinaOR250185OR661839OR832382OR832414
Tuber neoexcavatumexcavatumBJTC FAN184, holotypeYunnan, ChinaJX458715OM366169OM649574
Tuber neoexcavatumexcavatumBJTC FAN316Yunnan, ChinaOM256741OM366184OM649589OM584247
Tuber oligospermumpuberulumAH38984SpainJN392261JN392320
Tuber oligospermumpuberulumAH37783SpainJN392260JN392324
Tuber pseudobrumalemelanosporumYAAS L3181, holotypeYunnan, ChinaKJ742703
Tuber pseudobrumalemelanosporumBJTC FAN306Yunnan, ChinaOM287838OM366183OM649587OM584243
Tuber pseudobrumalemelanosporumBJTC FAN322Yunnan, ChinaOM287839OM366186OM649591OM584249
Tuber pseudobrumalemelanosporumYNAU0126Yunnan, ChinaOR665398OR825717OR832383OR832415
Tuber pseudobrumalemelanosporumYNAU0127Yunnan, ChinaOR665399OR661817OR832384OR832416
Tuber pseudobrumalemelanosporumYNAU0128Yunnan, ChinaOR665400OR661818OR832385OR832417
Tuber pseudobrumalemelanosporumYNAU0145Yunnan, ChinaOR665401OR661819OR832386OR832418
Tuber pseudobrumalemelanosporumYNAU0148Yunnan, ChinaOR665402OR661820OR832387OR832419
Tuber pseudobrumalemelanosporumYNAU0149Yunnan, ChinaOR665403OR661821OR832388OR832420
Tuber pseudobrumalemelanosporumYNAU0150Yunnan, ChinaOR665404OR661822OR832389OR832421
Tuber pseudobrumalemelanosporumYNAU0151Yunnan, ChinaOR665405OR661823OR832390OR832422
Tuber pseudobrumalemelanosporumYNAU0152Yunnan, ChinaOR665406OR661824OR832391OR832423
Tuber pseudobrumalemelanosporumYNAU0179Yunnan, ChinaOR665407OR661825OR832392OR832424
Tuber pseudobrumalemelanosporumYNAU0180Yunnan, ChinaOR665408OR661826OR832393OR832425
Tuber pseudobrumalemelanosporumYNAU0181Yunnan, ChinaOR665409OR661827OR832394OR832426
Tuber pseudobrumalemelanosporumYNAU0221Yunnan, ChinaOR665410OR661828OR832395OR832427
Tuber pseudobrumalemelanosporumYNAU0228Yunnan, ChinaOR665411OR661829OR832396OR832428
Tuber pseudobrumalemelanosporumYNAU0229Yunnan, ChinaOR665412OR661830OR832397OR832429
Tuber pseudobrumalemelanosporumYNAU0230Yunnan, ChinaOR665413OR661831OR832398OR832430
Tuber pseudobrumalemelanosporumYNAU0233Yunnan, ChinaOR665414OR661832OR832399OR832431
Tuber pseudobrumalemelanosporumYNAU0240Yunnan, ChinaOR665415OR661833OR832400OR832432
Tuber pseudobrumalemelanosporumYNAU0275Yunnan, ChinaOR665416OR661834OR832401OR832433
Tuber pseudobrumalemelanosporumYNAU0276Yunnan, ChinaOR665417OR661835OR832402OR832434
Tuber pseudobrumalemelanosporumYNAU0277Yunnan, ChinaOR665418OR661836OR832403OR832435
Tuber pseudobrumalemelanosporumYNAU0885Yunnan, ChinaOR665419OR661837OR832404OR832436
Tuber pseudofulgensexcavatumBJTC FAN399, holotypeYunnan, ChinaOM256757OM366196OM649601OM584259
Tuber pseudofulgensexcavatumBJTCFAN388Sichuan, ChinaOM256755OM366194OM649599OM584257
Tuber pseudohimalayensemelanosporumYNAU1608Yunnan, ChinaPP784763PP784589PP796851PP796861
Tuber pseudohimalayensemelanosporumBJTC FAN122Sichan, ChinaMF627983OM366162OM649567OM584213
Tuber pseudohimalayensemelanosporumYNAU1663Yunnan, ChinaPP784764PP784590PP796852PP796862
Tuber pseudohimalayensemelanosporumYNAU1664Yunnan, ChinaPP784765PP784591PP796853PP796863
Tuber pseudohimalayensemelanosporumYNAU1665Yunnan, ChinaPP784766PP784592PP796854PP796864
Tuber pseudohimalayensemelanosporumYNAU1666Yunnan, ChinaPP784767PP784593PP796855PP796865
Tuber pseudohimalayensemelanosporumYNAU1344Yunnan, ChinaPP784768PP784594PP796856PP796866
Tuber puberulumpuberulumZB1433HungaryJF261382JF261346
Tuber puberulumpuberulumZB1983 HungaryJF261390JF261357
Tuber regimontanummelanosporumITCV 909MexicoEU375838FJ809823JX022600JQ954520
Tuber shidianensepuberulumHKAS 88770, holotypeYunnan, ChinaKT444595KY174960OR832405OR832437
Tuber shidianensepuberulumHKAS 88771, paratypeYunnan, ChinaKT444596KY174961OR832406OR832438
Tuber sinensemelanosporumBJTC FAN108ChinaMF627968OM366160OM649565OM584211
Tuber sinensemelanosporumBJTC FAN110ChinaMF627970OM366161OM649566OM584212
Tuber sinoaestivumaestivumBJTC FAN522Sichuan, ChinaOM256774OM366210OM649615OM584271
Tuber sinoaestivumaestivumBJTC FAN487Yunnan, ChinaOM256773OM366209OM649614OM584270
Tuber sinoexcavatumexcavatumBJTC FAN130, holotypeYunnan, ChinaJX458717OM366163OM649568OM584216
Tuber sinoexcavatumexcavatumBJTC FAN166Yunnan, ChinaJX458718OM366165OM649571OM584221
Tuber sp. 1macrosporumK201JapanAB553344AB553512AB553532AB553552
Tuber sp. 5 KA-2010melanosporumK229JapanAB553381AB553536AB553556
Tuber sp. 5 KA-2010melanosporumK403JapanAB553382
Tuber sphaerosporumpuberulumJT12487USAFJ809853FJ809853JX022609
Tuber taiyuanenserufumBJTC FAN164Yunnan, ChinaOM311182OM366164OM649570OM584220
Tuber taiyuanenserufumBJTC FAN220Yunnan, ChinaMH115315OM366174OM649578OM584231
Tuber umbilicatumrufumBJTC FAN212Yunnan, ChinaOM311201OM366173OM649577OM584228
Tuber umbilicatumrufumBJTC FAN230Yunnan, ChinaOM311205OM366178OM649582OM584235
Tuber variabilisporummelanosporumBJTC FAN362, holotypeSichuan, ChinaOM287845OM366190OM649595OM584253
Tuber variabilisporummelanosporumYNAU0146Yunnan, ChinaPP784759PP784585PP796847PP796857
Tuber variabilisporummelanosporumYNAU0468Yunnan, ChinaPP784760PP784586PP796848PP796858
Tuber variabilisporummelanosporumYNAU0925Yunnan, ChinaPP784761PP784587PP796849PP796859
Tuber variabilisporummelanosporumYNAU1670Yunnan, ChinaPP784762PP784588PP796850PP796860
Tuber wenchuanenserufumBJTC FAN833Shanxi, ChinaOM311256OM366222OM649629OM584280
Tuber yigongensemelanosporumBJTC FAN729Tibet, ChinaMF663716OM649623OM584277
Tuber yigongensemelanosporumBJTC FAN731, holotypeTibet, ChinaMF663714OM366216
Tuber yunnanensemelanosporumYNAU019, holotypeYunnan, ChinaOK625306OR661811OR813081OR832407
Tuber yunnanensemelanosporumYNAU020Yunnan, ChinaOK625307OR661812OR813082OR832408
Tuber yunnanensemelanosporumYNAU0107Yunnan, ChinaOR665397OR661813OR813083OR832409
Tuber yunnanensemelanosporumYNAU0491Sichuan, ChinaOR250186OR661814OR813084OR832410

References

  1. Luo, Q.; Zhang, J.; Yan, L.; Tang, Y.; Ding, X.; Yang, Z.; Sun, Q. Composition and antioxidant activity of water-soluble polysaccharides from Tuber indicum. J. Med. Food 2011, 14, 1609–1616. [Google Scholar] [CrossRef] [PubMed]
  2. Splivallo, R.; Ottonello, S.; Mello, A.; Karlovsky, P. Truffle volatiles: From chemical ecology to aroma biosynthesis. New Phytol. 2011, 189, 688–699. [Google Scholar] [CrossRef] [PubMed]
  3. Li, S.N.; Li, X.A.; Zhang, Q.; Hu, Y.J.; Lei, H.R.; Guo, D.L.; Jiang, L.S.; Deng, Y. Chemical constitutes from Tuber indicum with immunosuppressive activity uncovered by transcriptome analysis. Fitoterapia 2024, 173, 105773. [Google Scholar] [CrossRef] [PubMed]
  4. Zambonelli, A.; Iotti, M.; Murat, C. True Truffle (Tuber spp.) in the World; Springer International Publishing: Berlin/Heidelberg, Germany, 2016; pp. 19–104. [Google Scholar]
  5. Bonito, G.; Smith, M.E.; Nowak, M.; Healy, R.A.; Guevara, G.; Cázares, E.; Kinoshita, A.; Nouhra, E.R.; Domínguez, L.S.; Tedersoo, L.; et al. Historical biogeography and diversification of truffles in the Tuberaceae and their newly identified southern hemisphere sister lineage. PLoS ONE 2013, 8, e52765. [Google Scholar] [CrossRef]
  6. Fan, L.; Li, T.; Xu, Y.Y.; Yan, X.Y. Species diversity, phylogeny, endemism and geography of the truffle genus Tuber in China based on morphological and molecular data. Persoonia-Mol. Phylogeny Evol. Fungi 2022, 48, 175–202. [Google Scholar] [CrossRef]
  7. Hall, I.; Brown, G.; Zambonelli, A. Taming the Truffle. The History, Lore, and Science of the Ultimate Mushroom; Timber Press: Portland, OR, USA, 2007; pp. 17–39. [Google Scholar]
  8. Murat, C.; Payen, T.; Noel, B.; Kuo, A.; Morin, E.; Chen, J.; Kohler, A.; Krizsán, K.; Balestrini, R.; Da Silva, C.; et al. Pezizomycetes genomes reveal the molecular basis of ectomycorrhizal truffle lifestyle. Nat. Ecol. Evol. 2018, 2, 1956–1965. [Google Scholar] [CrossRef]
  9. Schneider-Maunoury, L.; Leclercq, S.; Clément, C.; Covès, H.; Lambourdière, J.; Sauve, M.; Richard, F.; Selosse, M.A.; Taschen, E. Is Tuber melanosporum colonizing the roots of herbaceous, non-ectomycorrhizal plants? Fungal Ecol. 2018, 31, 59–68. [Google Scholar] [CrossRef]
  10. Trappe, J.M. The orders, families, and genera of hypogeous Ascomycotina (truffels and their relatives). Mycotaxon 1979, 9, 297–340. [Google Scholar]
  11. Antony-Babu, S.; Deveau, A.; Van Nostrand, J.D.; Zhou, J.; Le Tacon, F.; Robin, C.; Frey-Klett, P.; Uroz, S. Black truffle-associated bacterial communities during the development and maturation of Tuber melanosporum ascocarps and putative functional roles. Environ. Microbiol. 2014, 16, 2831–2847. [Google Scholar] [CrossRef]
  12. Albornoz, F.E.; Dixon, K.W.; Lambers, H. Revisiting mycorrhizal dogmas: Are mycorrhizas really functioning as they are widely believed to do? Soil Ecol. Lett. 2021, 3, 73–82. [Google Scholar] [CrossRef]
  13. Elliott, T.F.; Truong, C.; Jackson, S.M.; Zúñiga, C.L.; Trappe, J.M.; Vernes, K. Mammalian mycophagy: A global review of ecosystem interactions between mammals and fungi. Fungal Syst. Evol. 2022, 9, 99–159. [Google Scholar] [CrossRef] [PubMed]
  14. Le Tacon, F.; Zeller, B. Dubious intraspecific variability in the Tuber melanosporum genome revealed by sequence analysis of the internal transcribed spacer of the ribosomal DNA. Mycologia 2002, 94, 872–875. [Google Scholar]
  15. Bonito, G.M.; Gryganskyi, A.P.; Trappe, J.M.; Vilgalys, R. A global meta-analysis of Tuber ITS rDNA sequences: Species diversity, host associations and long-distance dispersal. Mol. Ecol. 2010, 19, 4994–5008. [Google Scholar] [CrossRef] [PubMed]
  16. Bonito, G.; Trappe, J.M.; Donovan, S.; Vilgalys, R. The Asian black truffle Tuber indicum can form ectomycorrhizas with North American host plants and complete its life cycle in non-native soils. Fungal Ecol. 2011, 4, 83–93. [Google Scholar] [CrossRef]
  17. Kinoshita, A.; Sasaki, H.; Nara, K. Phylogeny and diversity of Japanese truffles (Tuber spp.) inferred from sequences of four nuclear loci. Mycologia 2011, 103, 779–794. [Google Scholar] [CrossRef]
  18. Le Tacon, F.; Rubini, A.; Murat, C.; Riccioni, C.; Robin, C.; Belfiori, B.; De la Varga, H.; Akroume, E.; Deveau, A.; Martin, F. Certainties and uncertainties about the life cycle of the Périgord black truffle (Tuber melanosporum Vittad.). Ann. For. Sci. 2016, 73, 105–117. [Google Scholar] [CrossRef]
  19. Brongniart, J.L.É.A. Tableau des Champignons Fossiles de l’Ecole de Mines de Paris. [Translation: Table of Fossil Mushrooms from the School of Mines of Paris]; École de Mines de Paris: Paris, France, 1831; p. 24. [Google Scholar]
  20. Mello, A.; Murat, C.; Bonfante, P. Truffles: Much more than a prized and local fungal delicacy. FEMS Microbiol. Lett. 2006, 260, 1–8. [Google Scholar] [CrossRef]
  21. Riccioni, C.; Belfiori, B.; Rubini, A.; Passeri, V.; Arcioni, S.; Paolocci, F. Tuber melanosporum outcrosses: Analysis of the genetic diversity within and among its natural populations under this new scenario. New Phytol. 2008, 180, 466–478. [Google Scholar] [CrossRef]
  22. Vittadini, C. Monographia Tuberacearum; Typographia F. Rusconi: Milan, Italy, 1831; pp. 37–42. [Google Scholar]
  23. Merényi, Z.; Varga, T.; Geml, J.; Orczán, Á.K.; Chevalier, G.; Bratek, Z. Phylogeny and phylogeography of the Tuber brumale aggr. Mycorrhiza 2014, 24, 101–113. [Google Scholar] [CrossRef]
  24. Cooke, M.C.; Massee, G. Himalayan truffles. Grevillea 1892, 20, 67. [Google Scholar]
  25. Zhang, B.C.; Minter, D.W. Tuber himalayense sp. nov. with notes on Himalayan truffles. Trans. Br. Mycol. Soc. 1988, 91, 593–597. [Google Scholar] [CrossRef]
  26. Tao, K.; Liu, B.; Zhang, D.C. A new species of Truffle. J. Shanxi Univ. 1989, 02, 215–218. [Google Scholar]
  27. Hu, H.D. Tuber formosanum (new species) and its mycorrhizal relationship. Taiwan Univ. Exp. For. Res. Rep. 1992, 6, 79–87. [Google Scholar]
  28. Qiao, P.; Liu, P.G.; Hu, H.T.; Wang, Y. Typification of Tuber formosanum (Tuberaceae, Pezizales, Ascomycota) from Taiwan, China. Mycotaxon 2013, 123, 293–299. [Google Scholar] [CrossRef]
  29. Roux, C.; Séjalon-Delmas, N.; Martins, M.; Parguey-Leduc, A.; Dargent, R.; Bécard, G. Phylogenetic relationships between European and Chinese truffles based on parsimony and distance analysis of ITS sequences. FEMS Microbiol. Lett. 1999, 180, 147–155. [Google Scholar] [CrossRef]
  30. Zhang, L.F.; Yang, Z.L.; Song, D.S. A phylogenetic study of commercial Chinese truffles and their allies: Taxonomic implications. FEMS Microbiol. Lett. 2005, 245, 85–92. [Google Scholar] [CrossRef]
  31. Feng, B.; Zhao, Q.; Xu, J.; Qin, J.; Yang, Z.L. Drainage isolation and climate change-driven population expansion shape the genetic structures of Tuber indicum complex in the Hengduan Mountains region. Sci. Rep. 2016, 6, 21811. [Google Scholar] [CrossRef]
  32. Kinoshita, A.; Nara, K.; Sasaki, H.; Feng, B.; Obase, K.; Yang, Z.L.; Yamanaka, T. Using mating-type loci to improve taxonomy of the Tuber indicum complex, and discovery of a new species, T. longispinosum. PLoS ONE 2018, 13, e0193745. [Google Scholar] [CrossRef]
  33. Chen, J.; Guo, S.X.; Liu, P.G. Species recognition and cryptic species in the Tuber indicum complex. PLoS ONE 2011, 6, e14625. [Google Scholar] [CrossRef]
  34. Wang, Y.J.; Tan, Z.M.; Murat, C.; Jeandroz, S.; Le Tacon, F. Phylogenetic and populational study of the Tuber indicum complex. Mycol. Res. 2006, 110, 1034–1045. [Google Scholar] [CrossRef]
  35. Fan, L.; Zhang, J.L.; Li, T.; Sun, H.J.; Xiong, W.P.; Li, Y. Chinese black truffles: Tuber yigongense sp. nov., taxonomic reassessment of T. indicum sl, and re-examination of the T. sinense isotype. Mycotaxon 2018, 133, 183–196. [Google Scholar] [CrossRef]
  36. Moreno, G.; Manjón, J.L.; Díez, J.; García-Montero, L.G.; Di Massimo, G. Tuber pseudohimalayense sp. nov. An Asiatic species commercialized in Spain, similar to the “Perigord” truffle. Mycotaxon 1997, 63, 217–224. [Google Scholar]
  37. Wang, Y.J.; Moreno, G.; Riousset, L.J.; Manjón, J.L.; Riousset, G.; Fourre, G.; Massimo, G.D.; Garcia-Montero, L.G.; Diez, J. Tuber pseudoexcavatum sp. nov. a new species from China commercialised in Spain, France and Italy with additional comments on Chinese truffles. Cryptogam. Mycol. 1998, 19, 113–120. [Google Scholar]
  38. Manjón, J.L.; García-Montero, L.G.; Alvarado, P.; Moreno, G.; Massimo, G.D. Tuber pseudoexcavatum versus T. pseudohimalayense-new data on the molecular taxonomy and mycorrhizae of Chinese truffles. Mycotaxon 2009, 110, 399. [Google Scholar] [CrossRef]
  39. Chen, J.; Liu, P.G. Delimitation of Tuber pseudohimalayense and T. pseudoexcavatum based on morphological and molecular data. Cryptogam. Mycol. 2011, 32, 83–93. [Google Scholar] [CrossRef]
  40. Guevara, G.; Bonito, G.; Cázares, E.; Rodríguez, J.; Vilgalys, R.; Trappe, J.M. Tuber regimontanum, new species of truffle from Mexico. Rev. Mex. Micol. 2008, 26, 17–20. [Google Scholar]
  41. Li, S.H.; Zheng, L.Y.; Liu, C.Y.; Wang, Y.; Li, L.; Zhao, Y.C.; Zhang, X.L.; Yang, M.; Xong, H.K.; Qing, Y.; et al. Two new truffles species, Tuber alboumbilicum and Tuber pseudobrumale from China. Mycol. Prog. 2014, 13, 1157–1163. [Google Scholar] [CrossRef]
  42. Wang, L.; Tang, S.M.; Zhang, X.; Su, K.; Li, Y.; Yuan, T.; Wang, Y.; Dauner, L.; Li, S. Tuber melanoexcavatum sp. nov., an edible black truffle from China. Phytotaxa 2020, 477, 269–276. [Google Scholar] [CrossRef]
  43. Kumar, L.M.; Smith, M.E.; Nouhra, E.R.; Orihara, T.; Leiva, P.S.; Pfister, D.H.; McLaughlin, D.J.; Trappe, J.M.; Healy, R.A. A molecular and morphological re-examination of the generic limits of truffles in the Tarzetta-Geopyxis lineage-Densocarpa, Hydnocystis, and Paurocotylis. Fungal Biol. 2017, 121, 264–284. [Google Scholar] [CrossRef]
  44. Gardes, M.; Bruns, T.D. ITS primers with enhanced specificity for basidiomycetes-application to the identification of mycorrhizae and rusts. Mol. Ecol. 1993, 2, 113–118. [Google Scholar] [CrossRef]
  45. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed]
  46. Hall, T.A. BioEdit: A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp. Ser. 1999, 41, 95–98. [Google Scholar]
  47. Chen, J. Taxonomy and Phylogeny of the Genus Tuber in China. Ph.D. Thesis, Kunming Institute of Botany, Chinese Academy of Sciences, Kunming, China, 2007. [Google Scholar]
  48. Wang, Y.J.; Tan, Z.M.; Zhang, D.C.; Murat, C.; Jeandroz, S.; Le Tacon, F. Phylogenetic relationships between Tuber pseudoexcavatum, a Chinese truffle, and other Tuber species based on parsimony and distance analysis of four different gene sequences. FEMS Microbiol. Lett. 2006, 259, 269–281. [Google Scholar] [CrossRef]
  49. Tulasne, L.R.; Tulasne, C. Fungi Hypogaei; Friedrich Kfincksieck: Paris, France, 1851; pp. 19–22. [Google Scholar]
  50. Tulasne, L.R.; Tulasne, C. Selecta Fungorum Carpologia; Friedrich Kfincksieck: Paris, France, 1861; pp. 218–243. [Google Scholar]
  51. Bonito, G.; Smith, M.E. General description of the black truffle (Tuber melanosporum) genome. Methods Mol. Biol. 2012, 862, 47–58. [Google Scholar]
  52. Bond, J.E.; Stockman, A.K. An integrative method for delimiting cohesion species: Finding the population-species interface in a group of Californian trapdoor spiders with extreme genetic divergence and geographic structuring. Syst. Biol. 2008, 57, 628–646. [Google Scholar] [CrossRef] [PubMed]
  53. Monsen-Collar, K.J.; Dolcemascolo, P. Using molecular techniques to answer ecological questions. Nat. Educ. Knowl. 2010, 3, 1. [Google Scholar]
  54. Belfiori, B.; Riccioni, C.; Paolocci, F.; Rubini, A. Mating type locus of Chinese black truffles reveals heterothallism and the presence of cryptic species within the T. indicum species complex. PLoS ONE 2013, 8, e82353. [Google Scholar] [CrossRef]
  55. Hosford, D.; Pilz, D.; Molina, R. Ecology and Manageent of Commercially Harvested True Morels in the Pacific Northwest, USA: A Review. For. Ecol. Manag. 1997, 4, 145–161. [Google Scholar]
Figure 1. Randomized Axelerated Maximum Likelihood (RAxML) tree based on the sequences ITS, LSU, tef1-α, and rpb2 of T. yunnanense, T. melanoumbilicatum, T. microexcavatum, and related species, with C. sichuanensis as the outgroup. Bootstrap (BS) values obtained from maximum likelihood (ML) analysis (≥70%) and posterior probabilities (PP) from Bayesian inference (≥0.90) are indicated above or below the branches at the nodes. Newly obtained sequences are highlighted in bold font.
Figure 1. Randomized Axelerated Maximum Likelihood (RAxML) tree based on the sequences ITS, LSU, tef1-α, and rpb2 of T. yunnanense, T. melanoumbilicatum, T. microexcavatum, and related species, with C. sichuanensis as the outgroup. Bootstrap (BS) values obtained from maximum likelihood (ML) analysis (≥70%) and posterior probabilities (PP) from Bayesian inference (≥0.90) are indicated above or below the branches at the nodes. Newly obtained sequences are highlighted in bold font.
Jof 10 00640 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, R.; Dong, G.; Li, Y.; Wang, R.; Yang, S.; Yuan, J.; Xie, X.; Shi, X.; Yu, J.; Pérez-Moreno, J.; et al. Three New Truffle Species (Tuber, Tuberaceae, Pezizales, and Ascomycota) from Yunnan, China, and Multigen Phylogenetic Arrangement within the Melanosporum Group. J. Fungi 2024, 10, 640. https://doi.org/10.3390/jof10090640

AMA Style

Wang R, Dong G, Li Y, Wang R, Yang S, Yuan J, Xie X, Shi X, Yu J, Pérez-Moreno J, et al. Three New Truffle Species (Tuber, Tuberaceae, Pezizales, and Ascomycota) from Yunnan, China, and Multigen Phylogenetic Arrangement within the Melanosporum Group. Journal of Fungi. 2024; 10(9):640. https://doi.org/10.3390/jof10090640

Chicago/Turabian Style

Wang, Rui, Gangqiang Dong, Yupin Li, Ruixue Wang, Shimei Yang, Jing Yuan, Xuedan Xie, Xiaofei Shi, Juanbing Yu, Jesús Pérez-Moreno, and et al. 2024. "Three New Truffle Species (Tuber, Tuberaceae, Pezizales, and Ascomycota) from Yunnan, China, and Multigen Phylogenetic Arrangement within the Melanosporum Group" Journal of Fungi 10, no. 9: 640. https://doi.org/10.3390/jof10090640

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop